A “.mmu-u . '0 31516.? $ .1 an}. .Pfln < .115- .4 1.5:! in, I , . 5....“ r. $5.; use.» fihnfiua .hbilfisfls .! Z. In}. 3):.- is. . nu. amusu find gnu-411d .hérrm R... 3 a . Eu K. 1...; W LIBRARY ‘2 Michigan State W University This is to certify that the dissertation entitled ANALYSIS OF MICROBIAL COMMUNITIES IN A CONTAMINATED AQUIFER UNDERGOING URANIUM BIOREMEDIATION presented by ERICK CARDENAS POIRE has been accepted towards fulfillment of the requirements for the Ph.D. degree in Crop and Soil Sciences - Environmental Toxicology Majy Professor’s §Ignature Date MSU is an Affirmative Action/Equal Opportunity Employer 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 5/08 K:IProj/Aoc&Pres/ClRC/Dale0ue indd ANALYSIS OF MICROBIAL COMMUNITIES IN A CONTAMINATED AQUIFER UNDERGOING URANIUM BIOREMEDIATION By Erick Cardenas Poire A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Crop and Soil Sciences — Environmental Toxicology 2009 ABSTRACT ANALYSIS OF MICROBIAL COMMUNITIES IN A CONTAMINATED AQUIFER UNDERGOING URANIUM BIOREMEDIATION By Erick Cardenas Poire Contamination by metal and radionuclides is a problem for the Y-12 National Security Complex in Oak Ridge, Tennessee, and at other Department of Energy (DOE) facilities and waste sites. In contrast to organic compounds, metals cannot be degraded so their bioremediation may be possible by controlling their bioavailability. Iron-reducing bacteria such as Geobacter and sulfate-reducing bacteria such as Desulfovibrio, among others, are known to utilize metals and radionuclides as terminal electron acceptors. Uranium (VI) reduction decreases its solubility therefore its movement in water can be controlled. In this work, microbial communities of a uranium and nitrate contaminated aquifer undergoing bioreduction, and located at the Field Research Center (FRC) Area 3 at Oak Ridge, Tennessee, were analyzed through comparative 16S rRNA gene analysis using both traditional and pyrosequencing of clone libraries. Clone libraries of this gene revealed sequences belonging to genera known to contain uranium reducers and whose closest relatives are known to reduce U(VI) and utilize either ethanol, the carbon source injected as electron donor; acetate, a by-product of ethanol oxidation; or methanol, an impurity of the injected ethanol solution. The genera detected in clone libraries known to reduce uranium included Desulfovibrio, Geobacter, Acidovorax, Anaeromyxobacter, and Desulfosporosinus, among others. Hydrologic characteristics helped to explain the higher levels of activity of sulfate reducers (SRB), iron reducers (FRB), and nitrate reducers in certain zones of the site. Pyrosequencing, a novel tool for microbial ecology studies, was used in the analysis of the microbial communities of the treatment location in the PRC Area 3. Differences in community structure along the bioremediation zone were related mostly to chemical oxygen demand, an indication on how much of the stimulatory carbon source reached a sampled well. Sequences belonging to genera with U(VI)-reducing ability were detected throughout the Area 3. The genera detected were the same as in the clone libraries but the increased sampling depth revealed previously undetected Shewanella and Deinococcus sequences. Diversity-based clustering showed four clusters of sequences that were related to their bioactivity (SRB, FRB, and denitrifiers levels). Indicator species analysis in addition to the clustering showed several groups that were significantly higher in abundance in the most active uranium reducing zones. The most significant indicator species were related to Desulfovibrio, Desulfosporosinus, and Anaeromyxobacter. Based on indicator species analysis, patterns of abundance, and metabolic ability of their closest relatives, sulfate-reducing bacteria, especially Desulfovibrio, appear to play the most important role in uranium reduction at this location at the time of the sampling. This thesis is dedicated to my family for their support, To my grandparents, I am sure you are proud To Eliana, my wife, who started me with the idea of the Ph.D. and to Luciana, my daughter, for the inspiration to finish it iv ACKNOWLEDGEMENTS I would like to thank Mary Beth Leigh for her mentoring during my early years at the lab. Carlos Rodriguez-Minguela, Claribel Cruz-Garcia and Veronica Gruntzig helped me adapting to the lab, to the sometimes-cold Michigan, and to the research world, I will never forget that. I would like to thank the Ribosomal Database Project team with whom I worked for over a year. They really amazed me with the hard work they do and taught me much of the comparative 16S rRNA analysis I know. My thanks to Jim Cole, its director, one of the brightest people I have ever met. Nowadays, research in environmental microbiology is a multidisciplinary team effort, and this work was not an exception. This thesis would have not been possible without the contribution of scientists at the Department of Energy Field Research Center, Wei-min Wu and Craig Criddle from Stanford University, and Jizhong Zhou from Oklahoma University. I would also like to thank my labmates for providing a stimulating and fun work environment. This research was supported by the US Department of Energy under grants DE- FG02-97ER62469 and DE—FG02-99ER62848, and by the US National Institute of Environmental Health Sciences grant 5P42 ESOO4911-18/19. PREFACE Parts of the work for this thesis (chapters two and three) are part of a team effort between Michigan State University, Stanford University, and the Field Research Center (FRC) at Oak Ridge National Laboratory. Stanford University designed and managed the groundwater treatment system used at the Field Research Center Area 3 and mentioned in this work. This group also characterized the hydrology of the site. Wei-min Wu, Craig Criddle, J ian Luo, Matthew Ginder-Vogel, and Peter Kitanidis formed the Stanford team. Oak Ridge National Laboratory provided us the site, performed biological and chemical analysis, and retrieved the sediment samples used for microbial community analysis. Jizhong Zhou, Terry Gentry, Jack Carley, Sue Carroll, David Watson, Baohua Gu, and Phil Jardine were part of the Oak Ridge team. Christopher Hemme, also from Oak Ridge, provided the DNA I used to construct one of our gene libraries, and helped us with useful discussion with his preliminary data from a metagenome constructed for that sample. Michigan State University was in charge of the microbial community analysis. Mary Beth Leigh, Terence Marsh, James Tiedje and I were part of the team. Tony Gaca helped with the construction of the clone libraries as well as extracting DNA from sediment samples. Finally the staff from the Ribosomal Database Project (RDP) provided the website used for the pyrosequencing analysis, technical support, and custom java scripts vi used for this work. Jim Cole, Benli Chai, Qiong Wang, Siddique Kulam-Syed-Mohideen, and Ryan Farris were valuable members of the RDP for this project. A portion of this work was previously published in here (1) . This part is not included in this thesis. vii REFERENCES Wu, W. M., J. Carley, J. Luo, M. Ginder-Volgel, E. Cardenas, M. B. Leigh, C. Hwang, S. D. Kelly, C. Ruan, L. Wu, J. van Nostand, T. Gentry, K. Lowe, T. Melhoume, S. Carroll, M. W. Fields, B. Gu, D. Watson, K. M. Kemner, T. L. Marsh, J. M. Tiedje, J. Zhou, P. K. Fendorf, P. K. Kitanidis, P. M. Jardine, and C. S. Criddle. 2007. In situ bioremediation of uranium (V I) to submicromolecular levels and reoxidation by dissolved oxygen. Environ. Sci. Technol. 41:5716-5723. viii TABLE OF CONTENTS LIST OF TABLES xii LIST OF FIGURES xiv CHAPTER ONE Introduction ....................................................................................................................... 1 Physicochemical characteristic of uranium and its applications ....................................... l Uranium contamination .................................................................................................... 2 Uranium reduction as bioremediation strategy ................................................................. 3 Microbial communities involved in uranium bioremediation .......................................... 7 References ......................................................................................................................... 9 CHAPTER TWO Microbial communities in contaminated sediments, associated with bioremediation of uranium to submicromolar levels ...................................................................................... 14 Abstract ........................................................................................................................... 15 Introduction .................................................................................................................... 1 6 Materials and methods ................................................................................................... 19 Site description and bioremediation test ...................................................................... 19 Sampling for community analysis ............................................................................... 21 Microbial enumeration ................................................................................................. 24 Tracer test ..................................................................................................................... 25 DNA extraction and community analysis .................................................................... 25 Chemical and analytical methods ................................................................................ 28 Results ............................................................................................................................. 28 In situ biostimulation test ............................................................................................. 28 Groundwater flow pathway .......................................................................................... 31 Microbial enumeration ................................................................................................. 34 Sequence analysis ........................................................................................................ 35 Microbial community structure and major groups detected ........................................ 36 Diversity analysis ......................................................................................................... 42 Discussion ....................................................................................................................... 44 1. Remediation of uranium and metals in the subsurface ............................................ 44 2. Microbial communities ............................................................................................ 44 (2a) Major trophic groups detected ........................................................................... 44 (2b) Carbon source as selective agent and the community of ethanol degraders ..... 45 (2c) Other C sources and contaminants .................................................................... 46 (2d) Putative genera involved in U(VI) reduction .................................................... 48 ix 3. Dynamics of the community .................................................................................... 50 (3a) Patterns of diversity detected ............................................................................. 51 (3b) Emergence of previously undetected populations ............................................. 52 Acknowledgements ......................................................................................................... 54 References ....................................................................................................................... 55 CHAPTER THREE Massive parallel sequencing of microbial communities in a uranium remediation site ....62 Introduction .................................................................................................................... 62 Materials and methods ................................................................................................... 66 Site description and sampling ...................................................................................... 66 Chemical and analytical methods ................................................................................ 67 DNA extraction and direct sequencing ........................................................................ 68 Pyrosequencing analysis .............................................................................................. 69 Community analysis ..................................................................................................... 70 Multivariate analysis .................................................................................................... 71 Results ............................................................................................................................. 72 Chemical and hydrological characterization ................................................................ 72 Correlation of environmental variables ....................................................................... 78 Pyrosequencing results ................................................................................................. 82 Hierarchical clustering of samples ............................................................................... 86 Microbial community structure .................................................................................... 88 Presence of genera with members known to reduce uranium ...................................... 94 Indicator species ........................................................................................................... 98 Ordination of samples by principal coordinates analysis .......................................... 106 Discussion ..................................................................................................................... 109 Effect of hydraulic control on microbial communities of the treatment area ............ 109 Fate of electron donor injected and populations that responded to the treatment ..... 111 Uranium reduction and its correlation with microbial groups ................................... 113 Patterns of diversity ................................................................................................... 115 Conclusions ................................................................................................................... 1 16 Acknowledgements ....................................................................................................... 1 17 References ..................................................................................................................... l 18 CHAPTER FOUR New tools for discovering and characterizing microbial diversity .................................. 125 Summary ....................................................................................................................... 126 Introduction ................................................................................................................... 126 Impact of new sequencing technologies in microbial ecology ..................................... 128 A new look at metagenomics ........................................................................................ 131 Array-based approaches for studying microbial diversity ............................................ 132 Microfluidics, manipulations at micro(be) scale .......................................................... 133 The formerly unculturable — New approaches for isolation ......................................... 133 Metatranscriptomics and metaproteomics, fields in their infancy ................................ 135 Future challenges and directions ................................................................................... 138 Reference annotations ................................................................................................... 138 Acknowledgements ....................................................................................................... 142 References ..................................................................................................................... 143 CHAPTER FIVE Conclusions ...................................................................................................................... 149 APPENDIX ..................................................................................................................... 152 Appendix references ..................................................................................................... 153 xi LIST OF TABLES Table 2.1. Chemical properties of groundwater in wells of treatment area before and after remediation and in control well F W1 06“ ........................................................................... 23 Table 2.2. MPN for three major trophic groups in sediments of treatment zone (day 774) in comparison with groundwater from control well FW106 ............................................. 35 Table 2.3 Relative contribution of known uranium reducer genera in clone libraries from sediment samples in Area 3wells" ...................................................................................... 41 Table 2.4. Diversity indices in treated and untreated (FW 106) areas .............................. 42 Table 3.1. Hydrological characterization of the DOE FRC Area 3 wells ......................... 74 Table 3.2. Chemical composition of groundwater from wells from DOE FRC Area 3....75 Table 3.3. Chemical composition of sediments from DOE F RC Area 3 .......................... 77 Table 3.4. Correlation of chemical characteristics of groundwater, sediments and hydrology of the DOE FRC Area 3 wells .......................................................................... 80 Table 3.5. Sequences per sample and number of groups detected at different levels of taxonomy ............................................................................................................................ 83 Table 3.6. Detection of phyla poorly represented in 16S rRNA databases in libraries from FRC area 3 samples ........................................................................................................... 84 Table 3.7. Top 15 indicator species for sample cluster A, their classification and closest isolated relative ................................................................................................................ 102 xii Table 3.8. Top 15 indicator species for sample cluster B, their classification and closest isolated relative ................................................................................................................ 103 Table 3.9. Top 15 indicator species for sample cluster C, their classification and closest isolated relative ................................................................................................................ 104 Table 3.10. Top 15 indicator species for sample cluster D, their classification and closest isolated relative ................................................................................................................ 105 Table 3.11. Environmental factor that explain distribution of groups along the samples .......................................................................................................................................... 109 Table 4.1. Current available sequencing technologies and their features ....................... 129 Table 4.2. Selected resources available for metagenomics, meta transcriptomics and metaproteomics ................................................................................................................ 1 3 7 xiii LIST OF FIGURES Figure 1.1. Mechanism for biological mediated Uranium (VI) dissimilatory reduction .............................................................................................................................. 5 Figure 2.1. Map of the Area 3 treatment zone depicting the location of the samples wells and the control well. .......................................................................................................... 21 Figure 2.2. Typical biostimulation of U(VI) reduction by injection ethanol to the subsurface (days 704 to 714) ............................................................................................. 29 Figure 2.3. A tracer test performed with Br' on days 801 to 801 shows the hydraulic connection between injection well F W104, extraction well FW026, and MLS wells and biodegradation of ethanol .................................................................................................. 33 Figure 2.4. Rarefaction curves of the 16S rRNA gene libraries constructred .................. 36 Figure 2.5. Microbial composition of the clone libraries based on the RDP Classifier....38 Figure 2.6. Neighbor-joining tree showing the relationship of selected representatives (shown in bold) from groups similar to known U-reducing bacteria ................................. 40 Figure 2.7. Clustering of samples according to diversity patterns .................................... 43 Figure 3.1. Scheme of the DOE Field Research Center Area 3 treatment system ............ 67 Figure 3.2. Correlation plot of environmental variables measured for 15 locations at DOE FRC Area 3 ............................................................................................................... 79 Figure 3.3. Rarefaction curves for F RC Area 3 samples .................................................. 85 xiv Figure 3.4. Richness estimations for samples from DOE FRC Area 3 as influenced by evenness and sampling effort ............................................................................................. 86 Figure 3.5. Clustering of samples by abundance-based Sorensen index .......................... 87 Figure 3.6. Relative contribution of most abundant phyla to the microbial community from 15 samples from DOE FRC Area 3 ........................................................................... 89 Figure 3.7. Relative contribution of the five most abundant phyla present at samples from the FRC Area 3 .......................................................................................................... 91 Figure 3.8. Class composition of the Proteobacteria from FRC Area 3 samples ............ 92 Figure 3.9. Relative abundance of Deltaproteobacteria orders in FRC Area 3 samples .94 Figure 3.10. Relative contribution of most common uranium reducing genera found in the samples from DOE FRC Area 3 .................................................................................. 95 Figure 3.1 1. Location of maximum abundance for uranium reducing genera along the connectivity gradient at FRC Area 3 ................................................................................. 97 Figure 3.12. Principal coordinates ordination on abundance-based Sorensen distances for samples from the FRC Area 3 .......................................................................................... 107 Figure 4.1. Novel methods to discover and characterize microbial diversity ................. 128 XV INTRODUCTION Uranium Contamination and Strategies for its Bioremediation 1. Physicochemical characteristic of uranium and its applications Uranium (U) is the natural element with the highest atomic number, 92 (protons in its nucleus), and one of the densest (19.1 g-mL-l). Uranium occurs naturally as a mixture of several isotopes, mostly 238U (99.28%) followed by 235U (0.71%) and 234U (0.0054%). The half-life for these isotopes are 4.47 x 109 years, 7.0 x106 years and 2.55 x105 years, respectively. Because of its size, uranium’s nucleus is unstable and decays naturally 1n a known serles of elements emitting alpha particles. Uranium ranks 42n in abundance in the Earth crust, and it is distributed worldwide and even present in the oceans. Uranium is exploited mostly in mining operations in Canada, Australia and Russia. The main applications of uranium exploit two of its features: 1) its high density and 2) the ability of its nucleus to be broken. In the military world, uranium is used for armor plating and high penetration bullets. By being naturally fissible, it can be used for energy production in nuclear reactors, and for military purposes in bombs. Other uses include painting, radioisotope dating and transmission electron microscopy. In order to make uranium usable for energy production, uranium has to be transformed to increase the proportion of 235U isotope relative to 23 8U. This process generates depleted uranium [typically less than 0.2 % of 235U, low enriched uranium (0.2 235 . . . 235 . . to 3% of U) and ennched uran1um(typ1cally 3 to 5% U)]. Uraruum enrichment . . . . . 235 creates a high proport1on of depleted uranium dependlng on the natural U . . . . 235 concentratlon and the targeted final concentration. e.g., enrichlng for 3.6% U for fuel in a nuclear plant will need approximately 8 kg of natural uranium per kg of enriched uranium. 2. Uranium contamination Uranium it is a contaminant in facilities associated with uranium mining and enrichment. Uranium mining is the most frequent industry that has uranium contamination problems. In these sites, human exposure to uranium is due to inhalation and skin contact and the risks relate primarily to the chemical toxicity of U (as a heavy metal) and secondly to its radioactivity. However, 222Ra, a byproduct of U decay, is associated with lung cancer (7). Depleted uranium contamination is associated with uranium enrichment facilities as well as to the military industry where uranium is used in ammunition or heavy armor plating. The most frequent exposure route for depleted U is inhalation followed by ingestion through water (4). Acute exposure to uranium leads to renal toxicity, while chronic exposure can problems in lungs, the central nervous system, intestines, and kidneys (3, 7). Depleted uranium is 60% less radioactive than naturally occurring uranium, and its toxicity is more related to its heavy metal characteristics. Metals and radionuclides are frequent contaminants in the US Department of Energy facilities and waste sites (31). One of these sites, the Y-12 National Security Complex in Oak Ridge, Tennessee, USA, has groundwater contaminated with depleted uranium as well as nitrate, sulfate, and heavy metals. The main sources of contamination are four ponds (S-3 ponds) where the Y-l2 uranium enrichment plant stored its wastes for over 32 years. These unlined ponds had a 9.5 million liter capacity each and receive mainly acidic uranium nitrate as well as solvents, and wastes from other DOE facilities. In 1983, the ponds were neutralized with limestone, and their nitrate was removed by denitrification. Treated water was pumped out of the site and the ponds with their contaminated sediments were capped with a Resource and Conservation Recovery Act (RCRA) cover, and paved with asphalt to construct a parking lot (5). Nevertheless, the areas surrounding the ponds were already contaminated, and intrusion of groundwater to the ponds promoted the movement of soluble uranium from them (30, 40). Groundwater at this site is highly acidic (pH 3.5) and contains high levels of inorganic contaminants such as nitrate and sulfate; radionuclides such as uranium and technetium; and solvents such as acetone, trichloroethene and tetrachloroethene (41). This site is now used to develop and study uranium remediation strategies. 3. Uranium reduction as bioremediation strategy One of the earliest definitions of bioremediation was “ the utilization of microorganisms for the destruction of chemical pollutants” (1). Though this definition works fine for the degradation of organic compounds, it does not apply for metals as they cannot be degraded. However, bioremediation of metals is possible by altering their bioavailability. This can be done by sequestering metals into organic matrices, by assimilating the metals into the cells, by changing the solubility of the metals via oxidation/reduction or complextion with organic compounds, and by modifying the chemicals associated with uranium in situ (e. g., dissolving insoluble phosphate-uranium precipitates) (17, 23, 36). From these, one of the most promising strategies is dissimilatory metal reduction, a process by which microorganism use metals (e.g., Fe(III), Mn (IV), and Cr (VI)) and radionuclides (e.g., U(VI) and Tc(VII)) as terminal electron acceptors (18). The electrons can come from the breakup of organic compounds and, in some special cases, directly from electrodes (10). Electrons are then transferred from the organisms to the metals through direct contact such in Geobacter spp. (26); through electron shuttles such in Shewanella spp. (22); or through reduced products such as hydrogen sulfide (13), reduced humic acids (16, 27), nitrogen oxides (32), and reduced Fe(II) such as green rust(29) (Figure 1.1). While oxygen is usually the favored electron acceptor for microbial communities, these metals can be reduced in anaerobic conditions. Changes in the oxidation state can trigger changes in solubility in both directions. Reduction can yield more soluble products (e. g., F e(III) to Fe(II)) or more insoluble ones (Tc(VI) to Tc(V) and Tc(IV)). In the case of uranium the reduction is desirable because reduced uranium (U(IV)) is less water soluble; thus, uranium movement through water can be controlled if populations capable of reducing uranium are stimulated. Stimulation of uranium reducers can be achieved mainly by injecting an electron donor and/or carbon source into the aquifer and providing the right conditions for microbial activity in terms of pH, nutrients and minimizing competing electron acceptors such as oxygen and nitrate. More recently uranium reduction has been achieved through direct electron transfer by inserting electrodes, and electrodes coated with iron reducing bacteria (F RB) into aquifers (10). @ Organic electron donor Electrons in EICCIIOHS out Ii I 4 Electron” Figure 1.1. Mechanism for biological mediated Uranium (VI) dissimilatory reduction. Electron can be donated from electrodes directly (1) or from the decomposition of organic compounds (2). Form bacterial cells, electron can then be transferred to uranium by direct contact (3), or through electron shuttles of reduced compound such as humic acids or hydrogen sulfide (4). Dissimilatory uranium reduction was first demonstrated for the FRB Geobacter metallireducens, and the sulfate-reducing bacteria (SRB) Shewanella putrefaciens, both of them belonging to the Delta class of the Proteobacteria (20). Since then, the ability to reduce uranium has been detected in at least 18 other genera (3 8) whose affiliations range from F irmicutes, Deinococcus, Thermus, Actinobacteria, Beta-, Delta- and Gamma- Proteobacteria as well as Archaea (38). In terms of microbial diversity the ability to reduce uranium seems widespread through the tree of life; however, for management purposes the widespread distribution can create challenges in the stimulation of the desired populations. In addition, is not clear what differences in reduction efficiency exist among the groups, and how this efficiency relates to environmental conditions such as pH, oxygen concentrations, and electron donor availability among others. Uranium reduction as a way of controlling its movement has been tested in pure cultures of F RB and SRB (19, 21), batch cultures in serum bottles (8, 12, 43), microcosms (28), columns (11, 39) and in pilot scale projects (2, 15, 37, 41). Field studies have yielded diverse results depending on the carbon source used, the delivery method, and the characteristics of the site. In a former uranium mine, microcosm studies using contaminated sediments found the SRB Desulfosporosinus, and the fermenter Clostridium associated with uranium reduction (35). The medium used contained ethanol, lactate, acetate, benzoic acids, glucose, yeast extract and peptones (35). At a former uranitun ore processing facility at Rifle, Colorado, in situ stimulation via acetate injection promoted U reduction by FRB Geobacter, but in a second phase, SRB not capable of U reduction outgrew Geobacter (2, 37). Recently it was reported that uranium removal at Rifle continued even after acetate amendments stopped (24), and that this removal was associated with adsorption of uranium to the biomass of a F irmicutes- related bacteria that likely grew on dead cells from previously stimulated populations (24). At the Field Research Center at Oak Ridge, Tennessee, a pilot scale project showed associations between uranium reduction with SRB Desulfovibrio (14, 42, 43), F RB Geobacter (14, 42), FRB Anaeromyxobacter (6), SRB Desulfosporosinus (6, 14), and denitrifying Acidovorax (28). 4. Microbial communities involved in uranium bioremediation In order to study the microbial communities involved in the uranium reduction in situ, culture independent methods have been predominantly used. The preferred methods to track specific populations involved the use of the 16S rRNA gene in combination with terminal restriction fragment length polymorphism (TRFLP), restriction fragment length polymorphisms (RFLP), denaturing gradient gel electrophoresis (DGGE), and clone libraries. Restriction enzyme-based methods are simpler and cheaper and can provide a profile for the whole community that can be later consistently compared. However, it is hard to associate peaks or bands with specific populations unless a clone library for the same site is constructed. DGGE provides bands that can be tracked over samples and later sequenced for their phylogenetic information. However, the variability between and within gel gradients as well as a limitation in how many bands can be resolved (usually 20 to 40), can make this method challenging to analyze (25). Finally, clone libraries provide the most sequence information (usually 500 to 800 bases per clone). This information can be use for phylogeny inference and community reconstruction. However, library construction is time consuming, expensive, limited in scope, and in some occasions bias can be introduced in the cloning step (34). All PCR-based methods such as, TRF LP, DGGE and clone libraries can be biased due to primer coverage, differences in copy number of a gene in a genome, and extraction of DNA (9). One recent alternative that is both high throughput and provide sequence information is massive parallel sequencing. By sequencing hypervariable regions of the 16S rRNA gene, thousand of reads can be generated per sample, and the structure of the microbial community reconstructed by comparing the sequences to known databases (33). In this thesis I use comparative 16S rRNA gene studies of both traditional and pyrosequencing clone libraries, to study the diversity and structure of the communities associated with uranium bioremediation at the Oak Ridge National Laboratory Field Research Center, Area 3. The focal research questions are 1) what are the main populations associated with U reduction, and 2) how does their abundance, and presumably the features favoring their selection, relate to the geochemical conditions present at this site? REFERENCES Alexander, M. 1999. Biodegradation and Bioremediation. Academic Press, Incorporated, San Diego, California. Anderson, R. T., H. A. Vrionis, I. Ortiz-Bernad, C. T. Resch, P. E. Long, R. Dayvault, K. Karp, S. Marutzky, D. R. Metzler, A. Peacock, D. C. White, M. Lowe, and D. R. Lovley. 2003. Stimulating the in situ activity of Geobacter species to remove uranium from the groundwater of a uranium-contaminated aquifer. Appl Environ Microbiol 69:5884—91. Berradi, H., J. M. Bertho, N. Dudoignon, A. Mazur, L. Grandcolas, C. Baudelin, S. Grison, P. Voisin, P. Gourmelon, and I. Dublineau. 2008. Renal anemia induced by chronic ingestion of depleted uranium in rats. Toxicol Sci 103:397-408. Bleise, A., P. R. Danesi, and W. Burkart. 2003. Properties, use and health effects of depleted uranium (DU): a general overview. J Environ Radioact 64:93- 112. Brooks, S. C. 2001. Waste Characteristics of the Former S-3 ponds and outline of uranium chemistry relevant to NABIR Field Research Center studies ORNL/TM-2001/27. NABIR Field Research Center. Cardenas, E., W. M. Wu, M. B. Leigh, J. Carley, S. Carroll, T. Gentry, J. Luo, D. Watson, B. Gu, M. Ginder-Vogel, P. K. Kitanidis, P. M. Jardine, J. Zhou, C. S. Criddle, T. L. Marsh, and J. M. Tiedje. 2008. Microbial communities in contaminated sediments, associated with bioremediation of uranium to submicromolar levels. Appl Environ Microbiol 74:3718-29. Department of Protection of the Human Environment, W. H. O. 2001. Depleted Uranium: Sources, Exposure and Health Effects WHO/SDE/PHE/Ol .1. Finneran, K. T., M. E. Housewright, and D. R. Lovley. 2002. Multiple influences of nitrate on uranium solubility during bioremediation of uranium- contaminated subsurface sediments. Environ. Microbiol. 4:510-6. 10. ll. 12. 13. 14. 15. 16. 17. 18. Friedrich v. Wintzingerode, U. B. G., Erko Stackebrandt,. 1997. Determination of microbial diversity in environmental samples: pitfalls of PCR- based rRNA analysis. FEMS Microbiology Reviews 21:213-229. Gregory, K. B., and D. R. Lovley. 2005. Remediation and recovery of uranium from contaminated subsurface environments with electrodes. Environ Sci Technol 39:8943-7. Cu, B., W. M. Wu, M. A. Ginder-Vogel, H. Yan, M. W. Fields, J. Zhou, S. Fendorf, C. S. Criddle, and P. M. Jardine. 2005. Bioreduction of uranium in a contaminated soil column. Environ Sci Technol 39:4841-7. Holmes, D. E., K. T. Finneran, R. A. O'Neil, and D. R. Lovley. 2002. Enrichment of members of the family Geobacteraceae associated with stimulation of dissimilatory metal reduction in uranium-contaminated aquifer sediments. Appl. Environ. Microbiol. 68:2300-6. Hua, B., H. Xu, J. Terry, and B. Deng. 2006. Kinetics of uranium(VI) reduction by hydrogen sulfide in anoxic aqueous systems. Environ Sci Technol 40:4666-71. Hwang, C., W. Wu, T. J. Gentry, J. Carley, G. A. Corbin, S. L. Carroll, D. B. Watson, P. M. Jardine, J. Zhou, C. S. Criddle, and M. W. Fields. 2008. Bacterial community succession during in situ uranium bioremediation: spatial similarities along controlled flow paths. ISME J. Istok, J. D., J. M. Senko, L. R. Krumholz, D. Watson, M. A. Bogle, A. Peacock, Y. J. Chang, and D. C. White. 2004. In situ bioreduction of technetium and uranium in a nitrate-contaminated aquifer. Environ Sci Technol 38:468-75. Jang, J. H., B. A. Dempsey, and W. D. Burgos. 2008. Reduction of U(VI) by Fe(II) in the presence of hydrous ferric oxide and hematite: effects of solid transformation, surface coverage, and hurrric acid. Water Res 42:2269-77. Knox, A. S., R. L. Brigmon, D. 1. Kaplan, and M. H. Paller. 2008. Interactions among phosphate amendments, microbes and uranium mobility in contaminated sediments. Sci Total Environ 395:63-71. Lovley, D. R. 1993. Dissimilatory metal reduction. Annu Rev Microbiol 47:263- 90. 10 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. Lovley, D. R., and E. J. Phillips. 1992. Reduction of uranium by Desulfovibrio desulfuricans. Appl. Environ. Microbiol. 58:850-856. Lovley, D. R., E. J. P. Phillips, Y. A. Corby, and E. R. Landa. 1991. Microbial reduction of uranium. Nature 350:413-416. Lovley, D. R., E. E. Roden, E. J. P. Phillips, and J. C. Woodward. 1993. Enzymatic iron and uranium reduction by sulfate—reducing bacteria. Marine Geology 113. Marsili, E., D. B. Baron, I. D. Shikhare, D. Coursolle, J. A. Gralnick, and D. R. Bond. 2008. Shewanella secretes flavins that mediate extracellular electron transfer. Proc Natl Acad Sci U S A 105:3968-73. Merroun, M. L., and S. Selenska-Pobell. 2008. Bacterial interactions with uranium: An environmental perspective. J Contarn Hydrol 102:285-95. N'Guessan, A. L., H. A. Vrionis, C. T. Resch, P. E. Long, and D. R. Lovley. 2008. Sustained removal of uranium from contaminated groundwater following stimulation of dissimilatory metal reduction. Environ Sci Technol 42:2999-3004. Nakatsu, C. H. 2007. Soil Microbial Community Analysis Using Denaturing Gradient Gel Electrophoresis. Soil Sci Soc Am J 71:562-571. Nevin, K. P., and D. R. Lovley. 2000. Lack of production of electron-shuttling compounds or solubilization of Fe(III) during reduction of insoluble Fe(III) oxide by Geobacter metallireducens. Appl Environ Microbiol 66:2248-51. Nevin, K. P., and D. R. Lovley. 2000. Potential for Nonenzymatic Reduction of Fe(III) via Electron Shuttling in Subsurface Sediments. Environ. Sci. Technol. 34:2472-2478. Nyman, J. L., T. L. Marsh, M. A. Ginder-Vogel, M. Gentile, S. Fendorf, and C. Criddle. 2006. Heterogeneous response to biostimulation for U(VI) reduction in replicated sediment microcosms. Biodegradation V17:303-316. O'Loughlin, E. J., S. D. Kelly, R. E. Cook, R. Csencsits, and K. M. Kemner. 2003. Reduction of uranium(VI) by mixed iron(II)/iron(III) hydroxide (green rust): formation of U02 nanoparticles. Environ Sci Technol 37:721-7. 11 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. Phillips, D. H., D. B. Watson, Y. Rob, T. L. Mehlhorn, J. W. Moon, and P. M. Jardine. 2006. Distribution of uranium contamination in weathered fractured saprolite/shale and ground water. J Environ Qual 35: 1715-30. Riley, R. G., and J. M. Zachara. 1992. Chemical contaminants on DOE lands and selection of contaminant mixtures for subsurface science research. US. Department of Energy, Washington, DC. Senko, J. M., J. D. Istok, J. M. Suflita, and L. R. Krumholz. 2002. In-situ evidence for uranium immobilization and remobilization. Environ Sci Technol 36:1491-6. Sogin, M. L., H. G. Morrison, J. A. Huber, D. Mark Welcb, S. M. Huse, P. R. Neal, J. M. Arrieta, and G. J. Herndl. 2006. Microbial diversity in the deep sea and the underexplored. Proc Natl Acad Sci U S A 103:12115-20. Sorek, R., Y. Zhu, C. J. Creevey, M. P. Francine, P. Bork, and E. M. Rubin. 2007. Genome-wide experimental determination of barriers to horizontal gene transfer. Science 318: 1449-52. Suzuki, Y., S. D. Kelly, K. M. Kemner, and J. F. Banfield. 2003. Microbial populations stimulated for hexavalent uranium reduction in uranium mine sediment. Appl Environ Microbiol 69: 1337-46. Turick, C. E., A. S. Knox, C. L. Leverette, and Y. G. Kritzas. 2008. In situ uranium stabilization by microbial metabolites. J Environ Radioact 99:890-9. Vrionis, H. A., R. T. Anderson, I. Ortiz-Bernad, K. R. O'Neill, C. T. Resch, A. D. Peacock, R. Dayvault, D. C. White, P. E. Long, and D. R. Lovley. 2005. Microbiological and geochemical heterogeneity in an in situ uranium bioremediation field site. Appl Environ Microbiol 71:6308-18. Wall, J. D., and L. R. Krumholz. 2006. Uranium reduction. Annu Rev Microbiol 60: 149-66. Wan, J., T. K. Tokunaga, E. Brodie, Z. Wang, Z. Zheng, D. Herman, T. C. Hazen, M. K. Firestone, and S. R. Sutton. 2005. Reoxidation of bioreduced uranium under reducing conditions. Environ. Sci. Technol. 39:6162-9. l2 40. 41. 42. 43. Watson, D. B., W. E. Doll, T. J. Gamey, J. R. Sheeban, and P. M. Jardine. 2005. Plume and lithologic profiling with surface resistivity and seismic tomography. Ground Water 43: 169-77. Wu, W. M., J. Carley, M. Fienen, T. Mehlhorn, K. Lowe, J. Nyman, J. Luo, M. E. Gentile, R. Rajan, D. Wagner, R. F. Hickey, B. Gu, D. Watson, 0. A. Cirpka, P. K. Kitanidis, P. M. Jardine, and C. S. Criddle. 2006. Pilot-scale in situ bioremediation of uranium in a highly contaminated aquifer. 1. Conditioning of a treatment zone. Environ Sci Technol 40:3978-85. Wu, W. M., J. Carley, T. Gentry, M. A. Ginder-Vogel, M. Fienen, T. Mehlhorn, H. Yan, S. Carol], M. N. Pace, J. Nyman, J. Luo, M. E. Gentile, M. W. Fields, R. F. Hickey, B. Gu, D. Watson, 0. A. Cirpka, J. Zhou, S. F endorf, P. K. Kitanidis, P. M. Jardine, and C. S. Criddle. 2006. Pilot-scale in situ bioremediation of uranium in a highly contaminated aquifer. 2. Reduction of u(VI) and geochemical control of u(VI) bioavailability. Environ Sci Technol 40:3986-95. Wu, W. M., B. Gu, M. W. Fields, M. Gentile, Y.-K. Ku, H. Yan, S. Tiquias, T. Yan, J. Nyman, J. Zhou, P. M. Jardine, and C. S. Criddle. 2005 . Uranium (VI) Reduction by Denitrifying Biomass. Bioremediation 9:46-61. 13 CHAPTER TWO Cardenas, E., W. M. Wu, M. B. Leigh, J. Carley, S. Carroll, T. Gentry, J. Luo, D. Watson, B. Gu, M. Ginder-Vogel, P. K. Kitanidis, P. M. Jardine, J. Zhou, C. S. Criddle, T. L. Marsh, and J. M. Tiedje. 2008. Microbial communities in contaminated sediments, associated with bioremediation of uranium to submicromolar levels. Applied and Environmental Microbiology 74:3718-29. l4 Microbial Communities in Contaminated Sediments Associated with Bioremediation of Uranium to Submicromolar Levels ABSTRACT Microbial enumeration, 16S rRNA gene clone libraries and chemical analysis were used to evaluate the in situ biological reduction and immobilization of uranium (VI) in a long term experiment (more than two years) conducted at a highly uranium contaminated site (up to 60 mg/l and 800 mg/kg solids) of the US Department of Energy in Oak Ridge, TN. Bioreduction was achieved by conditioning groundwater above ground and then stimulating growth of denitrifying, Fe(III)-reducing, and sulfate-reducing bacteria in situ through weekly injection of ethanol to the subsurface. After nearly two years of intermittent injection of ethanol, aqueous U levels fell below the US EPA maximum contaminant level (MCL) for drinking water and groundwater (<30 ug/l or 0.126 pM). Sediment microbial communities from the treatment zone were compared with those from a control well without biostimulation. Most Probable Number estimations indicated that microorganisms implicated in bioremediation accumulated in the sediments of the treatment zone but were either absent or in very low numbers in an untreated control area. Organisms belonging to genera known to include U(Vl) reducers were detected, including Desulfovibrio, Geobacter, Anaeromyxobacter, Desulfosporosinus and Acidovorax spp. The predominant sulfate-reducing bacterial species were Desulfovibrio spp. while the iron reducers were represented by F erribacterium spp and Geothrix spp. Diversity-based clustering revealed differences between treated and untreated zones and also within 15 samples of the treated area. Spatial differences in community structure within the treatment zone were likely related to hydraulic pathway and electron donor metabolism during biostimulation. INTRODUCTION Metal and radionuclide are common groundwater contaminants at facilities and waste sites of the US Department of Energy (DOE) occurring in more than 50% of these locations (32). One of these sites, the former Y-12 National Security Complex at Oak Ridge, Tennessee, contains uranium in concentrations as high as 60 mg/l (252 pM) in groundwater and 800 mg/kg in sediments (48). To control the migration of the uranium, microbial reduction of U(VI) to sparingly soluble and immobile U(IV) has been proposed as a promising approach (1, 2, 12, 17, 22, 42, 48). Bioreduction of U(Vl) to U(IV) has been reported by pure and mixed cultures of iron(III)-reducers (FeRB) such as Geobacter spp., sulfate—reducers (SRB) such as Desulfovibrio, Desulfoporosz’nus, and Desulfotomaculum spp. and denitrifying bacteria such as Acidovorax spp. among others (45). In recent years, researchers have evaluated this approach through laboratory-scale experiments using batch serum bottles (10, 14, 51), microcosms (29), and sediment columns (11, 46). Reduction and immobilization of uranium in the mentioned laboratory experiments has been observed and confirmed by X-ray near-edge absorption spectroscopy (XANES) analysis (11, 29, 40, 41, 46). Geobacter spp. and Geothrix spp. were found to be associated with Fe(III) and U(VI) reduction in field conditions (5, 14, 16 46, 50). Control of the microbial community structure may be one of key issues for the long-term reduction and stabilization of uranium in situ (1). While bench-scale tests provide valuable information about the feasibility of bioreduction, they cannot replicate all of the heterogeneity and complexity of the subsurface. Hydrogeology is a factor that can contribute to the complexity of field treatments by creating gradients of resources in situ, affecting the microbial diversity and thus, potentially the remediation process. Little is known about the effect of groundwater flow on microbial community diversity during and after bioremediation, even when the kinetics of U(VI) reduction should be dependent on the microbial community and geochemical conditions. The effect that hydrogeology has on microbial diversity should be addressed. Pilot studies are the next step in demonstrating the feasibility of the in situ uranium remediation approach (12, 14, 17, 37, 48, 49). DOB does not provide specific uranium target levels, but US EPA regulates the maximum contaminant level (MCL) for drinking and groundwater at 30 ug/l (0.126 uM). At the Field Research Center (F RC) site of the Environmental Remediation Sciences Program (ERSP) of the DOE, Oak Ridge, TN, test facilities have been constructed for the remediation of uranium contaminated groundwater and sediments using various approaches (1, 17, 49). Prior to the tests, the microbial community structure was characterized by 16S rRNA gene clone libraries made from groundwater samples in Areas 1 and 3 and the majority of the sequences (>73% in Area 1 and >65% in Area 3) were found to be Proteobacteria belonging to the genera Azoarcus and Pseudomonas (9). Microbial community analyses of Areas 1 and 2 identified Geobacter spp. and Anaeromyxobacter spp. among others as the potential uranium reducers (28). In these areas of moderate uranium contamination (0 to 5.8 uM), l7 removal of nitrate and aqueous U(VI) was stimulated by the injection of ethanol or glucose to the subsurface using a push-pull approach (17). In the extremely contaminated Area 3, a long-term (> 2 years) bioremediation test has been performed. This area is located near former S-3 Ponds and contains high levels of nitrate (up to 200 mM) and U (up to 250 uM), aluminum(12-13mM) and calcium (22-25mM) (48, 49). Prior to remediation, the microbial levels in groundwater of this area were extremely low probably due to low pH (3.6) and high levels of nitric acid. A combination of remediation. approaches was used to remove U(VI) reduction inhibitors (nitrate and Ca) and condition the area for metal reduction by raising the pH and providing a carbon source. Using these approaches, low U(VI) concentrations below MCL were achieved and a new microbial community capable of uranium reduction was established (48-50). The objective of our study was to characterize the microbial community arising in a earlier study where successful U(VI) bioremediation was demonstrated (49). We used 16S rRNA genes sequence analysis to determine the structure of the bacterial consortia present in the treatment area where low uranium levels were achieved during biostimulation and in a control samples from outside the treated area. The community data were integrated with geochemical and hydraulic data to provide insight about environmental variables that profoundly influence the remediation process. We were able to identify key bacterial groups associated with successful reduction of U(VI) in the subsurface and correlate their spatial relationship with hydrogeology and geochemistry in the treatment zone. 18 MATERIAL AND METHODS Site description and bioremediation test. The bioremediation test was performed in Area 3 of the DOE F RC at the Y-12 National Security Complex, Oak Ridge, TN as previously reported (24, 48). The field system consisted of an outer groundwater recirculation loop (injection well FW024 and extraction well FW103) that isolates an inner groundwater loop (injection well FW104 and extraction well FW026) preventing penetration by highly contaminated groundwater from the source zone (Figure 2.1). The hydraulic control afforded by this system created a controlled in situ bioreactor. Reduction of U(VI) to U(IV) was accomplished through ethanol injection to the inner loop. Injection and extraction wells (FW104 and FW026) had a 10.2 cm diameter and a depth of 14.6 m below ground surface (bgs) with 2.5 m screened intervals between 11.28 and 13.77 m. Multilevel sampling (MLS) wells F W100, F W101 and FW102 were used to monitor hydrogeology and remediation performance. The MLS wells contained seven separate sampling tubes (diameter =1.9 cm) at different depths bgs. In this study, MLS wells FW101-2 (13.7 m bgs), FW101-3 (12.2 m bgs), FW102-2 (13.7 m bgs) and F W102-3 (12.2 m bgs) were selected for routine monitoring of remediation performance because of their hydraulic connection to FW104 (24, 48). The recirculation flow rates in the inner loop were 0.45 l/min (injection F W1 04, extraction F W026). The rates for the outer loop were 1.35 l/min (injection at FW024) and 0.45 l/rnin( extraction at FW103). Additional clean water was injected to F W024 at 0.7-0.9 l/min in order to minimize entry of ambient groundwater (48). This clean water was a mixture of tap water and groundwater treated by an aboveground system to remove nitrate via a bioreactor (48). The remediation test was started on August 23, 2003 by preconditioning the site 19 (day 0). During the initial 137 days, water was pumped from the subsurface, pH adjusted, treated to remove aluminum, calcium and nitrate (this by a denitrification bioreactor) and then re-injected. Calcium and nitrate was removed to avoid formation of stable Ca-U-CO3 products, U(IV) re-oxidation by nitrites, and nitrate competition as terminal electron acceptor. Ex situ treatment was used to avoid clogging by nitrogen gas, biomass (due to denitrification), and calcium and aluminum precipitates (due to pH adjustment) (48). As result, the pH increased from 3.6 to around 6.0, and nitrate and U(VI) concentrations decreased to around 0.5 mM and 5 uM, respectively (48). After the conditioning phase, ethanol was added as an electron donor to stimulate U(VI) bioreduction starting on January 7, 2004 (day 137) (49). An ethanol solution (industrial grade, containing ethanol 88.12%, methanol 4.65% and water 7.23%, w/w) with a Chemical Oxygen Demand (COD) to weight ratio of 2.1 was prepared in a storage tank with 6.9-9.8 g COD/l. This solution was normally injected at F W1 04 over a 48-hour period each week to give a COD 120-150 mg/l at FW104. 20 Outer loop \ FW102 lnnerloop\1 FW101 . FW103... / o FW024 FW105 / \ FWAHJZES III I:VV1(M4 FW100 Water flow 44' hfleters 0 1.5 3 6 FW106 0 Wells I Multisampling Wells Figure 2.1. Map of the Area 3 treatment zone depicting the location of the sampled wells and the control well. FW024: outer loop injection; F W1 04: inner loop injection; FW026: inner loop extraction; FW103: outer loop extraction well. Wells F W100, FW101 and FW102 are multilevel sampling wells. FW106: control well. Contamination source is approximately 20 m to the right. Sampling for community analysis. Sediment samples for microbial community analysis were collected on day 774 (October 5, 2005) from the inner loop injection FW104, extraction F W026, and two MLS wells at two depths F W101-2, F W101-3, FW102-2 and F W102-3. To collect the samples, a smooth PVC surge block (10 cm or 1.8 cm in diameter) was inserted into the well and then lifted up and down in a rapid plunging motion. This motion drew sediment from the soil matrix surrounding the well screens into the well. The surge blocks were rinsed with groundwater in between samples 21 to remove attached solids. The sediment slurry that settled to the bottom of the wells was pumped to a 2 L glass bottle under Ar atmosphere and the bottle was then sealed with a rubber stopper. The slurry was transferred to the laboratory and centrifuged to separate the sediments from the water. The pellets were frozen at -80 °C prior to being shipped on dry ice to Michigan State University. Fresh sediment slurry was also collected anaerobically in a 27-ml glass pressure tube for the most probable number (MPN) enumeration. Samples taken by this method are a mixture of the sediment along 1 m depth of the well (screened interval) and could contain sediment in the deep matrix where electron donor did not reach. Before remediation, bioactivity in the treatment zone was extremely poor due to low pH (3.2-3.6) and high levels of nitric acid and uranium. DNA extraction from untreated sediments repeatedly failed. In order to compare microbial communities before and after biotreatrnent, DNA was collected by filtering 1700 liters of groundwater from the F W106 well which is located 12 m away from the treatment zone in parallel with treated sediments. Groundwater from FW106 has similar composition to that in the treatment zone before remediation, i.e. high levels of nitric acid and uranium and low pH (Table 2.1). 22 80:838. bomb 2: no 833 CA: 35 :03 .838? on“ 8 5:00:88 0:28:83 mamas—8828 :05 9 888888 888885 28 8:83 .83 mmz86 no 2052,385805 88.8 ”m Axum >83 2: 3w 8.8 8588 Roam ._tg<-b88nom E :83 p.53 moifimmv 8288:8882 20.83 ”8.... a. 8882 88 8 8.8 88 2 888 8.2 8.8 8.8 88.8 3888 8;: .xév 88 88 88 88 888 88 888 88 82 88 88 m 8838 8 8882 :8 R8 888 88.8 8.2 88.8 888 88 8882 8.8 < 88 28 88 88 88 $8 88 :8 8.8 82 88.8 888 m 38;: 8 88.5 888 2 8.8 8.8 88 88.8 88.8 88. _ m 88.: _ 8.8 < 8 888 888 888 :8 888 88 £8 88 _8._ 88.8 88 m 8.88;: 8 88.: a. a: 8.: 8.8 a. 88.8 8.8 88.8 88.5 8.8 < t 88 88 888 :8 N88 88 8.8 :8 82 88.8 88 m 8-8.3.8 8 8882 a: a. 88 8.8 a. 88.8 88 8.8 88.5 8.8 < a 28 888 88 88 $8 888 88 8.8 88 88.8 88 m 8-2: 3.. 8 888.2 :8 2 8.2 8.8 8.2 88.8 88 8.2 888: 8.8 < 6 R8 .88 88 88.8 888 88 88 88 82 88.8 8.8 m 8;: 8 8882 :8 a. 88.: 8_ .8 88.2 88.8 88;: 88. _ m 8888 88.8 < r. 28 2.: 25 2:. 2:. 26 28 25 28 28 98:8 2 88 =2 82 8 2 am -6 28:88 2882 :8 .8238 =83 823.8 :83 75:8 2: can 8:860:88 coca new 883 828 8585828 .80 8:83 2: E 82828::on 80 828.59er 303.25 .~.N 25,—. 23 Microbial Enumeration. Denitrifying bacteria, FeRB and SRB were estimated using the MPN technique with five tubes for each dilution. Anaerobic pressure tubes (27 ml) containing 10 ml basal medium were sealed with butyl rubber stoppers with aluminum caps. The basal medium contained the following components (per liter): NH4C1, 0.5g; NaCl, 0.4g; NaHCO3, 0.55g; and mineral solution, 100 ml. The mineral solution contained (per liter): MnClz, 0.4 g; MgSO4, 1.5 g; CaClz, 0.5 g; and yeast extract, 0.02g. The medium was prepared under a Nz-COZ (99:1, vol/vol) atmosphere and distributed to each pressure tube (10 ml per tube). After autoclaving, a sterile trace element solution (0.4 ml) and a sodium trimetaphosphate solution (50 mM, 0.025 ml) was added into each tube to obtain pH 7.0. The trace element solution contained (per liter): HCl (12N), 6.4 ml; FeC12-4H20, 0.3g; ZnSO4-H20, 0.1g; MnSO4, 0.085g; boric acid, 0.06g; CoC12-6H20, 0.02g; CuSO4, 0.004g; NiSO4-6H20, 0.028 g; and NaMoO3-2H20, 0.04g. The electron acceptor solution was added to the tubes to obtain a final concentration of 5 mM Fe(III)-citrate for FeRB, 8.76 mM of sodium thioglycollate and 33 mM of F eSO4 for SRB, and 9.9 mM of KNO3 for denitrifiers, respectively. Ethanol solution (1 M) was added to each tube to give a final concentration of 10 mM. Groundwater from FW106 was pumped from the wells into anaerobic pressure tubes pre- filled with nitrogen gas. Sediment slurries were collected in pressure tubes under anaerobic conditions. The sample was then serially diluted in MPN tubes. The tubes were incubated at room temperature for two months. Tubes were compared to controls for scoring as positive or negative for production of gas in denitrifying tubes, color change in 24 F eRB tubes, and production of black Fe(II) sulfide precipitates in SRB tubes. Tracer test. A tracer study was performed to characterize the groundwater flow in the treatment zone. The hydraulic flow path is expected to affect the delivery of nutrients and thus, affect the metabolism of ethanol in the groundwater and the microbial community diversity. Sodium bromide, a conservative tracer, was injected through the FW104 well together with ethanol from days 801 to 803. An ethanol/NaBr tracer solution with COD/Br- of 2.46 g/g was prepared and injected into the recirculation line of the inner loop, resulting in injected concentrations in well FW104 of 50 mg/l bromide and 1.0 mM of ethanol. The tracer test continued for 50.75 h. Samples were periodically taken from the injection, extraction and MLS wells for analysis of Br-, COD, ethanol and ac etate . DNA extraction and community analysis. DNA was extracted from 0.5 g of sediments with the Fast soil prep kit (MoBio Inc., San Diego, CA, USA) following the manufacturer’s instructions. DNA was used to amplify 16S rRN A genes using the universal primers 27F (5’- AGAGTTTGATCMTGGCTCAF-3’) and 1392R (5’- ACGGGCGGTGTGTRC-3’) in a Stratagene thermal cycler (Stratagene, La Jolla, CA, USA). The PCR reaction (50 pl) contained 20 mM Tris-HCL (pH 8.4), 50 mM KCl, 3 mM MgC12, 0.1 pg/ul BSA, 10 pmol of each primer, 0.2mM each dNTP , 1.25 U taq polymerase (Invitrogen Life Technologies, Carlsbad, CA, USA) and 20 ng of purified 25 DNA. The PCR cycling conditions were: 95°C for 5 min; 95 °C for 1 min, 59 °C for 1 min, 72°C for 1min 405, for 28 cycles; 72 °C for 10 min. PCR products were analyzed in a 1.5% (wt/vol) agarose TAE gel to confirm the size of the product. Four replicate PCR reactions were generated for each DNA extract and then were compiled to address variability that may be introduced by PCR bias. The PCR products (200pl) were concentrated to 30 ill with a PCR purification kit (Qiagen Inc., Valencia, CA, USA). The concentrated products were then run in a 1% agarose gel, excised, extracted with the QIAquick Gel Extraction Kit (Qiagen Inc., Valencia, CA, USA) and eluted to 30 pl with EB buffer (10 mM Tris-Cl, pH 8.5). An additional step was taken to add poly-adenine overhangs to the PCR products to facilitate cloning (10 min at 72 °C, 18.8 mM Tris HCl pH8.4, 47mM KCl, 0.93 mM dATP and 0.5 U of taq polymerase; Qiagen Inc., Valencia, CA, USA). Products were cloned using the Topo TA Cloning kit for sequencing following manufacturer’s instructions (Invitrogen Life Technologies, Carlsbad, CA, USA). Single extension sequencing was conducted by Macrogen Inc (Seoul, Korea) from the internal primer 27F. The sequences were submitted to to the GenBank under accession numbers EF692646-EF693732. Raw sequencing files (Abl) were submitted to the myRDP application of the Ribosomal Database project where the nucleotide sequence was determined, the quality (Q score) of each base stored and vector sequence removed (7). The resulting sequences were aligned to the RDP model and rRN A distances matrices were generated with the Jukes-Cantor distance correction. Operational taxonomic units (OTUs) were defined at 97% sequence identity. The distance matrices were used to calculate alpha diversity indices (Chaol, Simpson and Shannnon) and rarefaction curves using Distance-Based 26 Operational Taxonomic Unit and Richness (DOTUR) determination program (35). The diversity indices were rarified to account for differences in number of sequences per library. The distribution of the OTUs in the different libraries was used in EstimateS (8) to make comparisons based on diversity patterns using Bray-Curtis and Sorensen beta diversity indices. These indices were used to cluster the samples according to their distances using MEGA 3.1 (20). The composition was normalized to account for differences in number of clones in the different libraries. I-LIBSHUFF (36, 38) was used to compare the clone libraries according to Good’s homologous and heterologous coverage. This approach provides a quantitative comparison of 16S rRNA gene clone libraries from environmental samples (3 8). A mask using the quality value (Q>20) was used and the resultant sequences were used for classification with RDP classifier using a 80% confidence value. MALLARD (4) was used to detect sequences with anomalies such as chimeras. The putative chimeras were later re-evaluated using the RDP Sequence Match with a suspicious-free and near-full-length data set and with the Pintail program (3). Sequences confirmed as anomalous with Pintail were excluded from the analyses. Phylogenetic trees were constructed using distance and maximum likelihood methods. Aligned sequences were downloaded from the RDP, aligned manually and the non-model positions were masked from the alignment. The Jukes-Cantor distance correction, full gap deletion and a bootstrap test using 10000 replicates and random seeds were used to construct the trees in MEGA v3.1 (20). 27 Chemicals and analytical methods. The source and quality of chemicals used in the field test were described previously (48, 49). COD was used as an overall indicator to monitor the consumption of electron donors (ethanol, its metabolite acetate and others). COD, sulfide and Fe(II) were determined using a Hach DR 2000 spectrophotometer (Hach Chemical, Loveland, CO). Anions (including N03: Br-, Cl-, S042. and P043.) were analyzed with an ion chromatograph equipped with an IonPac AS-14 analytical column and an AG-14 guard column (Dionex DX-120, Sunnyvale, CA), metals (Al, Ca, Fe, Mn, Mg, U and K etc.) were determined using an inductively coupled plasma mass spectrometer (ICPMS) (Perkin Elmer ELAN 6100), and U reduction state was determined with XANES as described elsewhere (49, 50). Ethanol and acetate were determined by a HP5 890A gas chromatograph equipped with a flame ionization detector and an 80/ 120% Carbopack BDA column (Supelco Division, Sigma-Aldrich Corp., St. Louis, MO) using Helium as carrier gas. RESULTS In situ biostimulation test. After one year of biostimulation, nitrate concentration in the treatment zone dropped to a level of nearly zero. Increase in sulfide concentration and appearance of F e(II) in groundwater suggested that sulfate and iron reduction were occurring (50). Microbial activity during biostimulation was determined by monitoring changes in aqueous concentration of electron donor added (as COD), sulfide and uranium. A representative time course of biostimulation is shown in Figure 2.2. On day 28 704, ethanol was injected to FW104, causing increase COD concentrations in all major MLS wells (Figure 2.2A). I80 +E -E ‘+E ‘_ 8150* E A Oi 000C) E12010 23° 020:: d90'I 0A .o O 060.3,.“2 A 030 --" 2""; "A O i o 053°010833138 p 15300 0935 aJ 704 706 708 710 71'2 714 71 1.00 B SUlfide, mM .9 S 0.01 704 706 708 710 7l2 714 716 0.37 2 Q 1 9. . . 8.02.130 9 o . g 800083. 0 0° 9; 0.. E A" 4344.13 0 A 0 A AA AA dooo 201%.. M . . .“ I. I I I . I ' I I I I - . 0.0 704 706 708 710 71‘2 71'4 716 Time, days 9FW101-2 AFWlOl-B I FW102-2 O FW102-3 El FW104 O FW026 Figure 2.2. Typical biostimulation of U(VI) reduction by injecting ethanol to subsurface (day 704 to 714). (A) COD concentrations. (B) Sulfide concentrations. (C) Uranium concentration changes in MLS. U concentration in the injection well FW104 was 0.5 pM during this test period and is not shown due to scale. +E indicates start and —E indicates stop of ethanol biostimulation. 29 The COD in FW104 was mainly attributed to ethanol (>80%) and a small amount of acetate (<30 uM) while in MLS and F W026 wells, the COD came mainly from acetate (data not shown). No ethanol was detected in the MLS wells and FW026 well and methanol concentrations were below detection limit (<50 uM) for all the cases. On day 706, ethanol injection was stopped and the COD concentrations in MLS wells decreased rapidly. The same pattern was observed again on day 711 to 713 when ethanol was reinjected. Sulfide concentrations in all MLS wells increased after ethanol was injected and decreased after injection stopped (Figure 2.2 B) but remained at detectable levels (>20 uM). Sulfate concentrations in MLS wells decreased when ethanol was injected and rebounded when the injection stopped (data not shown) indicating the presence of SRB activity in the subsurface. Uranium concentration in all MLS wells decreased after ethanol was delivered and slightly rebounded when ethanol was not injected (Figure 2.2C). However, the uranium levels after the rebound were lower than those before ethanol injection. This rebound was likely due to uranium being carried in the recirculated water (~0.5 um) and to the lack of electron donor for U(VI) reduction when ethanol injection stopped. During the ten-day test in figure 2.2, uranium concentration in the injection well F W104 was around 0.5 pM (data not shown in figure 2.2C) which was much higher than those in the MLS wells. Uranium levels in F W1 01-3 and F W102-2 were near or below EPA MCL (0.126 uM) throughout the 10 day-test period while U levels in FW101-2 and FW102-3 dropped even below 0.126 uM after ethanol injection. The relatively higher U levels in the latter two wells are likely due to the lack of enough electron donor after ethanol injection stopped. Lower U concentrations in the MLS wells 30 were achieved later as shown in Table 2.1. Aqueous F e(II) concentrations were 10-20 pM in the MLS wells before ethanol injection, dropped below 5 pM during ethanol injection as sulfide concentration increased and then slowly rebounded after ethanol injection stopped and sulfide concentration decreased (data not shown). The decrease in Fe(II) concentration during ethanol injection is likely due to formation of more F eS precipitates by reaction with the produced sulfide. The FeS precipitates likely accumulated in sediments. The reduced sediment samples showed black color (F W104) or dark green color (MLS). Reduced U(IV) was detected by XANES in the sediments samples from injection well and the MLS (Table 2.1). In the F W026 well, the U(IV) content was below the detection limit of XANES (<10% of total U). The highest content of reduced uranium was found in the injection well FW104 (50). Groundwater flow pathway. The injection of the NaBr'/ethanol solution to F W104 lasted for 50.75 h. Bromide concentration increased in the MLS and extraction (FW026) wells with different recovery ratio and mean travel time (Figure 2.3). These results indicate that all MLS wells were hydraulically connected to FW104 (Figure 2.3A). The Br' recovery percentages were: FW101-2, 93%; FW101-3, 60%; FW102-2, 94%; FW102-3, 93%; and FW026, 50% (Figure 2.3A). FW101-2, FW102-2 and FW102-3 received more than 93% of water injected to FW104, while F W101-3 and FW026 received 50% and 60% of water from FW104, respectively. The rest water was from surrounding areas. Mean travel times from F W104 to the different wells were: FW101-2, 31 2.84 h; FW101-3, 18.4 h; FW102-2, 11.6 h; FW102-3, 3.7 h; and FW026, 46 h (23). Therefore, groundwater injected to FW104 reached FW101-2 and FW102-3 more rapidly than other two wells. The fraction of groundwater flow from injection well to each MLS well and the mean travel time may significantly influence microbial community in the MLS wells as discussed later. During the tracer test, COD concentrations in MLS and FW026 wells increased after ethanol injection (Figure 2.38). The measured COD concentrations in the MLS wells were significantly lower than those calculated based on the Br' concentration and the COD/Br' ratio in the NaBr-ethanol tracer solution used. The difference was likely due to biodegradation of ethanol and acetate in the subsurface. Ethanol was observed only in the FW104 well during the tracer test. The COD in MLS and F W026 was mainly from acetate (>80%) based on acetate vs. COD measured. During the tracer test, sulfide concentrations increased in all MLS wells and U (V 1) concentrations decreased in the same trend as shown in figure 2.2 (data not shown). Another separate tracer test by injecting bromide to outerloop FW024 indicated that innerloop extraction well FW026 received 17% of water injected to FW024 (23). The infiltration of water from outerloop may also influence the microbial community. 32 80 704—A 'FWIO4 __ 5 AFw026 J 60 Emma _ O OFw101-3 ____. E 50 w to an I. O D 0FW102-2 é 40"?8‘9 o AFw102-3 # Cl 0 0 Q :2 30m: ..00 00 a o g 20 o ‘ A m JD; AAA I 9 Q 150 200 I FW104 A FW026 — cr FW101-2 Q . FW101-3 an O FW102-2— E, A FW102-3 a FW101-2— 8 ——r=wror-3 FW102-2 - - - FW102-3 0 010 20 30 4 50 60 70 80 90100 Time (hr) Figure 2.3. Tracer test performed with Bromide on day 801-803 shows the hydraulic connection between injection well F W104, extraction well FW026 and MLS and biodegradation of ethanol. A. Changes in bromide concentrations (23) B. Comparison of measured COD (marks) and calculated COD (lines) concentrations based on COD vs Br- ratio (2.46 g/g). 33 Microbial enumeration. MPN results indicated that after biostimulation for nearly two years, increased levels of denitrifiers, SRB and F eRB were present in the sediments in the bioreduced area in comparison with control well FW106 (Table 2.2) that did not undergo stimulation and shares similar geochemical composition with the active area wells (Table 2.1). After biostimulation, the microbial concentrations in the treatment . . . . 7 8 6 8 5 7 area were (1n cells/g sedrment): denrtrrfiers, 10 -10 ; SRB, 10 -10 ;and FeRB, 10 ~10 . he highest levels for all three trophic groups were found in inner loop injection well F W104 where ethanol was injected. Relatively low levels were found in extraction well FW026. Consistent with the tracer study, FW101-2 (more connected to the injection well F W104 based on the tracer study) showed higher bacterial counts than FW102-2 (less connected) (Table 2.2). MPN counts from FW106 (groundwater) showed levels of denitrifier as low as 3.3 cells/ml groundwater (49). Neither SRB nor F eRB were found in F W106. Comparison with sediment MPN at day 453 showed a decrease in SRB and denitrifiers and no change for FeRB in FW104. MPN count at day 774 for all three groups remained around 108 cell/g sediment in F W1 04 (49). Integration of MPN at day 774 and three earlier timepoints (days 278, 354, 453) for FW101-2 showed an increase in the counts overtime with microbial levels at 774 around 107 cell/g sediment, one order lower than F W104. Conclusions from the MPN analysis must be qualified because we contrast sediments with groundwater samples. However, prior MPN studies using samples from the same well show a one log difference in counts in sediment versus groundwater samples (data not shown). 34 Table 2.2. MPN for three major trophic groups in the sediments of the treatment zone (day 774) in comparison with groundwater from control well F W106. Cells / g sediment Well Denitrifiers FeRB SRB FW104 7.23 x 108 9.40 x 107 1.53 x 108 FW101-2 1.54 x 108 2.05 x 107 2.06 x 108 FW102-2 2.39 x 107 5.48 x 105 1.87 x 107 FW026 1.10x107 1.93 x106 1.07x106 FW106“ 3.3 cell/ml ND ND F eRB: Iron reducing bacteria; SRB: Sulfate reducing bacteria. ND: None detected. Wells are arranged according to their descending hydraulic connection to the injection well (FW104) based on the tracer studies. *Data on day 278 (49). Sequence analyses. After discarding putative chimeric sequences, an average of 155 sequences per sample were used for each library. Rarefaction analysis at 97% similarity levels (Figure 2.4) showed that the estimated coverage (rarified number of OTUs divided by rarified Chaol estimator) ranged from 36 to 58% in the stimulated area and was 93% in the background area (FW106). Even with low coverage in the biostimulated zone, clear differences in diversity were observed. Highest diversity was found in all the wells of the biostimulated area; working at 97% similarity for OTUs the number of OTUs ranged from 41.2 to 91.4 (rarified values). Only 6.6 OTUs were found in the background area FW106 (rarified value). 35 120 o —FW026 00 O -- 'FWlOl -2 1 - , I -—-—er01 -3 I’ll: CI -- --FW102 -2 x", .. I 5 80 -. ------ FW102 -3 ,r’ ['5 <>— ~er04 3 60 ‘ alt — -FW106 3' E :z b O 120 150 180 Number of Sequences Figure 2.4. Rarefaction curves of the 16S rRNA gene libraries constructed. OTUs were defined at 97% sequence identity. The library from the untreated control (FW106) was close to complete sampling (93% coverage). Microbial community structure and major groups detected. All libraries were dominated by members of the phylum proteobacteria, with Betaproteobacteria and Deltaproteobacteria being the most dominant proteobacterial classes in the biostimulated zone (Figure 2.5). In contrast, the background area was dominated by Gammaproteobacteria. Bacteria belonging to groups known to include U(VI)-, nitrate-, sulfate- and Fe(III)-reducers were detected in the active area but not in the background area library. Sequences from the genera Desulfovibrio, Geobacter, Anaeromyxobacter, 36 Acidovorax and Desulfosporosinus were detected in the libraries. These genera are known to include U(VI) reducers and additionally can contribute with one or more of the following activities: iron (III)-, sulfate- and nitrate reduction (Figure 2.6). The contribution of the putative species responsible for the uranium reduction and their richness (measured as number of OTUs in a genus) was highly variable (Table 2.3). Geobacter and Desulfovibrio were detected in all the libraries of the active area. In five of the six samples from the active area, the relative contribution of Desulfovibrio was higher than that from Geobacter (Table 2.3). Anaeromyxobacter, Acidovorax and Desulfosporosinus sequences were not present in all the libraries. Nitrate reducers from different taxonomic lineages were present in the libraries of the active zone but not in the untreated area (FW106). Most of the nitrate reducers were members of the Proteobacteria phylum e. g. F erribacterium (2-27%), T hiobacillus (0- 29%), Sphingomonas (0-6%), Desulfovibrio (4-l6%), Azoarcus (0-5%), Acidovorax (0- 4%), and Pseudomonas (0-1%) among others. The denitrifying Acidobacteria Geothrix was also detected in all the libraries of the active area (4-17%). Nitrate respiration is not exclusive to these groups and the mentioned bacteria are also know for using other electron acceptors, Iron(III) reducers were represented by bacteria from three groups: Deltaproteobacteria (Anaeromyxobacter and Geobacter), Betaproteobacteria (Ferribacterium and Thiobacillus) and the Acidobacteria Geothrix. F erribacterium was the most abundant FeRB contributing an average of 17% to the libraries. Other commonly found soil bacteria like Acidobacteria, Actinobacteria, 37 Planctomycetes and Verrucomicrobia were also present in the samples. These bacteria were present in the active area and not in the untreated area. l00°/o 8- ) r > > 890%" wig"; - :— 80% ‘ — E . g 709 -- . '- 0 I ..4} '3: ‘j E :31: '5 60%, . ..0_’i — 8 331: ho}; . LE 50% - -’ " '- coo - o 9.. = 0.. . .9 40% ‘ '- E E 30% " '- § I o 20% ‘ '— .2 § 10% 4 0% ‘ FW026 FW101-2 FW101-3 FW102-2 FW102-3 FW104 FW106 Wells U. Bacteria - Chloroflexi I Alphaproteobacteria :-I Other D Bactemides D Gammaproteobacteria 5“; . _ . . . R§ Actmobacterza Aczdobacterza @ Dellaproteobacterza a F irmicutes [IIIII] U. Proteobacteria I Betaproteobacleria Figure 2.5. Microbial composition of the clone libraries based on the RDP Classifier. The “other” group category includes the phyla Spirochaetes, Gemmatimonadetes, Verrucomicrobia, Chlamydia, Planctomycetes, Nitrospira, Cyanobacteria and the proposed phyla, OP] 1, OPlO, BRCl and TM7. Bacteria that could not be assigned with the 80% confidence bootstrap value were included in an artificial “Unclassified Bacteria” (U. Bacteria) taxon. Wells are arranged according to their descending hydraulic connection to the injection well (F W104) based on the tracer studies. 38 Figure 2.6. Neighbor joining tree showing the relationship of selected representatives (shown in bold) from the groups similar to known U-reducing bacteria. Metabolic abilities of the clones’ closest cultivated relatives are indicated. Non-model positions from the 16S rRN A were masked and J ukes-Cantor distance correction used. Type strains have a “(T)” label. Bootstrap values (10000 repetitions) are displayed if larger than 50%. (+) Present in closest relatives, (-) absent in closest relatives, * activity found some species of the genus, ? = unknown. The range of relative contribution to the different samples is also shown. 39 8... _ 82 9.8 2882.8 $888.8 HOQOU -0 HOHQOOOQ -0 E negmitau maemmoaomaoKNaaoQ _ 3?: .33 .flnaewnoaomnobaufillroc mu: «~33— +\. + + + + *1 + §=~ueaemue§auu§ _UaNN 82.883 3.3—=25 3. 3158:888— 8 €e§3odee¢Ro~oI 758288§JR .\ 24 n: E 2838:8886 82 - . - . $188.88.? 8 .. + A. + + 5.333 a- n? «:95 d8 omeauoNotvoSH—socD ma 81588388 r €153.38— ? 8 . . .x. I. 1 82E 83.388 823.35%. S A. A. 1T + I + + 50 9‘ Qagevg Aw ~"fivg— _INNAvfiBh—L . 8 8.. (TON AbacuzoMSefifi afloacoRAEoLoucV Gus—2. 383.,— 1.2: a 8815 538;?— . 33.: once. 28 Eaauflamotfiaobaqu 3 - - . - 2 .15 :333... 8 + 1. + + 8.8.8.835 88 28 83888.8 8,2838% 8 3158288 82 r «mac .8... 8.8.8.8833 8. 8 N m. - m0 AC aeooaheh-3o§o~oecoob I. FeSESEJH—r {cu—1c ~10 Soofizma 882586 2: - - 315.8385... - +\ + + + + 03336 AC nozowfifi 8.238% 9: cc— §158243E 88 Nm .2932 $3236 ..\o «1: _mm in mocgonaonk no no 315 8333.8: 8 822833. :38. E 8.838 822833... 88. _III ccvm buzgoaoeex 3.238333% mama—M a§~n£o3o§£~ LSoucozehoam 3 mad fl . + + + . + . fir I ' ‘ Methanol Ethanol Acetate U M) sulfate NItrate Fe (III) 8:288. meg—8:5 5923 8 88:88 338w Bob A23 5 :30an 8038858952 @2028 no 8:588:28 05 wages—m ooh wEEoT—onnwmoz .e.~ 0.53...— 40 _ o o o _ N Rm €§s§§£ m §§§uw .wzzmto .Q o v N o o _ .Emfimfitau .Q n mxzwmoaomwobzme m _ o. v o_ w 2 n:5§£am\ .b a 538mg 0 N m m w _ . azuzmmfiafih .V N amuuunoxkficxmuxw MEEQEQAMQ. D Emuzhmhfizx .D mmzmmfifi .b 333 6 _ _ S _ _ m $852? 6 n 333.3 mzoaofifau .Q o n v S 2 0 £83 .Q n etfiafiafi 3.95m m-_o_>>n_ N-No_>>n_ m-NS>?.m N‘s—33m veg/m 833—2 ax; :25 8a 8:33.80 gnu—om toga—0% “memo—U mDEO 3:00 68.8qu 03 and ® £03839 83o£ “memo? can Abra—Ea $3 v mDHO 26E: no .8583: 2F i=9: n no.2 2: 5 8.958 Beam—.3. SP: 8295: 28.9 E .223» Seacoaéaiaaua =32:— ue Aekvueuafihaeo «>323— .m.~ ~35. 41 Diversity analyses. All libraries from the stimulated zone showed greater diversity than the untreated zone based on both evenness and richness values as indicated by the diversity indices used (Table 2.4). One single OTU dominated the untreated area sample (FW106). This member of the Xanthomonadaceae family was also detected in a metagenome experiment for well FW106, where it was shown to carry a variety of metal resistance genes (13). Table 2.4. Diversity indices for the different wells in the treated and untreated area (F W106). Index Chaol Shannon (H) Simpson (l-D) Well (LCI, HCI) (LCI, HCI) FW104 125.5 (91.7, 200.6) 3.7 (3.5, 3.9) 0.97 FW101-2 76.5 (54.4, 137.9) 2.9 (2.7, 3.2) 0.90 FW102-3 87.8 (64.6, 147.7) 3.4 (3.2, 3.6) 0.95 FW102-2 250 (168.8, 415.1) 4.3 (4.2, 4.4) 0.99 FW101-3 114.3 (81.1, 192.9) 3.5 (3.3, 3.7) 0.95 FW026 124.9 (88.3, 210.4) 3.7 (3.5, 3.8) 0.97 FW106 6.0 (6.0, 6.0) 0.7 (0.5, 0.9) 0.31 LCI and HCI are rarefied 95% lower and higher confidence intervals (provided by the DOTUR application). Wells are arranged according to their descending hydraulic connection to the injection well (FW104) based on the tracer studies. Diversity-based clustering revealed that the F W106 community formed a cluster separated from the treatment zone wells (Figure 2.7). The two different clustering methods used resulted in congruent topologies. High similarity was observed between communities from FW102-3 and FW104, which were 53% similar according to a calculated Bray—Curtis index. l—LIBSHUFF comparison indicated that these two libraries were not significantly different (p=0.05). All the other pairwise comparison were non- 42 significant (libraries were significantly different, p=0.05).. Additional comparison of F W104 and FW102-3 communities with the LIBCOMPARE function of the RDP showed non—significant differences at (p=0.01) at all the different levels of taxonomy from phylum to genus. No clear relation was observed between the alpha diversity indices and the hydrology. FW026 FW101-3 FW102-3 FW104 FW101-2 FW102-2 FW106 *-—01——* Sorensen Distance FW026 FW101-3 { FW102-3 FW 104 FW101-2 FW102-2 FW106 T Bray-Curtis Distance ,1. Figure 2.7. Clustering of the samples according to diversity patterns. The topology was similar when using the Sorensen index (Presence/Absence) and the Bray-Curtis index (Presence/Absence and abundance). The indices were normalized to account for differences in total number of sequences per library. 43 DISCUSSION 1. Remediation of uranium and metals in the subsurface Microbial reduction of metals including uranium has been the proposed alternative to control the mobility of the contaminants in the groundwater (45). In this field experiment, ethanol injection successfully created a reducing environment capable of achieving uranium remediation and immobilization to levels below US EPA drinking water MCL (49). The biostimulation process changed the structure of the microbial communities from a small and low-diversity community to a more abundant and diverse community with bacteria capable of reducing the contaminants nitrate, sulfate and U(VI) in situ. 2. Microbial communities 2.a Major trophic groups detected Microbial enumeration together with clone libraries revealed that after biostimulation viable FeRB, SRB and denitrifiers had grown in the treated area where uranium, nitrate, sulfate, ferric compounds and perhaps other compounds served as electron acceptors. At least three major microbial trophic groups appear to be involved in the bioremediation of the area. The first group, denitrifiers, can remove nitrate and provide a favorable low redox environment for U(VI) reducers, FeRB and SRB. These latter two groups may contribute to the remediation process by having members capable of reducing not only Fe(III) and sulfate but also U(VI). Microbial enumeration indicated 44 that these three trophic groups were present in the sediments at high levels after biostimulation. In contrast, the untreated area sample (F W106) showed the presence of low numbers of denitrifiers (3.3 cells/ml) in groundwater but neither SRB nor FeRB. Previous reports on groundwater MPN for this stimulated area (same wells, earlier timepoints) also show these three groups but the microbial counts were lower that our timepoint (49). The cell count in the untreated sediments is unknown but it could be significantly low since DNA extraction was unsuccesful. Clone libraries permitted detailed speciation of these trophic groups. 2.b. Carbon source as selective agent and the community of ethanol degraders The use of ethanol as a carbon source likely increased the diversity of taxa and physiologies. Injected ethanol provided both the carbon source and the electron donor in the forms of ethanol, acetate (metabolic intermediate) and methanol (an impurity in industrial ethanol). Electron donor consumption and metabolism were likely performed by denitrifiers, FeRB and SRB (Figure 2.6). Based on groundwater analysis during tracer study, we hypothesize that the SRB were mainly responsible for ethanol consumption and acetate release between the injection well and MLS wells (where no nitrate was detected and sediments showed black or dark green color), while FeRB and denitrifiers utilized mainly acetate between MLS wells and the extraction well. At this well there was evidence of groundwater infiltration from the outer loop and the sediments were green- yellow in color. Close relatives of the SRB detected (Desulfovibrio spp. and Desulfosporosinus spp.) utilize ethanol and not acetate when doing sulfate reduction 45 though few strains of both groups can also use acetate (19, 33). The release of metabolic acetate can be used later by other species, including Geobacter. This hypothesis is supported by sulfide accumulation during ethanol injection (SRB activity) (Figure 2.2) and the detection of mostly acetate (and not ethanol) as carbon source in less connected wells. The sequential utilization of ethanol and release of acetate likely create a niche for acetate-consuming F eRB and acetate-consuming denitrifiers. The first group was represented in our libraries by Geobacter spp., Geothrix spp. and Anaeromyxobacter spp. These bacteria are capable of acetate and ethanol degradation mainly by Fe(III)-reduction as well as by denitrification (2, 6, 34). The second group is represented by denitn'fiers such as Acidovorax spp., and T hauera spp. These Betaproteobacteria members utilize acetate and ethanol as electron donors and have been previously found in the aboveground denitrification reactor used to pre-treat groundwater at this site (16). Methanol utilization may play a minor role because its small contribution (5% in the ethanol solution used), nevertheless it can be used for denitrification by almost all denitrifiers and also for sulfate reduction by some Desulfosporosinus spp. (33) and Desulfovibrio carbinolicus (25). 2.c. Other C sources and contaminants In addition to the C sources used to stimulate the microbial activities, other C sources such as aromatic and chlorinated compounds, humic acids and cellular components are present in the contaminated area. Aromatic (phenols) and chlorinated compounds (TCE) were detected in the 46 groundwater prior to biostimulation but mainly removed during site preconditioning (48). These compounds are still present in the groundwater outside the treatment area and could have infiltrated to the inner loop (based on tracer test results) and provided additional carbon sources, electron donor and acceptors to the underground communities. Several detected groups of different phylogenetic affiliation show some potential for dechlorination or aromatic compounds degradation. In the case of dechlorination, at least 50% of the Geobacteraceae sequences were related to G. lovleyi, an isolate known to use chlorinated compounds such as TCE and PCB as electron acceptors (39). Close relatives of the chloro-phenol respirer Anaeromyxobacter dehalogenans 2CP-1 (34) and PCB respirer Desulfosporosinus meridiei were also detected (33). Aromatic compounds can also be used as carbon sources by some bacteria of the Acidovorax genus which was detected in the sediments. The Betaproteobacteria Pseudomonas sp. strain P51 can degrade chlorinated benzenes (43). Aromatic compounds and solvents present at the beginning of the bioremediation operation could have influenced the initial communities before the biostimulation began though their concentration is generally low and it is not clear if they can support growth at levels present at the site. Humic substances can be used as electron acceptors (21). Sequences related to the humic-acid reducers Geothrix sp. (6) and G. humireducens (21) were detected in the libraries. Reduced humic substances can potentially abiotically reduce UWI) far from the bacteria and they have been shown effective in increasing F e(III)-reduction in subsurface environments (27). In addition, other bacterial groups not directly related to bioremediation, as far as we know, were found in the sediments e. g. Planctomycetes, C hloroflexi, Actinobacteria, 47 etc. They are likely involved in the degradation or digestion of dead cells, soil humics, and extracellular substances produced during biostimulation in the subsurface. Chloroflexi members are known to grow in filaments, produce hydrogen and are proposed to be involved in dead cell recycling (52). This group can potentially contribute to the stability to the bacterial community by promoting the formation of biofilms and can facilitate bioremediation by transferring electrons in the form of hydrogen to other groups more likely involved in bioremediation such as Desulfovibrio spp. The Chloroflexi group was the only group not belonging to the F eRB, SRB or denitrifiers present in all the libraries from the active zone. 2.d. Putative genera involved in U(VI) reduction Analyses of uranium oxidation state in sediments by XANES confirmed that the decrease in groundwater U concentration was due to reduction of U(VI) to U(IV). Elevated total U and U(IV) percentages of total U (18, 49, 50) were found in the inner loop injection well FW104 consistently with effective bioremediation (Table 2.1). In this study, we found a variety of previously reported UWI)-reducing bacteria present in the reduced sediments including FeRB Geobacter spp. and Anaeromyxobacter spp., SRB Desulfovibrio spp. and Desulfosporosinus spp. as well as the denitrifier Acidovorax spp. Our results suggest that uranium reduction cannot be attributed to a single group and it is very likely that this role is taken by several different bacteria. The contribution to the uranium reduction based solely on the relative frequency in our libraries and the reported literature ranks the contributor groups as following: 48 Desulfovibrio > Geobacter > Anaeromyxobacter > Desulfosporosinus > Acidovorax. Desulfovibrio spp. were detected in all the samples of the active area and their contribution peaked in wells with high percentage of reduced uranium. This group was the most abundant group with reported uranium reducing abilities and was found in the sediment samples. Even though there are no reports of U(VI)-reduction by the closest relatives of the cloned sequence, the high frequencies of these bacteria in the sediment clones (up to 16%) suggest a substantial role in the groundwater ecosystem. Geobacter spp. were found in all sediment sample of the active area. Our sequences were closely related to the uranium reducer G. lovleyi (39) and to the humic- acid reducer G. humireducens (21). Additionally, Geobacter spp. was found to be associated with U(VI) reduction at sites contaminated with uranium and U ore where acetate was added (2, 28, 44). Anaeromyxobacter sequences were detected in all sediment samples from the active area except for FW026. These sequences were related to the known U(VI) reducer A. dehalogenans (47) and to a clone from a uranium mine sediment were uranium reduction was demonstrated (40). Desulfosporosinus sequences were present in half the libraries of the active area and the closest isolate to most of the sequences was D. orientis, a SRB also known for reducing Fe(III), nitrate and U(VI) (26, 40). Despite their relative low contribution to the fatal community, Desulfosporosinus spp. may play a minor role in U(VI) reduction and a ‘0) gger role in the long term stability of the reduced uranium as they can form spores and survive under starvation conditions. 49 Acidovorax sequences were found in half the libraries of the active area. An Acidovorax sp. was shown capable of U(VI) reduction in microcosms tests with sediments from the FRC (29). Nitrate and nitrite has shown being able to re—oxidize and remobilize Fe(III) (3 7) and the presence of this denitrifier could contribute to the removal of these competing electron acceptors and ensure the stability of the reduced uranium. Geothrix spp. sequences were found in all the libraries of the active zone. This iron(III) reducer can use humic acids as electron acceptor, and has been found during uranium reduction events in the F RC Area 2 using 16S rRNA gene microarrays and enrichment studies (5). The Geothrix genus has not been characterized for U(VI) reduction, but based on the number of clones retrieved, it was an important member of the community. It is possible that its contributed to U(VI) reduction indirectly via reduced humic acids or reduced iron(II) compounds. Abiotic U(VI) reduction may also play a role under our operational conditions (i.e. pH 5 .8-6.6 and HCO3- < 5 mM). Sulfide, the end product of SRB; and green rusts, products of F eRB can both reduce U(VI) to U(IV) (15, 30). Indirect U(VI) reduction by reduced humic substances can also be contributing an important piece given the constant presence of F eRB humic-reducing Geothrix spp. in all the tested sediments from the active area. Therefore, the activity of SRB and FeRB could also indirectly contribute to the reduction of U(VI) and maintenance of a stable and low level of uranium. 3. Dynamics of the community 50 3.a. Patterns of diversity detected Microbial diversity varied in the sediments of the treated zone (Table 2.4). The utilization of standardized techniques in all of our community analyses allows us to conclude that the observed differences were due to real differences in the community, and not the consequence of PCR bias. The irregular pattern observed is likely due to the heterogeneity of groundwater flow and the distribution of contaminants as shown by the tracer studies. Microbial enumeration analyses for the three trophic groups studied (denitrifiers, FeRB and SRB) showed a decline in cell counts going from the injection well and its more connected wells, to the less connected wells (Table 2.2). Even though replication did not allow for statistical conclusions, some apparent trends were observed. The different approaches used to study the diversity and to compare the communities yielded consistent results and the clustering analyses showed a topology that was in agreement with the tracer studies depicting the groundwater flow. Because the flow of the water is not homogenous, gradients of electron donors are expected as the microbial communities consume and convert the injected ethanol solution. The hydrology clearly affected the microbial counts (MPN) but no clear relationship between the water flow and the alpha diversity indices was found. Though this may be due to incomplete sampling, it is more likely that the ethanol injection created a selective pressure for specific functions (like iron reduction, ethanol utilization, etc.) but not for specific bacteria. Bigger gradients and more divergent communities would be expected in natural systems where the water flow is not controlled. Having more divergent communities in an area of remediation adds additional layers of complexity that can make the monitoring of the performance more difficult. Thus, control of the hydrology is key to have a more 51 homogenous response to the bioestimulation. When analyzing the carbon and electron acceptors, more ethanol-consuming organisms were detected in the more connected wells. Consistent with the expected sequential electron acceptor utilization (49), more denitrifiers where present in highly connected wells and FeRB were more abundant in the wells with lower connectivity. The exception to this pattern was Desulfovibrio, a SRB that showed high relative abundance in more connected wells, possibly due to its ability to utilize the injected ethanol. Desulfovibrio spp. were more abundant than Desulfosporosinus spp., also a SRB capable of using ethanol. The presence of Desulfovibrio in the aboveground bioreactor (16) could have given Desulfovibrio spp. an initial competitive advantage over Desulfosporosinus spp. by a continuous inoculation of Desulfovibrio spp. carried in the treated water from the bioreactor. On the other hand the ability of Desulfosporosinus to sporulate and degrade methanol may account for its survival. Overall the microbial methods were in agreement with the hydrological studies. However, more study is needed to understand the relation quantitatively. 3.b. Emergence of previously undetected populations After preconditioning by pH adjustment and removal of inhibitors, the ethanol injection increased the microbial diversity of the subsurface as shown by diversity indices that consider richness and abundance (Table 2.4). The activity of the new complex bacterial community created favorable conditions where F eRB, SRB and denitrifiers thrive in contrast with the untreated area where they were either absent or in very low 52 levels. It is very unlikely that the microbial community observed was fully derived from indigenous species that survived the extreme conditions of the area. The most likely source of new colonists was the upper soil where nitrate is low and pH neutral. These bacteria could have been transported by either natural groundwater infiltration or forced recirculation during the treatment. Geobacter and Anaeromyxobacter spp. have been found in FRC Areas 1 and 2 using iron-reducer enrichments with acetate and lactate (31). Geobacter spp. and Geothrix spp. have been also detected in the F RC Area 2, using high density 16S rRNA gene arrays (5). Pseudomonas spp. and Azoarcus spp. have been detected in the FRC Area 3 with 16S rRNA clone libraries prior to biostimulation (9). Therefore, these microorganisms are present in pH neutral soils at the site. A second likely microbial source could be the aboveground bioreactor system. This reactor worked for 400 days to remove nitrate from the groundwater and the treated nitrate-free water was reinjected into the subsurface (48, 49). The reinjected water likely carried some bacteria to the treated area although it was filtered. Sequences related to Desulfovibrio, Thauera, Azoarcus, F erribacterium and Acidovorax have been previously reported for the bioreactor (16) and were detected in the collected sediments of the active area in this study. Our results suggest that biostimulation efforts successfully fostered communities comprised of a variety of bacterial groups (Geobacter, Desulfovibrio, Geothrix, Anaromyxobacter, Desulfosporosinus and Acidovorax) involved in the groundwater remediation process. This study provides a detailed view of the differences and 53 similarities among the microbial communities throughout the active area that correlated with the path of groundwater flow depicted by tracer studies. The results demonstrated that microbial communities can be established by in situ biostimulation with an electron donor to achieve successful reduction of U(VI) concentration below US EPA MCL. These findings contribute to an improved understanding of the composition, variability and controls on microbial communities in the subsurface associated with a successful bioremediation process and provide a foundation for future implementation and monitoring efforts applied to this and other contaminated sites. Acknowledgements The microbial study was funded by the US DOE under grant DE-FG02-97ER62469 and the fieldwork was funded by the US. DOE Environmental Remediation Sciences Program (ERSP) under grant DOE C05-OOOR22725. Dr. Mary Beth Leigh was supported by US NSF Postdoctoral Fellowship during this study. The authors thank Hui Yan, Kenneth Lowe and Tonia Mehlhorn for help in field and analytic work and Benli Chai from the Ribosomal Database Project for his help in formatting data. We appreciate Dr. Chris Hemme for providing a groundwater DNA sample and Anthony Gaca for the assistance in the construction of the clone libraries. 54 REFERENCES Anderson, R. T. 2006. DOE genomics: Applications to in situ subsurface bioremediation. Remediation Journal 17:23-38. Anderson, R. T., H. A. Vrionis, I. Ortiz-Bernad, C. T. Resch, P. E. Long, R. Dayvault, K. Karp, S. Marutzky, D. R. Metzler, A. Peacock, D. C. White, M. Lowe, and D. R. Lovley. 2003. Stimulating the in situ activity of Geobacter species to remove uranium from the groundwater of a uranium-contaminated aquifer. Appl. Environ. Microbiol. 69:5884-91. Ashelford, K. E., N. A. Chuzhanova, J. C. Fry, A. J. Jones, and A. J. Weightman. 2005. At least 1 in 20 168 rRN A sequence records currently held in public repositories is estimated to contain substantial anomalies. Appl. Environ. Microbiol. 71:7724-36. Ashelford, K. E., N. A. Chuzhanova, J. C. Fry, A. J. Jones, and A. J. Weightman. 2006. New Screening Software Shows that Most Recent Large 16S rRNA Gene Clone Libraries Contain Chimeras. Appl. Environ. Microbiol. 72:5734-5741. Brodie, E. L., T. Z. Desantis, D. C. Joyner, S. M. Back, J. T. Larsen, G. L. Andersen, T. C. Hazen, P. M. Richardson, D. J. Herman, T. K. Tokunaga, J. M. Wan, and M. K. Firestone. 2006. Application of a high-density oligonucleotide microarray approach to study bacterial population dynamics during uranium reduction and reoxidation. Appl. Environ. Microbiol. 72:6288-98. Coates, J. D., D. J. Ellis, C. V. Gaw, and D. R. Lovley. 1999. Geothrix fermentans gen. nov., sp. nov., a novel Fe(III)-reducing bacterium from a hydrocarbon-contaminated aquifer. Int. J. Syst. Bacteriol. 49 Pt 4:1615-22. Cole, J. R., B. Chai, R. J. Farris, Q. Wang, A. S. Kulam-Syed-Mohideen, D. M. McGarrell, A. M. Bandela, E. Cardenas, G. M. Garrity, and J. M. Tiedje. 2007. The ribosomal database project (RDP-II): introducing myRDP space and quality controlled public data. Nucleic Acids Res. 35:D169-72. 55 10. 11. 12. 13. 14. 15. 16. Colwell, R. K. 2005. EstimateS: Statistical estimation of species richness and shared species from samples. Version 7.5. Persistent url . Fields, M. W., T. Yan, S. K. Rhee, S. L. Carroll, P. M. Jardine, D. B. Watson, C. S. Criddle, and J. Zhou. 2005. Impacts on microbial communities and cultivable isolates from groundwater contaminated with high levels of nitric acid- uranium waste. FEMS Microbiol. Ecol. 53:417-28. Finneran, K. T., M. E. Housewright, and D. R. Lovley. 2002. Multiple influences of nitrate on uranium solubility during bioremediation of uranium- contarninated subsurface sediments. Environ. Microbiol. 4:510-6. Cu, B., W. M. Wu, M. A. Ginder-Vogel, H. Yan, M. W. Fields, J. Zhou, S. F endorf, C. S. Criddle, and P. M. Jardine. 2005. Bioreduction of uranium in a contaminated soil column. Environ. Sci. Technol. 39:4841-7. Hazen, T. C., and H. H. Tabak. 2005. Developments in Bioremediation of Soils and Sediments Polluted with Metals and Radionuclides: 2. Field Research on Bioremediation of Metals and Radionuclides. Reviews in Environmental Science and Biotechnology V4: 157-183. Hemme, C. L., T. J. Gentry, L. Wu, M. W. Fields, C. Detter, K. Barry, D. Watson, C. W. Schadt, P. Richardson, T. Hazen, J. Tiedje, E. Rubin, and J. Zhou. 2006. Analysis of a Microbial Community Metagenome from Uranium- Contaminated Groundwater, 106th Gen. Meet. Am. Soc. Microbiol. May 21-25, 2006, Orlando, FL. Holmes, D. E., K. T. Finneran, R. A. O'Neil, and D. R. Lovley. 2002. Enrichment of members of the family Geobacteraceae associated with stimulation of dissimilatory metal reduction in uranium-contaminated aquifer sediments. Appl. Environ. Microbiol. 68:2300-6. Hua, B., H. Xu, J. Terry, and B. Deng. 2006. Kinetics of uranium(V I) reduction by hydrogen sulfide in anoxic aqueous systems. Environ. Sci. Technol. 40:4666- 71. Hwang, C., W. M. Wu, T. J. Gentry, J. Carley, S. L. Carroll, C. Schadt, D. Watson, P. M. Jardine, J. Zhou, R. F. Hickey, C. S. Criddle, and M. W. Fields. 2006. Changes in bacterial community structure correlate with initial 56 17. 18. 19. 20. 21. 22. 23. 24. 25. operating conditions of a field-scale denitrifying fluidized bed reactor. Appl. Microbiol. Biotechnol. 71:748-60. Istok, J. D., J. M. Senko, L. R. Krumholz, D. Watson, M. A. Bogle, A. Peacock, Y. J. Chang, and D. C. White. 2004. In situ bioreduction of technetium and uranium in a nitrate-contaminated aquifer. Environ. Sci. Technol. 38:468-75. Kelly, S. D., K. M. Kemner, J. Carley, C. Criddle, P. M. Jardine, T. L. Marsh, D. Phillips, D. Watson, and W.-M. Wu. 2008. Speciation of Uranium in Sediments before and after In situ Biostimulation. Environ. Sci. Technol. 42:1558-1564. Krieg, N. R., and J. G. Holt. 1984. Bergey's manual of systematic bacteriology. Williams & Wilkins, Baltimore. Kumar, S., K. Tamura, and M. Nei. 2004. MEGA3: Integrated software for Molecular Evolutionary Genetics Analysis and sequence alignment. Brief. Bioinform. 5: 150-63. Lovley, D. R., J. D. Coates, E. L. Blunt-Harris, E. J. P. Phillips, and J. C. Woodward. 1996. Humic substances as electron acceptors for microbial respiration. Nature:445-448. Lovley, D. R., E. J. P. Phillips, Y. A. Gorby, and E. R. Landa. 1991. Microbial reduction of uranium. Nature 350:413-416. Luo, J., W.-M. Wu, J. Carley, C. Ruan, B. Gu, P. M. Jardine, C. S. Criddle, and P. K. Kitanidis. 2007. Hydraulic performance analysis of a multiple injection-extraction well system. Journal of Hydrology 336:294. Luo, J., W. M. Wu, M. N. Fienen, P. M. Jardine, T. L. Mehlhorn, D. B. Watson, 0. A. Cirpka, C. S. Criddle, and P. K. Kitanidis. 2006. A Nested-Cell Approach for In Situ Remediation. Ground Water 44:266-274. Nanninga, H. J., and J. C. Gottschal. 1987. Properties of Desulfovibrio carbinolicus sp. nov. and Other Sulfate-Reducing Bacteria Isolated from an Anaerobic-Purification Plant. Appl. Environ. Microbiol. 53:802-809. 57 26. 27. 28. 29. 30. 31. 32. 33. 34. Nevin, K. P., K. T. Finneran, and D. R. Lovley. 2003. Microorganisms associated with uranium bioremediation in a high-salinity subsurface sediment. Appl. Environ. Microbiol. 69:3672-5. Nevin, K. P., and D. R. Lovley. 2000. Potential for Nonenzymatic Reduction of Fe(III) via Electron Shuttling in Subsurface Sediments. Environ. Sci. Technol. 34:2472-2478. North, N. N., S. L. Dollhopf, L. Petrie, J. D. Istok, D. L. Balkwill, and J. E. Kostka. 2004. Change in bacterial community structure during in situ biostimulation of subsurface sediment cocontaminated with uranium and nitrate. Appl. Environ. Microbiol. 70:4911-20. Nyman, J. L., T. L. Marsh, M. A. Ginder-Vogel, M. Gentile, S. Fendorf, and C. Criddle. 2006. Heterogeneous response to biostimulation for U(VI) reduction in replicated sediment microcosms. Biodegradation V17 :303-316. O'Loughlin, E. J., S. D. Kelly, R. E. Cook, R. Csencsits, and K. M. Kemner. 2003. Reduction of uranium(VI) by mixed iron(II)/iron(III) hydroxide (green rust): formation of U02 nanoparticles. Environ. Sci. Technol. 37 :721-7. Petrie, L., N. N. North, S. L. Dollhopf, D. L. Balkwill, and J. E. Kostka. 2003. Enumeration and characterization of iron(III)-reducing microbial communities from acidic subsurface sediments contaminated with uranium(VI). Appl. Environ. Microbiol. 69:7467-79. Riley, R. G., and J. M. Zachara. 1992. Chemical contaminants on DOE lands and selection of contaminant mixtures for subsurface science research. US. Department of Energy, Washington, DC. Robertson, W. J., J. P. Bowman, P. D. F ranzmann, and B. J. Mee. 2001. Desulfosporosinus meridiei sp. nov., a spore-forming sulfate-reducing bacterium isolated from gasolene-contaminated groundwater. Int. J. Syst. Evol. Microbiol. 51:133-40. Sanford, R. A., J. R. Cole, and J. M. Tiedje. 2002. Characterization and description of Anaeromyxobacter dehalogenans gen. nov., sp. nov., an aryl- halorespiring facultative anaerobic myxobacterium. Appl. Environ. Microbiol. 68:893-900. 58 35. 36. 37. 38. 39. 40. 41. 42. 43. Schloss, P. D., and J. Handelsman. 2005. Introducing DOTUR, a computer program for defining operational taxonomic units and estimating species richness. Appl. Environ. Microbiol. 71: 1501-6. Schloss, P. D., B. R. Larget, and J. Handelsman. 2004. Integration of microbial ecology and statistics: a test to compare gene libraries. Appl. Environ. Microbiol. 70:5485—92. Senko, J. M., J. D. Istok, J. M. Suflita, and L. R. Krumholz. 2002. In-situ evidence for uranium immobilization and remobilization. Environ. Sci. Technol. 36: 1491-6. Singleton, D. R., M. A. Furlong, S. L. Rathbun, and W. B. Whitman. 2001. Quantitative Comparisons of 168 rRNA Gene Sequence Libraries from Environmental Samples. Appl. Environ. Microbiol. 67:4374-4376. Sung, Y., K. E. Fletcher, K. M. Ritalahti, R. P. Apkarian, N. Ramos- Hernandez, R. A. Sanford, N. M. Mesbah, and F. E. Loffler. 2006. Geobacter lovleyi sp. nov. strain 82, a novel metal-reducing and tetrachloroethene- dechlorinating bacterium. Appl. Environ. Microbiol. 72:2775-82. Suzuki, Y., S. D. Kelly, K. M. Kemner, and J. F. Banfield. 2004. Enzymatic U(VI) reduction by Desulfosporosinus species. Radiochimica acta 92:11-16. Suzuki, Y., S. D. Kelly, K. M. Kemner, and J. F. Banfield. 2003. Microbial populations stimulated for hexavalent uranium reduction in uranium mine sediment. Appl. Environ. Microbiol. 69:1337-46. Tabak, H. H., P. Lens, E. D. van Hullebusch, and W. Dejonghe. 2005 . Developments in Bioremediation of Soils and Sediments Polluted with Metals and Radionuclides - 1. Microbial Processes and Mechanisms Affecting Bioremediation of Metal Contamination and Influencing Metal Toxicity and Transport. Reviews in Environmental Science and Biotechnology V4:115-156. van der Meer, J. R., R. I. Eggen, A. J. Zehnder, and W. M. de Vos. 1991. Sequence analysis of the Pseudomonas sp. strain P51 tcb gene cluster, which encodes metabolism of chlorinated catechols: evidence for specialization of catechol 1,2-dioxygenases for chlorinated substrates. J. Bacteriol. 173:2425-34. 59 44. 45. 46. 47. 48. 49. 50. 51. Vrionis, H. A., R. T. Anderson, I. Ortiz-Bernad, K. R. O'Neill, C. T. Resch, A. D. Peacock, R. Dayvault, D. C. White, P. E. Long, and D. R. Lovley. 2005. Microbiological and geochemical heterogeneity in an in situ uranium bioremediation field site. Appl. Environ. Microbiol. 71:6308-18. Wall, J. D., and L. R. Krumholz. 2006. Uranium reduction. Annu. Rev. Microbiol. 60: 149-66. Wan, J., T. K. Tokunaga, E. Brodie, Z. Wang, Z. Zheng, D. Herman, T. C. Hazen, M. K. Firestone, and S. R. Sutton. 2005. Reoxidation of bioreduced uranium under reducing conditions. Environ. Sci. Technol. 39:6162-9. Wu, Q., R. A. Sanford, and F. E. Loffler. 2006. Uranium(VI) reduction by Anaeromyxobacter dehalogenans strain 2CP-C. Appl. Environ. Microbiol. 72:3608-14. Wu, W. M., J. Carley, M. Fienen, T. Mehlhorn, K. Lowe, J. Nyman, J. Luo, M. E. Gentile, R. Rajan, D. Wagner, R. F. Hickey, B. Gu, D. Watson, 0. A. Cirpka, P. K. Kitanidis, P. M. Jardine, and C. S. Criddle. 2006. Pilot-scale in situ bioremediation of uranium in a highly contaminated aquifer. 1. Conditioning of a treatment zone. Environ. Sci. Technol. 40:3978-85. Wu, W. M., J. Carley, T. Gentry, M. A. Ginder-Vogel, M. Fienen, T. Mehlhorn, H. Yan, S. Carol], M. N. Pace, J. Nyman, J. Luo, M. E. Gentile, M. W. Fields, R. F. Hickey, B. Gu, D. Watson, 0. A. Cirpka, J. Zhou, S. Fendorf, P. K. Kitanidis, P. M. Jardine, and C. S. Criddle. 2006. Pilot-scale in situ bioremediation of uranium in a highly contaminated aquifer. 2. Reduction of u(VI) and geochemical control of u(VI) bioavailability. Environ. Sci. Technol. 40:3986-95. Wu, W. M., J. Carley, J. Luo, M. Ginder-Volgel, E. Cardenas, M. B. Leigh, C. Hwang, S. D. Kelly, C. Ruan, L. Wu, J. van Nostand, T. Gentry, K. Lowe, T. Melbourne, S. Carroll, M. W. Fields, B. Gu, D. Watson, K. M. Kemner, T. L. Marsh, J. M. Tiedje, J. Zhou, P. K. F endorf, P. K. Kitanidis, P. M. Jardine, and C. S. Criddle. 2007. In situ bioremediation of uranium (VI) to submicromolecular levels and reoxidation by dissolved oxygen. Environ. Sci. Technol. 41:5716-5723. Wu, W. M., B. Gu, M. W. Fields, M. Gentile, Y. K. Ku, H. Yan, S. Tiquias, T. Yan, J. Nyman, J. Zhou, P. M. Jardine, and C. S. Criddle. 2005. Uranium (VI) Reduction by Denitrifying Biomass. Bioremediation 9:46-61. 60 52. Yamada, T., Y. Sekiguchi, S. Hanada, H. Imachi, A. Ohashi, H. Harada, and Y. Kamagata. 2006. Anaerolinea thermolimosa sp. nov., Levilinea saccharolytica gen. nov., sp. nov. and Leptolinea tardivitalis gen. nov., sp. nov., novel filamentous anaerobes, and description of the new classes Anaerolineae classis nov. and Caldilineae classis nov. in the bacterial phylum Chloroflexi. Int. J. Syst. Evol. Microbiol. 56:1331-40. 61 CHAPTER THREE Massive parallel sequencing of microbial communities in a uranium remediation site INTRODUCTION Metals and radionuclides are two major contaminants at Department of Energy (DOE) facilities and waste sites (45). One of these sites, the former Y-12 National Security Facility at Oak Ridge, Tennessee, stored its waste from uranium enrichment operation in four unlined ponds of 9.5 million liter capacity each for over 31years. The ponds received acidic uranium nitrate (~30% uranium nitrate) from its uranium enrichment operations as well as waste from other DOE sites. After 31 years of operations the ponds were neutralized, subject to nitrate removal, and its water treated for nitrate removal before the ponds were capped (the contaminated sludges remained). The contamination, however, was not contained within the ponds but expanded to nearby areas through groundwater movement. The nearby Area 3 contains uranium in concentrations as high as 250 uM in groundwater and 800 mg/kg in sediments (8, 58). Since metals cannot be degraded, other strategies are needed to reduce their bioavailability. One of these alternatives is microbial dissimilatory reduction, an anaerobic process in which microbes use metals such as Fe(III), Mn(IV), Cr(VI) , Se(VI) and radionuclides such as U(VI), Tc(VII) as final electron acceptors to support their growth. Metal reduction involves changes in oxidation state than can trigger changes in water solubility making some metals more soluble and easier to remove from the system (e.g. Mn), and others less soluble, hence less mobile and less bioavailable, like in the 62 case of uranium. Since uranium reduction was first discovered it has been proposed as an alternative to control U bioavailability (2, 3, 23, 29, 34, 58). The ability to reduce U(VI) to U(IV) was first shown in the Fe(II) reducing bacteria (FRB) Geobacter metallireducens (34) and later in the sulfate- reducing bacteria (SRB) Desulfovibrio desulfuricans (33). Since then, U(VI) reducers have been found in a variety of phylogenetic groups such as Delta-, Beta— and Gamma-Proteobacteria, Firmicutes, Deinococci, Actinobacteria among others groups (55). Beside the ability to reduce uranium, there is no common theme among uranium reducing bacteria as they represent metabolically different groups such as FRB, SRB, ferrnenters, and denitrifiers. Uranium reduction for bioremediation purposes has been tested successfully with pure cultures of Fe(III)-reducing bacteria such as Geobacter spp.; sulfate-reducing bacteria such as Desulfovibrio, Desulfoporosinus, Shewanella, and Desulfotomaculum spp.; ferrnenters such as Clostridium; heterotrophs such as Deinococcus spp. and denitrifying bacteria such as Acidovorax spp., among others (55). More recently, batch serum bottles (18, 25, 61), microcosms (43), and sediments columns (20, 56) have been used to study this approach in controlled conditions. Field scale studies have recently started at a former uranium ore processing facility (Old Rifle site, now part of the Uranium Mill Tailings Remedial Action program of the DOE) at Rifle, Colorado (54), and at the Field Research Center (FRC) at Oak Ridge, Tennessee, the site containing residue from the former Y-12 uranium enrichment facility (3, 29, 58). Studies of uranium reduction have shown response of F RB and/or SRB, though the stimulatory conditions were different. Geobacter was shown to be abundant at the 63 Old Rifle site, where uranium reduction was promoted through acetate injection into the subsurface (3, 25, 42, 54). A different study also at the Old Rifle Site found enrichments of FRB (Geobacter spp.) and also SRB (Desulfobacterales) (54). A recent publication from this site has shown that uranium removal from groundwater continued even after acetate injection stopped due to uranium absorption into biomass of F irmicutes-related bacteria (40). Though not expected to be uranium reducers but dead-cell recyclers, their effect in controlling uranium bioavailability is real and promising. In field studies from the FRC Area 3 where ethanol was used as stimulatory electron donor, Desulfovibrio spp. and Geobacter spp. have been associated to the uranium (10, 27); and Geobacter and Geothrix have been found in enrichments using Area 2 samples with lactate as electron donor (6). In anaerobic enrichments of contaminated sediments of a uranium mine, sequences related to the SRB Desulfosporosinus, and fermentative Clostridium (53) appeared to be major contributors to U(VI) reduction. This present work studies the approach taken at the F RC Area 3 for uranium and nitrate bioremediation. The bioremediation strategy was based on a hydraulic control system that creates an area, isolated from groundwater intrusion and which allows electron donors to be injected into the system to promote growth of local communities of uranium reducers. The system was created using a series of wells that pump and recirculate the groundwater. Over a two year period, the treatment was successful in removing nitrate and conditioning the subsurface for the establishment of uranium reducing communities (58). After conditioning, ethanol was injected and uranium levels in groundwater dropped. X-ray Absorption Near Edge Structure (XANES) confirmed that uranium was being reduced, discarding the possibility that uranium removal from water 64 was due to sorption or complexation with inorganic compounds (59). Microbial surveys of sediments from six wells from Area 3 detected genera known to harbor U(VI)- reducing members such as Geobacter, Desulfovibrio, Aneromyxobacter, Desulfosporosinus, and Acidovorax after uranium reduction was established in the zone (10) . Microbial counts, measured as Most Probable Number, correlated with the hydraulic path; however, the proportion of U present as U(IV) did not follow a linear relationship with the hydraulic connection (10). This suggests that factors other than electron donor availability are influencing the presence and/or activity of the uranium reducing populations. A more recent study tracked the microbial communities of F RC Area 3 groundwater over a 1.5 year period and found that environmental factors such as nitrate, uranium, sulfide, and ethanol were correlated with particular bacterial populations (27). More interestingly, the engineering adjustments to control dissolved oxygen and deliver nutrients to the subsurface were also found to be significant in explaining the biological variability (27). In this study we explored the role of environmental conditions, including hydrology, on the structure of the sediment microbial communities of FRC Area 3 in a period of high uranium reduction. The use of sediments allows for a more precise spatial characterization than using groundwater samples where filtering large volumes of water is required because of low biomass. Additionally, groundwater samples may not reflect the attached communities present in sediments, especially in oligotrophic aquifers such as the FRC (22). For this purpose we thoroughly surveyed the microbial communities present in the area to include a wider range of environmental conditions. We used massive parallel 65 16S rRNA gene sequencing to obtain a deeper survey of the microbial communities, and higher confidence in the detection indicator species. We were able to detect groups associated with uranium reduction, to explain differences in community structure with environmental factors, and to manage with high confidence the massive number of sequences to provide candidate groups likely important for the bioremediation of the Area 3. MATERIAL AND METHODS Site description and sampling. Samples were retrieved from Area 3 of the DOE F RC at the Y-12 National Security Complex, Oak Ridge, Tennessee, USA. In this area, a hydraulic control system was used to promote in situ bioremediation in a controlled fashion. The system consisted of an outer groundwater recirculation loop (injection well F W024 and extraction well FW103) that isolated an inner groundwater loop (injection well F W104 and extraction well FW026) preventing penetration by highly contaminated groundwater from the source zone (Figure 3.1). Well F W105 was used to feed water to the above ground treatment system and provided additional hydraulic control to the outer loop (58). The outer loop was expected to respond to changes in groundwater movement, while the inner loop’s shape and flow was expected to remain stable (5 8). The remediation started on August 23th, 2003 (day 0). During the first 137 days, groundwater was pumped and treated ex situ to adjust pH, remove Al, Ca and nitrate (via a denitrification bioreactor). This process removed nitrate competition as a terminal electron acceptor, avoided formation of stable Ca-U-CO3 products and clogging by gas and biomass due to denitrification. Reduction of U(VI) to U(IV) was accomplished 66 through ethanol injection to the inner loop, starting on day 137. Multilevel sampling (MLS) wells, F W100, F W101 and F W102 were used to monitor hydrogeology and remediation performance over seven levels (-1 to -7). Except for the multisarnpling wells, the wells were 14.6 m depth with screens between 11.28 m and 13.77 m. Multisampling wells retrieved sediments from 15.2 , 13.7, 12.2, and, 10. 7 m for levels -1, -2, -3, and -4, respectively. Water and sediments samples were retrieved on October 5th, 2005(day 775) as previously described (10). FW101 Groundwater I flow f FW026 I O FW105 fl \ FW103 3 . Multilevel sampling wellsL Milfilfi , O 0 Well 1L 2 ;- Figure 3.1. Scheme of the DOE Field Research Center Area 3 treatment system. Wells FW104 and FW024 are injection wells, FW026,F W105 and F W103 are extraction wells. Wells F W100, F W101, and F W102 are multilevel sampling wells. The controlled hydraulic scheme creates an outer cell (2) that protects an inner cell (1) from intrusion of contaminated groundwater from the background (3). The stimulatory electron donor, ethanol, is injected through FW104 into the inner cell. The terms inner/outer loop and inner/outer cells are used interchangeably throughout this article. Chemicals and analytical methods. Chemical oxygen demand (COD), sulfide 67 and F e(II) were determined using a Hach DR 2000 spectrophotometer (Hach Chemical, Loveland, CO). Anions (including NO3-, Br-, Cl-, SO42- and P043) were analyzed with an ion chromatograph as previously described (10). Metals (Al, Ca, Fe, Mn, Mg, U and K etc.) were determined using an inductively coupled plasma mass spectrometer (ICPMS) (Perkin Elmer ELAN 6100), and U reduction state was determined with XANES as described elsewhere (59, 60). The proportion of Fe(II) to the total Fe was measured as HCl (10%) extractable. Two hydrology parameters that described the connectivity of the different locations to the injection well were measured with a tracer study that introduced a conservative tracer (bromide) at the injection well (FW104) were previously determined (3 5). Tracer recovery (%) was considered a proxy for the amount of the stimulatory electron donor received at each location. Similarly, mean travel time (from FW104) was used as proxy for the composition of the electron donor as ethanol is converted to acetate by the microbial communities. DNA extraction and direct sequencing. DNA was extracted from 0.5 g of sediments with the Fast soil prep kit (MoBio Inc., San Diego, CA, USA) following the manufacturer’s instructions. Ribosomal RNA genes were amplified using a primer set that flanked the v4 hypervariable region of the 16S rRNA gene at corresponding Escherichia coli positions 563 and 802. Primers 563F (5’- GCCTCCCTCGCGCCATCAG(barc0de)AYTGGGYDTAAAGVG-3’) and 802R (5’- GCCTTGCCAGCCCGCTCAGTACNVGGGTATCTAATCC-3’) were a combination of universal eubacterial primers and sequence adaptors (in bold) used in the amplification 68 step of the 454 Roche pyrosequencing process. The forward primer additionally contained a short run of nucleotides used as bar codes. The bar codes were designed to be distinct from each other by at least two nucleotide changes. Primers were dual HPLC-purified (Integrated DNA Technologies, Coralville, IA). Each PCR mixture contained 1 uM of each primer (dual HPLC-purified, Integrated DNA Technologies, Coralville, IA), 1.8 mM MgC12, 0.2 mM for each dNTP, 3 ug BSA (New England Biolabs, Beverly, MA), 1 unit of F astStart High Fidelity PCR System enzyme blend (Roche Applied Science, Indianapolis, IN), and 10 ng of DNA template. Each sample was run in triplicate. The PCR cycling conditions were: 95°C for 3 min; 95 °C for 453, 57 °C for 455, 72°C for 1min, for 30 cycles; 72 °C for 4 min. PCR products were run in a 1% (wt/vol) agarose Tris-acetate-EDTA gel and bands between 270-300 bp were excised. Bands for each triplicate were pooled together for gel purification with the QIAquick Gel Extraction Kit (Qiagen Inc., Valencia, CA, USA). After gel extraction, products were cleaned for a second time with the QIAquick PCR purification kit (Qiagen 3 Inc., Valencia, CA, USA) and eluted with 20 ul of EB buffer (10 mM Tris-Cl, pH 8.5). Clean products were quantified using the Nanodrop ND-IOOO spectrophotometer (N anoDrop Technologies, Wilmington, DE) and mixed in equal amounts, 20 mg each for direct sequencing by Genome Sequencer FLX System (454 Life Sciences). Pyrosequencing analysis. Sequences were processed with the Ribosomal Database Project (RDP) Pyrosequencing Pipeline (14). Sequences were first trimmed to remove the pyrosequencing adaptor sequences, and then sorted according to their bar 69 codes (used in the PCR amplification). Two quality filters were applied to remove sequences with aberrant lengths and to discard sequences with more than two changes in the primer portion. The RDP also provided a complete linkage clustering tool and secondary-stucture—based alignment tool based on INFERNAL version 8.1 (http://inl‘emal.ianeliaorgf) and a 16S rRNA secondary structure model (9). Community analysis. Sequences from different samples were classified using RDP classifier with a 80% bootstrap confidence (57). To rapidly separate sequences into taxonomic groups, RDP’s Classifier was used at a 50% confidence. e.g. to obtain all F irmicutes from sample A. Sequences from that particular taxon from different samples were later aligned and clustered (complete linkage) in order to track the distribution of a specific group among samples. Operational taxonomic units (OTUs) were defined from 0% to 20% sequence dissimilarity (distance) to calculate richness of individual samples at different levels of taxonomy. The results were rarified to the lowest number of sequences per sample from the group of samples being compared. When an individual phylum or class was studied, OTUs were defined at 5% distance (~genus level). When an individual genus was studied, OTUs were defined at 1% distance (species level). To compare individual samples according to their diversity patterns (beta diversity analysis), a modified Sorensen index was used. This index has been corrected by using abundance information to account for unseen species in the survey (11). Clustering of samples was done using the average neighbor clustering algorithm on the 70 modified Sorensen index matrix. To find groups that represent a specific branch of the clustering tree, indicator species analysis was used on the resulting topology (16). This method was preferred because it provides a consistent approach to analyze the high number of sequences per sample. The indicator value algorithm considers both frequency of occurrence, and relative abundance, and its higher when a species is present in all the samples and highly abundant in one specific group (16), this avoids considering indicator species groups that appear only on one location. Since we test each OTU for significance in the indicator species analysis, and multiple tests could increase the chance of finding significant results by chance, we used a false discovery rate (FRD) approach (5). The FDR estimates the chance of reporting a false positive in all the significant results (q- value) instead of using excessively conservative corrected p-values. Q values were generated from p-values using the freely available software package QVALUE Ver. 1.0 (51, 52) using the bootstrap method. Multivariate analysis. Environmental variables were standardized (z-score) and tested for correlation using a principal component analysis (PCA) on their correlation matrix. This result was used to find variables correlated to uranium reduction, sulfate reduction and iron reduction, to create a model on how the treatment changes the geochemical characteristics, and to select environmental for direct gradient analysis. For direct gradient testing, environmental variates were normalized using a log(x+1) or arcsin(\/x) transformations, pH was left unchanged. Hellinger transformation was used in some cases on OTUs abundance data to make it suitable even if when many zeros were present (32). Principal coordinates analysis (PCoA) and non-dimensional scaling, was performed on untransformed species data. 71 Statistical analyses were conducted using the R environment v2.80 (http://www.R-project.0rg.) with packages Vegan v1.8-8 (50), BiodiversityR (13), Labdsv v1.3-1 (46) (indicator species analysis), and QVALUE v1.1 (available at http://genomics.princeton.edu/storeylab/qvalue/) (51, 52). RESULTS Chemical and hydrological characterization. Analysis of DOE FRC Area 3 hydrology (Table 3.1), groundwater (Table 3.2) and sediments (Table 3.3) showed differences in geochemical and hydrological composition due to the creation of the three zones (inner loop, outer loop and background) by the treatment system. One of the hydrology parameters, percentage of tracer recovery, an indication of how much water from the injection well FW104 reached to a particular well, was significantly higher in the inner loop (p<0.01) than in the outer loop. Mean travel time, the second hydrology parameter and an indication on how long does it takes groundwater to move from the injection well FW104 was not significantly different in the two loops. When grouping the two multilevel sampling (MLS) wells from the inner loop (FW101 and FW102) according to the depth they sampled, wells at level -1 and -4 received significantly less water from FW104 (p<0.05), and were less connected (p<0.001) than wells from levels - - 2 and -3. This is likely because the injection at F W1 04 occurred at depths between levels -2 and -3. When compared to the outer loop, inner loop groundwater was richer in sulfate (1K0 - 01), Fe (p<0.01), K (p<0.01), and chloride (p<0.05); and more depleted in sulfide (1K0 - 05). This trend remained when compared to the background area (FW105) though 72 no statistical comparison was possible (only one sample from background well existed). Samples from depths 2 and 3 of the inner loop were significantly higher in K (p<0.01), and Fe (<0.05) and lower in chloride (p<0.05) than samples from levels 1 and 4. Sediment analysis showed reduced U(IV) only in wells from the inner loop, and a higher proportion of reduced Fe(II) in the inner loop (p=0.08). Sulfide levels of sediments were also higher in the inner loop, although not significantly. These three trends remained when adding the background well FW105, though no statistical test was possible. Overall, the bioremediation system seems to have influenced the composition of the groundwater by creating gradients in electron donor availability (due to differences in connectivity) and maintaining the isolation of the inner loop from contaminated background groundwater. The differences found can be related to the activity of SRB and F RB as observed in changes in sulfate/sulfide and total Fe/reduced Fe(II). 73 Table 3.1. Hydrological characterization of the DOE FRC Area 3 wells. Tracer studies results Well* Mean travel time Tracer recovery (hr) (%) FW104 0 100 FW101-1 387 18 FW101-2 2.9 93 FW101-3 17.9 60 FW101—4 298 18 FW102-l 222 3 FW102-2 1 1.6 94 FW102-3 3.7 94 F W102-4 790 6 FW026 8 1 50 "#07024 """"""""""""""""" 7345 """"""" FW100—1 21 1 9 F W100-2 223 12 F W100-3 43 18 FW100-4 211 8 "Ell/.193. __________________________________ 7-8 ......................... 1.9 .............. FW105 710 6 Inner loop (meand: SE.) 174.1 :h 82.3 53.6 i 12.6 Outer loop (meant SE) 1 17.3 d: 44.1 10.3 :1: 1.9 Background 710 6 P-value* * 0.622 0.008 Results from hydrology studies using bromide (conservative tracer) on days 801 to 802 (November 1St to 2nd, 2005). Tracer was injected at well F W104, tracer recovery and travel time at this well are 100% and 0 h by definition. The treatment system created three different areas: inner loop (FW104, FW026 and multilevel wells F W101 and FW102), outer loop (FW024, FW103 and multilevel well FW100), and a local background (represented by FW1 05).*F or multilevel sampling wells the well location is followed by the depth where the sample was taken. * *Comparison of inner and outer loops was done with a two-tailed t-Test, alpha=0.05. 74 42 ~26 83 a; mac who a? 33 -.~ 8; so A: 8.4 8;: who -o.o 33 ~55 :3 a3 mn~ 83 :.~ Sod So _2 ~qm SSE of Sod 82. 83 :3 ~26 -.~ 82 e~.~ 32 8d a: 9% «.825 Ed Rod 83 :3 as 22 $~ 83 -.~ -o.o cod 9: an 22.5 who Sad 83 ~23 moo wed 8.~ 88d wm~ Sod cod 42 com woes/a 3m ~23 Sod ~23 86 3o 32 m~qo m~.~ 83 cod m~._ Se 33: a; Ego Sod 23 $3 26 $~ ~85 ~m.~ _ 5d 23 w? :3 48.5 85 ES 89o 23 ~26 3o 3~ m~no 82 Sod 36 o~._ Ed 89?. go 985 ~85 83 _~.o who :1 5% 3..~ ~56 cod m~._ 3.... 4-823... ~90 $3 Sod ~23 .3 m2 ~o.~ 53 :.~ 53 one o: as ~-~o§m 23 ES 98¢ :36 ad a: 8; $3 ~m.~ _ :3 e _ .c :2 as 32.5 2.2 $3 :3 23 -.~ 2.0 a; an; 3~ 23% cod 3o £8 78;: $5 5.3 83 83 was 93 :.~ 83 a..~ 83 85 :1 3m 4-8.3.2 $5 Sod mood ~35 m2 :3 o~.~ _ _ 3 on~ 83 :.o 8.. Se 32.5 $5 E3 486 ~55 one on: ~o.~ 9.3 -.~ Sod as 2: m3 323... om.~ Sod wood ~55 was 93 ~02 $3 ~v.~ o~no cod n3 who 78;: 85 ~86 ES :3 ~26 3o S.~ $3 v~.~ coed one w: 26 3;: 8 :2 on 2 m2 v. «z a -6 32.2 8.2.5 Beam 2; .133 N225 533.6555 mo :oEmomEoU 325.20 .m 8:. um... won :3... £2. 8...: 3:55....» .3 858388 .3255 .~.m 03.. s 75 .modumna—m 48,—L 3:8 -95 a was: 25¢ 33 302 8:5 25 55: mo camtmmfioouw * dew—8 $3 29:8 2: 20:3 fiaoc 05 an 330:8 2 5:80. :03 2: 233 wEEEmm 35:22: 8m * A3. 3m 3 338835 vcsohwv—omn was .89 3m :95 ESE—:8 can 8:5”— .vmoamv 82 530 Aoomam Ea 9:23 m=oB 65:22: ES amok/n— .vogzmv Q02 55: ”£8“ “:80ng 085 @888 88?? Sesame; .Eq 38x3 2:. 5 Ba mcosmbcoocoo .Amoom .Em 89809 SE mac no 3552 803 838nm wnmd cued w 56 _Nwd Sod mood god 026 mmod Cvd god mood omod **o=_w> -m S: A as .o coed mum ~ .0 mmd mmd 3N; nwmmd voomd So. .N c Q..— 86 cczewxomm To 36 o de vod mod Ed ommd Nod womd cod mod Ed Amm HSBEV H H H H H H H H H H H H H Q8— 330 mm; vod mood Rod mvd amd Sum $6 and Nomd cod 9: vafi mm; god mood Sod mud Ed wmd .36 mod mmva 36 _d _md Amm HamoEV H H H H H H H H H H H H H 82 BE: S.N owed fled _mod 36 wwd Wm cwwd own mama 36 no; 36 «0 £2 on _< 32 M «Z D -5 @8me 0325 Sufism in *5? 92:1? $339580 mo $583.00 32825 )/ Q. .53 .~.~ 2...; 76 Table 3.3. Chemical composition of sediments from DOE FRC Area 3. Chemical Characteristics of Sediments U 1v / Total Fe 11 / We11* Bioactivity (mg/kg) To(tal)U Fe T6191 )Fe (cg/(111;) fig: (%) (g/kg) (%) FW104 High 10.3 61 199.1 0.53000 337.9 23.000 FW101-1 Low 0.051 0 48.3 0.02640 8.6 0.000 FW101-2 High 1.249 54 43.3 0.27400 23.9 0.667 FW101-3 High 1.825 51 29.4 0.29400 24.2 0.265 FW101-4 Low 0.619 0 42.6 0.04660 27.1 0.000 FW102-1 Low 0.041 0 39.4 0.00560 22.6 0.000 FW102-2 High 0.517 17 33.7 0.30900 20 0.557 FW102-3 High 0.875 30 36.4 0.28700 22.5 0.431 FW102-4 Low 0.481 0 37.1 0.00660 12.7 0.000 "If-W939 ............. 9.4.3.9193? ......... L??- .......... 9 ......... 9121.-----9219999 ...... 9.2.27. ...... 99.3.9--- FW024 Low 0.371 0 50.8 0.21900 33.2 0.070 FWIOO-l Low 0.215 0 55.6 0.00142 10.5 0.000 FW100-2 Medium 0.98 0 32.4 0.12640 7.5 0.000 FW1 003 Medium 1.098 0 36.2 0.06360 13.4 0.000 FW100-4 Low 1.501 0 34.4 0.05450 23.8 0.000 “Ii-“(ESP ............... 9.9-“Y .......... 9.26.9.3 .......... 9 ......... 9i-2-----9.-.99§1.9 ...... 3.2.-.1 ....... 9.9-99--- FW105 Low 0.978 0 47.8 0.00480 30.7 0.000 Inner loop 1.718 21.3 55.6 0.19272 52.2 2.495 (mean :1: SE) i i i i i i - 0.97 8.1 16 0.05515 31.8 2.28 Omr 100p 0.8 42.4 0.07834 20.1 0.012 (mean t SE) i i i i i - 0.197 0 3.9 0.03374 4.6 0.012 Background - 0.978 0 47.8 0.0048 30.7 0 P-value* * - 0.377 ND 0.441 0.157 0.343 0.304 Samples retrieved on day 774 (October 5th, 2005). Treatment system created three different areas: inner loop (FW104, F W026 and MLS FWIOO and FW200), outer loop (FW024, F W103 and MLS FW100), and a local background (represented by FW105). *F or multilevel sampling wells the well location is followed by the depth where the samme was taken. * *Comparison of inner and outer loops with a two tailed t-Test, a1Pha=0.05.ND, mathematically not defined. Bioactivity is a qualitative descriptor based on Most Probable Number (MPN) results for iron reducing bacteria, sulfate reducing bacteria and denitrifiers. This variable is used more as a general guide rather than a Strictly quantitative indicator. The MPN methodology is described elsewhere (10). 77 Correlation of environmental variables. Correlation analysis showed that the environmental variables measured were highly correlated. PCA on the correlation matrix (Figure 3.2) showed that most of the variables were collinear with at least one other variable as shown in the correlation matrix (Table 3.4). Hydrology factors (mean travel time and tracer recovery) seemed to influence key variables such as pH. Mean travel time was correlated with pH, A1 and U in groundwater. It is not clear if U levels in groundwater were influenced by pH directly or controlled by microbial activity. The percentage of tracer recovery was correlated with sulfide levels in groundwater, potassium and iron in groundwater. Reduced uranium in sediments (U(VI)/Total U) was strongly correlated with bioactivity and reduced iron and moderately correlated with sulfide levels in sediments, uranium levels in sediments, iron in sediments and chemical oxygen demand (COD). Nitrate was highly correlated (r>0.9) with Na, chloride, Mg, Mn, and Ca. Nitrate was also correlated with chloride and negatively with sulfate but with smaller strength (F 0.72 and -O.72 respectively). The correlation of reduced U(IV) with variables related to FRB and SRB and not with nitrate reduction suggests that FRB and SRB are more important for U(VI) reduction at this location than nitrate reducing populations. Based on the correlation analysis, the following variables were kept: pH, uranium in sediment, tracer travel time, nitrate, and COD in sediments. 78 01-7 Sulfate bu P U(Sed) "‘ Fe(sed) Sulfide(Sed) COD(Se ) w 9 Red.U(Sed) .c-J Red.Fe(Sed) ° Bioactivity - Sulfide (Gw) . -° Tracer recovery Chloride U" i Nitrate Ca Mg I l i 1 1 -1.0 -O.5 0.0 0.5 1.0 1.5 PCI Figure 3.2. Correlation plot of environmental variables measured for 15 locations at DOE FRC Area 3. Variables were z—score transformed prior to correlation analysis. Bioactivity was converted as following, High=3, Medium=2 and Low=1. 79 ts and imen f groundwater, sed lcharactenstics 0 Table 3.4 Correlation of chemica hydrology of the DOE FRC Area 3 wells. .0269: Bow:- .mm 053 $295 :60: . _ N :83 6.653. .8 $63 moo .2 :83 tom museum .2 sea em .2 663 .5 Bose-é .e. :63 3 .2 $3385 .3 6 .2 :2 .2 is em .: _< .2 w: .6 v. L NZ 5 9,3 2 .e 625.6 .m 35:2 .v 633 .m 63:3 .~ :d ._ 03min? dmd vod- vod- odd- 2d Nod- vmd mod- :d mmd 2d hmd end- dmd _ _d _md 5.? mod- dmd mmd mmd- dd.— 1 mmd- N. d mdd- _dd- odd- edd- mNd- :3 Ed- mud- Ed- 3d- mvd dud- de- vod- nmd mod- 2.d- mmd- dd.— dad mvd- dvd _vd _vd mmd ovd mvd _md mmd- mod and Ed- 3d- and m _ d- de- w _ d- w _ d- dd.— emd- add mod- N. d- de- med- 2 d- m _ d- mmd- wad Nod ddd fidd- odd Ed- 3d 5d- mud dd. _ cmd- vmd Ndd- fidd- bod mod- N_ d wdd- N _ d mod med mod- _dd- wed o _ d- mud dd 8. _ dvd- Ndd add- add- mmd- udd- mmd- hod- mmd- odd- 3d 2d- mwd d. d- _ _d- de- dd.— dmd- _Nd- mdd- add- udd- Nod- _ _d- Ed- 2d- wwd bud Nod Rd- mod mmd- do.— and mmd- Ed 2d 2d 3d oNd m2... 2d 3d- mod- 3d 3 d- und- 8.. dmd- and. and cod dmd- 3d and cvd- odd bdd add- :d- mmd ddd 2d- N_d- dod- «.md Ndd- Ed- de- 3.? «_d 2d- mmd- Ndd- :d- and Vdd Nod- Ed- 2d- dmd Vdd- N_d- 2d- hod- _md vod- :d- mmd- de- odd Ed- mmd- odd Cd- mdd- dwd dd.— wwd vod mmd dd.— Ndd- 3.? dc.— 2d- do.— do.— 9 10 l l 1 2 l3 80 Table 3.4 (cont’d) 5‘68. sum; .mm 9:: _o>m.= cam—2 .5 3me 853m .8 €me 80 .2 8% ..£ 803% .3 :3 ma .2 €me .2 “.838”. .9 sad 2 .9 €885 .3 8 .9 :2 .NF is 8 .2 _< .2 92 .m x .Q «2 s is a Q $8.5 .m 222 L 853 .m 03.5 .N In d 03a€m> cmd mvd vvd wmd 3d ovd mvd de- dmd- 2d- wmd wad mmd mod mad dud mmd 3d mad 3d mvd dd. _ dd.— Zd- 3d- and mod and mod mad and dmd dd.— dd.— Rd. 3.? 2d- dd.— wod odd dd.— ccd dd.— dd.— dd. _ 14 15 l6 17 100 19 20 21 22 *Percentage of U(IV) to total U. * *Percentage of Fe(II) to total Fe. Variables 1 to 13 were measured in groundwater (gw) samples, variables 14 to 20 in sediments (sed), and variables 21 and 22 in tracer studies. Pearson correlation is significant if greater than 0.482 (two tailed, alpha=0.05) 81 Pvrosecmencing results The 16S rRNA gene surveys produced a total of 97, 610 sequences for 17 samples that covered a wide range of conditions in the F RC Area 3 including wells from the inner and outer loops, three MLS wells at four different depths and one well (F W105) outside the treatment area. The high throughput provided a wide view of the microbial diversity present at the site (Table 3.5). The surveys were powerful enough to detect phyla poorly represented in the databases (Table 3.6). In some of these cases there were more sequences in the libraries for a given phyla than in the complete Ribosomal Database Project dataset, though RDP Classifier tools is better trained for well studied groups. Two samples yielded a low sequence count and were discarded from the analysis (F W100-l and F W101-1). Without considering those samples there was no significant difference in microbial richness recovered between inner and outer loops. Rarefaction curves (Figure 3.3) suggests that even with thousands of sequences per sample the sampling would be incomplete. This was not surprising given the high microbial diversity expected for non-extreme environments (17, 26). Estimates of coverage (data not shown) were similar to previous experiments that used ~100 clones per library (10). Given that we obtained 20 to 80 times more sequences per sample, it is clear that rarefaction curves are only informative when a plateau has been reached, otherwise the coverage estimation is deemed to fail, especially if using non-parametric models such as Chaol. 82 Table 3.5. Sequences per sample and number of groups detected at different levels of taxonomy. Well Sequences Phylum Class Order Family Genus FW104 6671 1 8 20 34 66 99 FW101-1* 52 7 10 11 16 15 FW101-2 6769 15 19 34 67 94 FW101-3 6458 18 20 34 34 107 FW101-4 5450 17 21 40 91 146 FW102-1 8566 18 21 41 36 165 FW102-2 8881 18 20 39 92 170 FW102-3 5954 16 20 37 75 118 FW102-4 7118 20 25 43 91 143 EYE/92.6 _______ §L3§ _____ 1.5. _ _ - _ .22 ..... 3 £5 ..... 7.4 _____ 1.09 _ _ FW024 8121 15 18 36 68 92 FW100-1* 83 13 16 13 16 17 FW100-2 2050 17 21 40 79 95 FW100—3 5564 18 24 41 82 121 FW100-4 5451 20 34 40 89 169 -EVYLQ} _______ 23.4.5 _____ L7_____29 _____ 3 .7 ..... 8.0 ..... 1.2.3-- FW105 6889 17 21 39 77 122 Whole set 97610 26 29 57 151 251 Classification based on RDP Classifier at 80% bootstrap confidence. *Discarded from later analyses. Table 3.6. Detection of phyla poorly represented in 168 rRNA databases in libraries from FRC area 3 samples Found in Present in Average identity to closest Phyla library RDP* isolated relative WS3 7 203 84.5% TM 7 185 883 77.1% Tenericutes 7 2570 91 .8% 0p] 0 2 369 82.0% OD] 545 195 74.6% Nitrospira 6 2045 95.6% Lentisphaerae 1 6 271 92.7% Gemmatimonas 2099 1 572 89.9% F usobacteria 23 1929 86.6% F ibrobacter 23 3 64 95 .6% Deinococcus— T hermus 13 1305 95.0% C hlamydiae 4076 654 91.3% BRC] 121 86 82.2% *RDP Release 10, Update 6 of 12/3/2008 containing 715,637 aligned 16S rRNA gene sequences. Comparison is against the complete RDP dataset (Both partial and near-full length sequences, from clones or isolates, of good or suspicious quality. Sequences were assigned to a particular phylum using RDP Classifier with 80% bootstrap confidence. Richness (number of different OTUs at 95% similarity) correlated with the . 2 . number of sequences per Site (r =0.6) as expected when coverage 1S not complete. However, richness in some samples was high albeit low sequence counts. This is probably due to a more even community structure that makes it easier to capture a wider range of richness (Figure 3.4), as well as incomplete sampling. Nevertheless, the communities were generally dominated by few OTUs, in 10 most abundant OTUs in any sample accounted for 40% of the sequences for that particular sample. Singletons, sequences with abundance equal to one in one library, represented on average 50.1% of the richness for any given site. 84 1400 .\ 1200 «“31 a "FW103 E1000 FW100-3 8 FW102-4 g 800 - FWIOO-Z g FW105 3.600 _ FW100-2 '3 FW104 D FW102-3 400 - _FW026 0 l I r l I I I I I 1 O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O O .— N m V W \O l\ 00 05 2 Sequences Figure 3.3. Rarefaction curves for FRC Area 3 samples. The shape of the curves suggests that a plateau has not been reach and incomplete sampling may occurred. 85 'D'S/Smax + gowsvwnscvwwsmw ;§§§§§§§§§§§§§§ u—BBBBBBBLHLHBBBWU- mmmmmmm [amt-L. Figure 3.4. Richness estimations for samples from DOE FRC Area 3 as influenced by evenness and sampling effort. S = Richness (number of different groups), n = number of sequences sampled, J-eveness = Eveness indicator (range 0 to l), 11max and Smax refer to the maximum number of sequences and richness the complete set of samples, respectively. Changes in sampling correlated with changes in richness. In some wells, e.g.FW100—2, richness was still high even when sampling was low due to a more even community. Hierarchical clustering of samples Clustering of the samples based on their microbial community yielded four clusters (Figure 3.5). The clustering was consistent with differences in bioactivity and connectivity. Cluster A contained only samples with low bioactivity and poor connectivity to the injection well. Some samples from this group came from the inner loop but at depths (levels -1 and -4) that make them less likely to have received the electron donor (injected between levels -2 and -3). Cluster B contained all the samples with the highest bioactivity and U(VI) reduction ability. Three subgroups were present in cluster B, with subgrouping consistent with pH levels suggesting that even when the Communities respond to the treatment in a similar fashion, there is variation that can be 86 attributed to differences in pH, non homogenous starting populations or other variables. Clusters C and D contained wells with medium to low bioactivity and intermediate connectivity. Samples present in clusters B, C and D showed similar pH levels within clusters but cluster A did not, having a wide range of pH. — FW105 (4.66) — FW103 (5.92) — FW102-1(6.4) A — FW101-4(5.6) — FW100-4(5.8) ~-— FW102-4(4.4) _, FW104(5.8) —_{: FW026 (5.7) B1 — FW101-3 (6.1) _L— FW101-2(6.2) B2 B FW102-3 (6.2) ,_ FW102-2 (6.5) B3 — FW100-3 (5.9) — FW024(5.9)| J C — FW100-2 (5.5) D l l l l l 0.2 0.3 0.4 0.5 0.6 Sorensen abundance-based distance Figure 3.5. Clustering of samples by abundance-based Sorensen index. Well name and (pH) shown. Four clusters were defined (A to D). Cluster A was comprised of wells with low activity and connectivity to the carbon injection well. Cluster B has the most active communities and the highest connectivity. Clusters C and D contained wells with medium activity. In clusters B and CD, the subgroups seemed to be influenced by pH. Hierarchical clustering used an average neighbor algorithm. 87 Microbial communig structure The microbial communities were dominated by Proteobacteria and Acidobacteria, this agrees with previous surveys of the site (10, 27). On average Proteobacteria and Acidobacteria together contributed 52% of the sequences, with combined contributions ranging from 24% at FW100-2 to 74% at F W024. Proteobacteria contributed 39% on average, with a maximum of 71% at F W024 and minimum of 18% at FW100-2. At F W024, 46% of the sequences were related to Sulfuricurvum, an anaerobic sulfide oxidizer that can use nitrate and hydrogen as electron donor (30). Acidobacteria contributed 13% of the sequences with a maximum of 33% at FW026 and minimum of 3% at FW026.The Actinobacteria contribution was around 6% with the exception of one well (FW100—2) where its contribution was 36%. The Proteobacteria contribution varied the most, followed by Actinobacteria, Acidobacteria and F irmicutes (Figure 3.6). 88 O\ \l O O l 0 (%g, 0 A O l |--l I I I O 0 Relative abundance N U) o d l l 0 -- I 0 # O l O l |-- '[li' .115 " Inj-"I (DI [i' o 0 III HI .0 II 01d 13V ”V Figure 3.6. Relative contribution of most abundant phyla to the microbial community from 15 samples from DOE FRC Area 3. Boxplots show medians, lower and upper quartile (box limits), range of observations (whiskers) and outliers if any (circles). Pro = Proteobacteria, Aci = Acidobacteria, Act = Actinobacteria, Fir = F irmicutes, Chla = Chlamydiae, Ver = Verrucomicrobia, Bac= Bacteroidetes, Gem = Gemmatimonadetes, Chlo = Chloroflexi, Pla = Planctomycetes, Other = other phyla. Unclassified bacteria are not shown. 89 When comparing the structure of samples in the inner and outer loops, Acidobacteria (p<0.01), Chloroflexi (p<0.05) and Chlamydiae(p<0.05) were significantly higher in the inner loop, while F ibrobacter was significantly lower (p<0.05) in this zone. F ibrobacter is a rare phylum originally found in the gut of herbivores where it plays a role in the anaerobic degradation of cellulose. It has been detected recently outside the gut in landfills (3 7). Our putative F ibrobacter sequences were on average 95.6% identical with its closest isolated relative in the RDP. When comparing the four sample clusters produced in the clustering analysis, cluster D was dominated by Actinobacteria and cluster C by Proteobacteria (Figure 3.7). Clusters A and B shared a similar profile, however cluster B had a significantly higher proportion of Chloroflexi (p<0.01), Chlamydiae (p<0.01), and BRCI (p<0.01), and a significantly lower proportion of Verrucomicrobia (p<0.05), Actinobacteria (p<0.05), TM7(p<0.05), and Bacteroidetes (p<0.05) than cluster A. Chloroflexi are a group of facultative aerobes known to subsist on carbohydrates and yeast extract and thought to have a role in cellular matter degradation (62), therefore, its abundance in wells with higher bacterial biomass could be due not to its ability to utilize the stimulatory carbon source but to recycle dead cells. Verrucomicrobia are ubiquitous in soil, with tolerance to acidic conditions and oligotrophy and thus a good competitor under these limiting conditions at the FRC. 90 > CD 0 U 100% 90% --....... ‘ U.;-__—_—_—— 80% $5: I I O .\ V — ‘ E’ ' 4 " — Ad E 70% I - — - 2,: .9 'N a o W I I I 3 60A) _, I II. M m .. '.' - a 3:1 «"3 N” “M \A. - vs. 3500/44-W M.. MW av-w — o .. U. A, H. M. H, a. M:- .,. I = :': I ‘A " .'. . U' V '. V V‘ V _ “V V‘ -:-..v~we-v~ Mw-MN m 40%.;.;...;.vx. .9“ MWWM .~ ;. -——- = In. 0.: In 1'. m ' g -:- :-: -: w ..v~ “WW“ -2- 0 1:3 :3: 1' P53 :1: -' :3' W N '11 "I 0 30% ~3:3~:1: :5 3:3 :3: .3 ': W W '-- 5 555 33: 5= *1‘~ M 20% n.’ a) 10% 0% :-: H-:- :u ':.:'u:-':-' '-:.':- ':-:'-:-I ' I I New“ , W?’ W” B” sslsiiaa tis‘aaag‘“ a ‘1“ at“ I Other Bacteria El Verrucomicrobia I F irmicutes D Actinobacteria Acidobacteria El Proteobacteria Figure 3.7. Relative contribution of the five most abundant phyla present at samples from the FRC Area 3. Samples are ordered into four clusters (A, B, C and D) according to the clustering analysis described earlier. Other Bacteria include Chlamydiae, Chloroflexi, Bacteroidetes, Gemmatimonadetes, Planctomycetes, Tenericutes, Nitrospira, F usobacteria, Deinococcus-Thermus, Deferribacteres, TM 7, Spirochaetes, WS3, 0D], 0P10, F ibrobacteres, BRCI, Cyanobacteria, Lentisphaerae, and unclassified Bacteria. Not all the mentioned phyla were not present in all sites. 91 1 w '70 A 90% 80% 70% 40% l N u 2 o. .\ 0\ Relative contribution to 10% 0% 5» D U. Proteobacteria I Epsilon I Delta 13 Alpha E] Gamma a Beta Figure 3.8. Class composition of the Proteobacteria from FRC Area 3 samples. U. Proteobacteria accounts for sequences classified as Proteobacteria but that could not be assigned with a 80% confidence to any of the classes of the Proteobacteria using RDP’s Classifier. Samples are ordered into four clusters (A, B, C and D) according to the clustering analysis described earlier. Taxonomic classes of the Proteobacteria phyla varied among the sites (Figure 3.8). When comparing sample clusters A and B fiom the clustering analysis (inactive vs active), cluster B was enriched in Deltaproteobacteria (p<0.01), while cluster A had a significantly higher proportion of Gammaproteobacterz'a and Alphaproteobacteria. Cluster D was similar to cluster A in terms of abundance of Gammaproteobacteria. 92 When comparing the inner and outer loops, we found Deltaproteobacteria in higher proportions (p<0.05) at the inner loop, while Epsilonproteobacteria were significantly lower (p<0.05) in this zone. The great variability observed in Proteobacteria is probably due to the range of conditions surveyed in terms of pH and carbon availability (measured as chemical oxygen demand due to connectivity differences. Also the proteobacteri are the most metabolically and ecologically diverse group of bacteria (19). Given that the Deltaproteobacteria class has many of the known uranium reducers such as Geobacter (Desulfomoronales), Desulfovibrio (Desulfovibrionales), and Anaeromyxobacter (Myxococcales), we evaluated if orders of this class were present in the samples and if they were distributed in the same abundance in samples from inactive wells versus samples from active wells. The results showed that the cluster B’s Deltaproteobacteria were enriched in members of the orders Myxococcales (p=0.01), Desulfobacterales (p<0.05) and Desulfovibrionales (p<0.01) in contrast to low activity wells. 93 A B C D (0.88%) (8.46%) (0.80%)(0.54%) A100% ”1!!! , I \- _ '5? 35:-2:} ., -Z:--I 2% 90% 23:}: “5:31;” ‘ ‘25:: :Ij-I:j::'i—:j:§:-I:‘ —:::II:' 3.... :2: 1.::: «2:2: 2:: :.:r::- :. 1: f § 70% — ~ — ZfiEj- - ~ —+ '7'- s 60% :1: . .::2: 333:: -—— - :3 500/ . _2-31: ~_: _.¥::-_ . _:-;:3::-_ __ _ __ _ 3 ° -::-- __ - -::--: -:=:--: = ;_-_', .555; 1. .1} {3. :2”: 2:3} f}. .3 40% ‘ - — .-_. - 5 '. ...-._ .5}:- ‘E 30% - - _' —: u - 3: ~ I;:-;:-.; —— ~ —- - =1 -.'-'. 4:? -I 3'3 5 *5 i-‘-' t: . _ ‘5 °‘ 0% 2E- .. at b -I'- at t» a 2% it" at ::e .22 e? I Unc. Deltaproteobacteria El Syntrophobacterales ’00? El Desulfovibrionales Cl Desulfomoronales El Desulfobacterales I Bdellovibrionales Myxococcales Figure 3.9. Relative abundance of Deltaproteobacteria orders in FRC Area 3 samples. Samples are ordered in the four clusters (A to D) that were defined previously in clustering analysis. The average contribution of the Deltaproteobacteria to the total community in that specific cluster of samples is given in parenthesis. U. Deltaproteobacteria are sequences that cannot be assigned with >80% confidence to one of the orders of the Deltaproteobacteria class by RDP classifier. Presence of genera with members known to reduce uranium 94 Genera known to contain U(VI) reducers were detected in the samples. Desulfovibrio was on average the most abundant one followed by Desulfosporosinus, Geobacter, Anaeromyxobacter and CIostridium. The first four genera were present in all wells with high bioactivity (Figure 3.10), Clostridium was present in these wells but its abundance was higher at a low activity well from cluster A. 12 I Geobacter Desulfovibrio B A naeromyxobacter Desulfosporosinus I Clostridium g—o o a I co \\\\\\\\\\\\ \\\\\\\\\\\\\\\\\\\ I Relative contribution to to library (%) h 0‘ \\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\ IIIII——_— IIIIIIIIIIII'IIIIIII’IIIIIII. \\\\\\\\\\\\\\\ , : i E: ’ l '5 . ‘S:: l O 1 W :9 1.: ’5.‘ . ‘ 9" @‘M'v x," o) my @‘W xwwfi'v Q6” \Q \Q \Q \Q \Q .\Q \Q \Q N‘éeas si‘ksaa $8 at“ C D Figure 3.10. Relative contribution of most common uranium reducing genera found in the samples from DOE FRC Area 3. Samples are ordered in the four clusters (A to D) defined previously in clustering analysis. Cluster B contains samples with highest bioactivity. 95 Abundance of Geobacter, Anaeromyxobacter ,Desulfosporosinus, and Anaeromyxobacter sequences were higher at wells with higher connectivity, as measured by tracer recovery (Figure 3.11). The location where the population peaked in abundance differed for each group (order: Desulfovibri0> Geobacter > Anaeromyxobacter > Desulfosporosinus), suggesting that some niche specialization may be have occurred along the electron donor gradient. The location of the peak for Geobacter is consistent with a study that found elevated Geobacter abundance in clone libraries when organic carbon (measured as COD) was elevated (27). Four other genera with known uranium reducers were also detected: Desulfitobacterium, Shewanella, Acidovorax, and Deinococcus. They, however, were not present in all the wells where uranium reduction was established and their relative abundance was generally low (0.1 to 1%). Hence, their role in uranium reduction is difficult to establish and for that they were not considered in the analysis. 96 -O- Geobacter 100 d ‘o. 4 -D-Desu(fi)vibrio :3 ' . -&' Anaemmyxobacter E 80 t .0-Desulfospomsinus ‘ g S V ' -I- C lostndzum ,1 g “m- Tracer recovery as t s2 w - ii 5% 4° ‘ ‘ é‘ a. 20 o . $5®W3Wb® 153' $5.1? Wg‘“ Q9§4§§y$4® We“ $3®§®<¢Q<€M M3 Miss Figure 3.11. Location of maximum abundance for uranium reducing genera along the connectivity gradient at FRC Area 3. Plotted are the ratio of abundance to maximum abundance for selected groups and the proportion of tracer recovery, all in a 0 to 100% scale. 97 Indicator species Indicator species analysis is an approach that can greatly reduce the complexity of the dataset by finding sequences that represent samples, or cluster of samples. Additionally, the indicator value is provided with an indication of statistical confidence. We used abundance information for 4719 OTUs over 15 samples to obtain indicator species. The first level of analysis focused on the four sample clusters found by clustering analysis (Clusters A, B, C, and D). Cluster A (mostly inactive wells) was represented mostly by Betaproteobacteria and Gammaproteobacteria OTUs (Table 3.7), two classes known to have nitrate reducing organisms. The top 15 representatives (ranked according to descending indicator value) contained three Rhodanobacter and two Castellinella OTUs. Rhodanobacter, a Gammaproteobacteria, was previously detected in high abundance in F W106, a location in Area 3 meters away from the treatment zone (10). FW106 has not received any stimulatory electron donor and it is still acidic @H 3.6) and highly contaminated. A metagenome analysis of a FW106 groundwater sample additionally revealed a Rhodanobacter sp. with a wide variety of metal resistance genes and indication that these genes were horizontally transmitted (24). Castellinella is a denitrifying Betaproteobacteria that has been found in biostimulated wells from the FRC Area 1 where it was the dominant nirK-containing denitrifier (49). Area 1 is also close to the main contamination source at the PRC, and has acidic conditions as well as high nitrate levels. The main differences between these areas 1 and 3 reside in the U levels, with Area 3 being higher, and the treatment system used. Area I experienced a pull-push 98 approach with ethanol while Area 3 used the hydraulic control system previously described. Castellinella isolates from FRC Area 1 were recovered under acidic (pH 4.5) and neutral (pH 7.5) conditions. It was suggested that this group may outcompete other denitrifiers under acidic conditions characteristic of the FRC contaminated areas (49). Castellinella has also been detected in microcosm enrichment from the FRC Area 2 (1). The rest of the top 15 representatives of cluster A did not include genera known to be involved in bioremediation of U or nitrate. A Comamonas sp. was also found in top 15 ranking at #2, this genus is known for their ability to grow on organic compounds such as phenols. These compounds were present in the contamination source (58) and may be still present in locations poorly connected to the injection well. The main physiological theme in cluster A representatives was their resistance to harsh conditions like those from the FRC. The top 15 representative OTUs of cluster B contained seven OTUs related to U(VI)-reducing genera (Table 3.8). The wells in this cluster are the most active in microbial activity, and are also the only locations where U(VI) reduction has been detected. Out of the seven OTUs, four belonged to Desulfosporosinus (SRB F irmicutes), two to Desulfovibrio (SRB Deltaproteobacteria) and one to Anaeromyxobacter (FRB Deltaproteobacteria). Two denitrifiers, Vogesella (#3) and T hauera (#12) were also present as top indicator species. Thauera spp. have been previously found in the denitrifying bioreactor used ex situ to precondition the site by removing nitrate (28). Since aboveground treatment system was connected to the subsurface, it is likely that T hauera spp, were injected with the treated water. 99 Bacteria related to iron reduction/oxidation had a high indicator value for this cluster of sequences: Gallionella (#15, Ind.value 0.93, p=0.001), Thiobacillus (#16, Ind.value=0.928, p=0.001), Geothrix (#22, Ind.value=0.883, p=0.001), and F erribacterium (#28, Ind.value=0.883 p=0.115). Geobacter, a F RB also known to reduce U(VI), ranked slightly lower at position 53 (indicator value 0.73, p=0.065). A special case is that of T hiobacillus, a bacteria capable of reducing iron and nitrate but also able to oxidize iron and uranium (the later while reducing nitrate) (4). Its presence in the top of the indicator list suggest that it plays a role in the most active uranium reducing communities and may influence the mobility of uranium in situ. Two amoeba endosymbiont relatives were found in high in the indicator species ranking. The putative endosymbiont come from two different phyla ,Chlamydiae and Proteobacteria, and share the same likely host, Acanthamoeba, a common free-living protozoa that feeds on bacteria, algae and yeast in the environment (3 6). Considering that the sediments with high activity in occasions clog the wells with biomass (10), it is quite possible that predatory populations of amoeba exist at these locations, though this remains to be established directly by microscopy, cultivation, or PCR based methods. Clusters C and D did not contain known U(VI) reducers in their top 15 ranking taxa (Table 3.9 and Table 3.10, respectively). In cluster D, Geobacter was positioned #21 (p=0.065) and Desulfovibrio #99 (p=0.109). These results suggest that even when populations capable of reducing U(VI) exist in these locations, they are not good representatives of the main selection trends in these sites. In both C and D, the most frequent groups in the top of the list were related to Alpha- and Gammaproteobacteria. 100 Indicator species analysis was also performed on the cluster B. Subgroup B2 was formed by three samples, FW101-2, FW101-2 and FW102-3, and was consistently high in uranium reduction. In these three samples Desulfovibrio was a top indicator species where it ranked #1 (p=0.05 5) and #12 Q3=O.O3), respectively. Desulfosporosinus ranked #8 at B2 but the statistical confidence was low (p=0.317). In these samples Desulfovibrio spp. were always in the 10 most abundant species. Interestingly, Geothrix, an iron reducer Acidobacteria was also highly present in these samples, and always more abundant than Desulfovibrio. It is also likely that there are taxa not previously described as U(VI) reducers present and actively reducing uranium at this site. Candidates for these roles are iron reducers such as Geothrix (Acidobacteria) and F erribacterium (Betaproteobacteria). The indicator list for this group of samples could potentially be used to target specific groups in isolation efforts since the uranium reduction ability seems to be more widespread than originally described. 101 Table 3.7. Top 15 indicator species for sample cluster A, their classification and closest isolated relative. 3 Ind. - 1 2 . Rank Value vailue Classification Closest isolate in RDP ”:22? Treponema Parvibaculum sp. psclO; 0‘ ”’00 0-002 (spirochaetes) EU930870 90.7 C omamonas Bacterium r19; ABOZl327 02 0'997 0'002 (Betaproteobacteria) 99_ 5 U. Sphingobacteriales Niastella sp. RHYL-67; 03 09” 0'008 (Bacteroidetes) EU917053 94.] C astellaniella C astellam'ella defragrans; 04 0'985 0‘00] (Betaproteobacteria) 62Car; AJ005449 99. 5 Rhodanobacter Cimanggu media isolate 88; 05 0.973 0.005 (Gammaproteobacteria) Cimanggu media 88; B; AF229452 99.0 Rhodanobacter Rhodanobacter ginsengisoli; 0‘ “'9‘“ “'0‘” (Gammaproteobacteria) GRl7-7; EF166075 99.0 U. Rhizobiaceae Ensifer xinjiangensis; Rx22; 07 0'938 0'004 (Alphaproteobacteria) AF250353 99.0 Opitutus Opitutus sp. VeSml3; X99392 08 0'904 0'064 (Verrucomicrobia) 97.1 Rhodanobacter Cimanggu media isolate 88; 09 0.886 0.002 (Gammaproteobacteria) Cimanggu media 88; B; AF229452 99.0 Gp6 (Acidobacteria) Unidentified eubacterium clone ‘0 0.872 “-004 sto7; A1229181 339 C astellaniella C astellam'ella sp. 7.5A2; 1 1 0'866 0'002 (Betaproteobacteria) EF175378 99. 5 Gp6 (Acidobacteria) Marine gamma proteobacterium ‘2 0-847 0-006 HTCC2246; AY386337 g5_3 13 0.843 0.065 U. Bacteria Sigifigsogrggnas spJQl-3; 32 3 U. Verrucomicrobia/es Bacterium Ellin516; AY960779 l4 0'833 0'07] (Verrucomicrobia) 33_7 U. Hypomicrobiaceae Devosia ginsengisoli; Gsoil 326; 15 0‘833 0'00? (Alphaproleobacteria) ABZ71045 9&0 Ind.value = Indicator value, p-value = probability of obtaining as high an indicator values as observed over 1000 iterations. 1 Classification based on RDP Classifier with a 80% bootstrap confidence. 2 Closest relative was found with RDP SeqMatch function using an isolates-only dataset. 3 Sequence identity with the closest isolate from the RDP. Unclassified sequences (U) were not able to be assigned to a lower taxon with the 80% confidence. Wells in this cluster: F W100-4, F W101-4, FW102-1, FW102-4, FW103, and F W1 05. Groups in bold were significant (q-value <0.05) under the false discovery rate analysis. 102 Table 3.8. Top 15 indicator species for sample cluster B, their classification and closest isolated relative. 3 Ind. - l 2 ' Rank Value vzilue Classification Closest isolate in RDP Ides/Sty 0 Desulfosparosinus Desulfosporosinus sp. Sapy; 0‘ "00° 0‘00] (Firmicutes) AF159120 99.0 Desulfosporosinus Desulfosporosinus sp. Sapy; 02 M97 “'0‘” (Firmicutes) AF159120 99,0 03 0 979 0 001 U. Burkholderialas Vogesella indigofera; ATCC ' ' (Betaproteobacteria) 19706T; A8021385 95,6 U. Peptococcaceae Anaerobic bacterium Prop2; 0“ 0377 0'00] (Firmicutes) AY756143 92 7 U. Chlamydia/es Endosymbiont of Acanthamoeba sp. 05 0974 “002 (Chlamydiae) uwczz; AF083616 91,7 Desulfosporosinus Desulfosporosinus sp. Sapy; 06 0'97] 0'00] (Firmicutes) AF159120 98.0 Desulfovibrio Desulfovibrio putealis (T); DSM 07 0'97" 0'00] (Deltaproteobacteria) 16056; AY574979 99,5 08 0.968 0.00] U. Firmicutes Clostridium sp. 9B4; AY554416 92.7 Desulfovibrio Sulfate-reducing bacterium F 1- 09 0367 0'00] (Deltaproteobacteria) 7b; AJ012594 99,5 . Desulfomonile limimaris (T); 10 0.962 0.002 U. Bacteria AF230531 86.3 11 0.959 0.001 U. BRCI Bacterium Ellin37l; AF498753 82.8 12 0.957 0.006 U. Betaproteobacteria Thauera sp. R-28312;AM084110 99.0 Anaerom xobacter Anaeromyxobacter dehalogenans 13 0.948 0.001 ( Del 1‘ roJI’eobacteria) (1‘); 2CP-l; ATCC BAA-258; “p AF382396 96.1 . Endosymbiont of Acamhamoeba 14 0.943 0.001 U. Alphaproteobactena polyp haga; AF132138 90.2 . Gallionellaferruginea subsp. 15 0.929 0.001 U. Betaproteobactena caps iferr iformans; DQ386262 96.6 Ind.value = Indicator value, p-value = probability of obtaining as high an indicator values as observed over 1000 iterations. 1 Classification based on RDP Classifier with a 80% bootstrap confidence. 2 Closest relative was found with RDP SeqMatch function using an isolates-only dataset. 3 Sequence identity with the closest isolate from the RDP. Unclassified sequences (U) were not able to be assigned to a lower taxon with the 80% confidence. Wells in this cluster: FW026, FW101-2, FW101-3, FW102-2, FW102-3, and FW104. Groups in bold were significant (q-value <0.05) under the false discovery rate analysis. 103 Table 3.9. Top 15 indicator species for sample cluster C, their classification and closest isolated relative. 3 1nd. - 1 2 . Rank Value vzilue Classification Closest isolate in RDP [dig/n? 0 U. Chlamydiales Neochlamydia hartmannellae (T); 01 1000 0009 (Chlamydia) All-15p; AF 1 77275 84'9 02 1.000 0.013 g . Legionella sp. 1129; ABOS8910 90.7 ammaproteobacterta . _ Tuber borchii symbiont b-Z43; b- 03 1.000 0.010 U. Bacterozdetcs Z431; AF233292 89.3 Rhodocista zospirillum sp. DA10-2; 04 1 'OOO 0'0 1 3 (A Iphaproteobacleria) AY 1 l 8225 93 '2 . Pseudophormidium sp. 05 1.000 0.008 GpV (C yanobacterza) ANT.PROGRESS2.2; AY 493 583 95.6 . C handromyces [anuginosus (T)' 7 9 06 1.000 0.01.. U. Bacteria Sy t2; A1233939 82.9 . C lostridium paradoxum; DSM 07 0.981 0.043 U. Bacteria 7308T; para2;Z69941 79.0 08 0.968 0.001 S“bd'v's'°'i' 3 . Bacterium Ellin514; AY960777 94.6 ( Verrucomicrobia) . Neochlamydia hartmannellae (T); 09 0.964 0.003 U. Bacteria Alep; AF177275 85.2 Geobacter sulfurreducens PC A; 10 0.957 0.00] 0D] AE017180 75.6 Rhabdochlamydia Rhabdochlamydia crassificans; ' 1 ”“947 0'008 (Chlamydiae) CRIBOI; AY928092 93'2 . Sulfate-reducing bacterium R- 12 0.946 0.010 U. Bacteria PropAl;A.101259l 82.0 13 0.943 0.010 Gp3 (Acidobacteria) Bacterium Ellin37l; AF498753 89.3 . Desulfohalobium utahense (T); 14 0.931 0.086 U. Proteobacteria EtOH3; DQ067421 82.9 15 0.931 0.006 U. Alphaproteobacteria 305p "”1”" amazoneme; 21R; 89.3 AY741 146 Ind.value = Indicator value, p-value = probability of obtaining as high an indicator values as observed over 1000 iterations. 1 Classification based on RDP Classifier with a 80% bootstrap confidence. 2 Closest relative was found with RDP SeqMatch function using an isolates-only dataset. 3 Sequence identity with the closest isolate from the RDP. Unclassified sequences (U) were not able to be assigned to a lower taxon with the 80% confidence. Wells in this cluster: F W024 and FW100-3. Groups in bold were significant (q-value <0.05) under the false discovery rate analysis. 104 Table 3.10. Top 15 indicator species for sample cluster D, their classification and closest isolated relative. 3 1 d. - 2 ‘ Rank Vglue vzilue Classification Closest isolate in RDP Ide(r01/t01)ty . Candidatus Brocadia sp. 1 1.000 0.070 U. Bacteria 40; AM285341 81.5 U. Lachnospiraceae C lostridium herbivorans 2 1'000 00% (Firmicutes) (T); 54408; L34418 95'6 Verrucomicrobia 3 1.000 0.064 U. Bacteria bacterium YM31-066; 87.8 AB372850 . Unidentified eubacterium 4 1.000 0.069 U. Bacteria clone BSV28; A1229190 82.0 Flavimonas Pseudomonas sp. SV4; 5 Low 0072 (Gammaproteobacteria) AF251334 92'2 . Candidatus Brocadia sp. 6 1.000 0.070 U. Bacteria 40; AM285341 80.5 . Bacillus sp. BR (T); 7 1.000 0.053 U. Bacteria AM050346 82.0 . Butyrate-producing 8 1.000 0.060 fggffutr’fnicmbia) bacterium A2-194; 95.6 0 AJ270473 . Treponema porcinum (T); 9 1.000 0.064 U. Bacteria 14V28; AY518274 88.3 U. Actinomycetales Actinobaculum sp. 10 1'0“) 0073 (Actinomycetes) P2P_19 P1; AY207066 94'6 lsosphaera Isosphaera -like str. C2- 1 1 1000 0062 (Planctomycetes) 3; AF239692 99.5 Desulforomonas Delta proteobacterium 12 1'000 0'066 (Deltaproteobacteria) 112 (T); AY835388 95'6 Sphaerobacter 13 1.000 0.079 U. Bacteria thermophilus; DSM 78.4 20745T; AJ420142 U. C lostridiales C lostridium termitidis; ‘4 1'0“) mm (Firmicutes) DSM 5396; X71854 9L7 Treponema T reponema bryantii (T); 15 ”mo 0'064 (Spirochaetes) RUS-l; M57737 88'7 Ind.value = Indicator value, p-value = probability of obtaining as high an indicator values as observed over 1000 iterations. ] Classification based on RDP Classifier with a 80% bootstrap confidence. 2 Closest relative was found with RDP SeqMatch function using an isolates-only dataset. 3 Sequence identity with the closest isolate from the RDP. Unclassified sequences (U) were not able to be assigned to a lower taxon with the 80% confidence. Cluster formed by well FW100-2. Groups in bold were significant (q-value <0.05) under the false discovery rate analysis. 105 Ordination of samples by principal coordinates analysis Principal coordinates analysis showed an ordination that was consistent with previous clustering analysis as shown with the minimum spanning tree overlaying the ordination (Figure 3.12 left panel). When testing the ordination against gradients of environmental factors, chemical oxygen demand (COD) in sediments, sulfide in sediments, total iron in sediments, and uranium levels in sediments were significant in explaining the ordination (p<0.05). However, we previously established that colinearity was present in these variables. We hypothesize that COD is the independent variable of this group. COD represents the organic carbon content in sediments, and also probably an indication of electron donor availability in this electron-donor—limited environment. COD likely influenced sulfide, iron, and uranium levels through the activity of sulfate-, iron- and uranium-reducing microorganism that used the organic compounds. This results agrees with our hypothesis that the injection of the stimulatory carbon, ethanol, helped shape the community structure. 106 A FW024 ‘ FW1 00-3 (%L0 91) zvooa to,1'0.0'011'0.Z'0.£'0‘v'019'0 1 FW026 FW102-4 FW100-4 FW101-2 FW101 FW101-3 0.4 0.2 010 0'2 0'.4 PCOAl (22.07%) B Fw’Tz4 ‘ COD 3 Sulfide (Sed) 1 FW100-3 U (Sed) i e(Sed) w ‘1 8 8 B « mime-2 f; 5 ‘ii’ . FW103 S o“ FW026 3‘: z... ‘ FW102-4 .0 FW104 °_ FW100-4 ' I p o 0 FW105 FW102-2 FW101-2 “ o w102-1 - FW101-4.1; FW102-3. FW101-3 9 l 1 1 N 0.4 0'2 0.0 0.2 0.4 PCOAl (22.07%) Figure 3.12. Principal coordinates ordination on abundance-based Sorensen distances for samples from the FRC Area 3. First two dimensions together explain 38% of the original variability. The grouping of samples is consistent with previous clustering results as shown with a minimum spanning tree on top of the ordination (panel A). Vectors that explained significantly the ordination (p<0.05) are shown (panel B), though vectors were collinear (correlated). Cluster A (0), cluster B (I), cluster C (A), and D (*). 107 We also used non metric dimensional scaling (NMDS) to study the distribution of specific taxonomic groups likely related to the U(VI) reduction process. This model was preferred because it does not assume a specific distribution of species along a gradient. It is not clear if the measured environmental gradients were pronounced enough for our groups of interest to react to; thus, we cannot assume that our populations have a unimodal distribution (single peak along a gradient). In this case we tested only five main groups of variables, they were called pH, nitrate, chemical oxygen demand (COD), mean travel time and U in sediments. Mean travel time is a hydrology terms that refers to the time water from the injection wells takes to travel to a given location, the more connected the location the shorter the travel time. Because the variables were correlated, these names actually represent several collinear variables. The names used are to our best estimation of what is the independent variable for the group, though it is possible that a collinear variable is responsible for the response. Four of the five groups of variables were significant for at least one group of organisms (Table 3.11). COD and Uranium levels in sediment were the most common factor to be found significant in explaining the variability of the groups of interest. COD explained significantly (p<0.05) the distribution of Bacteria, Anaeromyxobacter, and Desulfovibrio along the samples. Uranium levels in sediment were highly significant in explaining the distribution of Deltaproteobacteria, Desulfovibrio and Desulfosporosinus (Table 3.11). However, this variable is correlated with sulfide and iron levels in sediments, so part of the response could be due to changing levels of sulfate and/or iron. Nitrate was weakly significant (p<0.1) in explaining the distribution of Geobacter OTUs. Nitrate was highly 108 correlated with sulfate, na, mg, Mn, chloride and Ca. The results suggest that SRB Desulfovibrio and Desulfosporosinus, as well as FRB Anaeromyxobacter are responsible for the uranium levels, and that they responded to the injection of the stimulatory electron donon Table 3.11 Environmental factor that explain distribution of groups along the samples. COD Travel Group Nitrate U(sed) Jsed) time Acidobacteria - . _ Gp8 - * . - Betaproteobacteria - . . - Deltaproteobacteria - - - . Anaeromyxobacter - ** * * Desulfovibrio - ** * * Geobacter . - - - F irmicutes - * * . Desulfosporasinus - ** . - Verrucomicrobia - — - - Bacteria - . * - Environmental factor that significantly explain the distribution of clusters in NMDS ordination space. °p<0.1,*P<0.05, **p<0.01. Similarity threshold: Genus level analysis=99%, Phyla and class level = 95%. pH was not significant in explaining the distribution of any of the groups studied. DISCUSSION Effect of hydraulic control on microbial communities of the treatment area The establishment of a controlled hydraulic system created an opportunity to study the bioremediation of a contaminated aquifer that went through different stages from 109 denitn'fication to iron reduction to sulfate reduction (27, 58, 59). We observed differences in the spatial scale that are a consequence of the hydraulic flow regime and the injection of ethanol as the stimulatory electron donor. The geochemical gradients shaped the microbial communities to resist toxicity, denitrify, reduce iron and reduce uranium in different zones of the study area, creating differences in the spatial patterns. Other work that tracked the succession of communities in FRC Area 3 since the stimulation started determined the establishment of these different types of communities (toxicity resistant/denitrifiers/metal reducers) but in a temporal scale (27). In biological terms, the establishment of spatial gradients increased the diversity of metabolism for the site creating a biological transition zone. This zone potentially works as buffer zone, protecting the inner loop from intrusion of contaminated water if the hydraulic system is perturbed and could be important in the long term bioremediation of the site. Besides the hydraulic control and ethanol injection, an additional factor in shaping the microbial communities of the site is the conditioning phase that used a fluidized bed bioreactor to remove nitrate (competing electron acceptor). This bioreactor worked for over 4 months treating water ex situ and reinjected treated water as well as microbial cells into the study area (28). This bioreactor was initially inoculated with biomass from the Y- 12 wastewater treatment plant (300 m away from Area 3, also at the FRC) (28), and biomass from a pilot study FBR that showed denitrification as well as U(VI) reduction (61) . The pilot scale reactor was also inoculated with site water (TBP-16 well the FRC). In both reactors, pilot and full scale, ethanol was used as electron donor and populations of SRB such as Desulfovibrio spp. were detected (28, 61) . In the pilot scale bioreactor, uranium reduction was also present and additions of molybdate (a SRB inhibitor) 110 decreased sulfate reduction as well as U(VI) reduction, suggesting that SRB were partly responsible for the uranium reduction (61). Other groups detected in the bioreactors that were also detected in the sediment samples include Acidovorax (denitrifier, also U(VI) reducer), Thauera (denitrifier Betaproteobacteria), Sporomusa (acetogen F irmicutes), Stenotrophomonas, Pseudomonas, Decholormonas (denitrifier Betaproteobacteria), Hydrogenophaga (hydrogen oxidizing Betaproteobacteria), and Clostridium among others (28, 61). No FRB known to reduce uranium where detected in either bioreactor. However, Geobacter and Anaeromyxobacter, both U(VI)-reducing F RB, were detected in enrichments of sediments of Area 1 of the FRC (42). The bioreactor is likely a source inoculum in the subsurface ecosystem, helping U(VI)-reducing SRB to establish in the aquifer and compete with local FRB populations that are more likely in lower numbers. The F BR inoculums is probably a reason why SRB are predominant in this location, and U(VI) reduction works for this site. Since not all SRB reduce U(VI) and many SRB can utilize ethanol as electron donor, the success formula of this site may not occur in other sites. In a different site also contaminated with uranium, injection of acetate stimulated an early response of F RB (Geobacter spp.) that reduced uranium but later succession with SRB Desulfobacterales decreased uranium reduction (25). Fate of electron donor injected and populations that responded to the treatment Several studies have shown that ethanol is the best electron donor for uranium reduction . at the FRC (1, 59). Ethanol, however, is rapidly converted into acetate at the injection point, acetate accumulates and it is later consumed (1 , 10, 39). The oxidation of ethanol 111 to acetate and its release to the environment is consistent with the activity of SRB such as Desulfovibrio, and Desulfosporosinus (31, 47). Additionally when ethanol has been used as electron donor in studies at the F RC, sulfate and uranium reduction occur simultaneously (1, 38, 59). This suggests that at the FRC, SRB are more important in uranium reduction than FRB. In our analysis, SRB Desulfovibrio and Desulfosporosinus correlated with uranium reduction in ordination analysis and are characteristic of sites with high U(VI) reducing abilities in the indicator species. This however does not discount that FRB partially contributed to the reduction even though acetate provides less energy than ethanol. Akob and colleagues found that F RB Geobacter spp. reacted to ethanol amendments in microcosm enrichment from Area 2 that showed U(VI) reduction (59). In our case F RB Geothrix, F erribacterium, and Geobacter were also abundant in sites with high UW I) reducing activity. The three of them are known to utilize acetate to reduce iron but U(VI) reduction has been demonstrated for Geobacter only (34). In all the active wells, the F RB Geothrix was the more abundant than any SRB. At the injection well, F W1 04, where ethanol is the main electron donor, Geothrix and F erribacterium where the most dominant FRB, being more abundant than Geobacter. In this well, reduced iron was at its maximum, so competition for oxidized iron, or unknown abilities of the specific F erribacterium and Geothrix spp. could have excluded Geobacter. In later sample wells when acetate is the main electron donor, Geobacter peaked in abundance. After Geobacter, Anaeromyxobacter (also FRB) and Desulfovibrio (SRB) peaked in abundance. The fact that the SRB were less abundant in these wells could be due to sulfate reduction being less favorable than iron reduction when acetate is the 112 electron donor. Another possible explanation is that since acetate oxidation is not a frequent characteristic of Desulfovibrio, this SRB could be growing as the hydrogen- consuming partner in a syntrophic association with an acetate-consuming partner. This type of association has been demonstrated for pure cultures of Geobacter and Desulfovibrio (15). In general, it seems that SRB dominate these samples because of their ability to oxidize ethanol, the time of sampling in the succession event, and the initial advantage given by the F BR. FRB likely use mainly acetate to reduce iron and benefit from the presence of iron oxidizers (aerobic and anaerobic) at the site. This dynamic seems to be present in our site since many iron oxidizing genera such as Gallionella and T hiobacillus were detected in the most active wells. This dynamic could also be applied to sulfur since Sulfuricurvum, an anaerobic sulfide oxidizer was present in high abundance but only in one site (FW024). Uranium reduction and its correlation with microbial groups Comparison of the levels of uranium reduction in different wells with the locations where the putative iron reducers peaked, suggests that multiple populations are responsible for uranium reduction. The highest uranium reduction (61%) was at the injection site FW104, where ethanol was rapidly converted to acetate (likely by SRB) and FRB Geothrix and F erribacterium dominated. In slightly less connected wells, FW102- 2, and FW102-3, Geobacter peaked in abundance. In these wells the average uranium reduction was 24%. Finally at wells FW101-2 and FW101-3, uranium-reducing 113 candidates with sulfate-reducing activities peaked in abundance. Those wells had 54% and 51% of reduced uranium, respectively. We hypothesize that F RB are responsible in the more connected wells for uranium reduction, and SRB are responsible for uranium reduction at the injection site and less connected points. This agrees with previous findings of Geobacter and Geothrix being associated with high levels of COD and no dissolved oxygen control (27). In our case more connected wells are the equivalent of that because they have more carbon available and a higher chance of having dissolved oxygen (due to injection process). Desulfovibrio spp. were associated with dissolved oxygen control in previous studies (27), this agrees with our findings of Desulfovibrio peaking in less connected wells where oxygen is less likely to be present due to its consumption in earlier wells. It remains unclear if the FRB F erribacterium and Geothrix are involved directly in uranium reduction or if they are only benefitting from the added electron donor while reducing iron or humic acids. Since reduced iron compounds (from FRB )(44) as well as reduced humic acids (from a humic reducer such as Geothrix) (12, 41) can potentially reduce uranium abiotically, F erribacterium and Geothrix activities could be linked to uranium reduction. Geothrix has been associated with uranium reduction in other studies of the FRC, though no direct evidence of uranium reduction exists to date (7, 10, 27). The role of humic acids as electron shuttles for uranium reduction also remains unknown (39). Clustering, direct gradient ordination, and indicator species analysis suggests that in more active communities, Deltaproteobacteria, and more specifically its SRB members seemed to be playing a bigger role in the immobilization of uranium at the time of the sampling. In less active communities, the lack of electron donors has selected for species 114 capable of life under acidic and toxic conditions. Our indicator species analysis suggests that the groups more likely to reflect the uranium-reducers are related to Desulfovibrio, Desulfosporosinus, Anaeromyxobacter, and Geobacter. Differences in uranium reducing populations at the active locations can be attributed mainly to differences in connectivity to the injection well that likely influenced that carbon available for the microbes. There are populations that were not related to uranium reduction but were a consequence of the establishment of high number of bacteria in the active zone, e. g. populations likely involved in recycling dead cell material such as Chloroflexi were significantly higher in active sediments. Patterns of diversity The communities coming from samples with reduced uranium were not the richest communities. This is likely due to a combination of incomplete sampling and differences in evenness. Uneven communities reflect on one extreme a thriving community with highly enriched groups (e. g. SRB in FW104) and on the other extreme a community struggling to survive toxic conditions (e.g. F W105). Clustering by diversity patterns was more informative, and showed that communities responsible for uranium reduction were similar in composition suggesting that the same main groups in these communities were reacting to the treatment. The treatment changed the subsurface by creating better conditions for microbes in terms of pH, heavy metals, and electron donor availability. Electron donor availability was probably the most important change since energy source is commonly the limiting factor in uranium reduction in groundwater (59). The impact of 115 the treatment can be tracked to groups that increased in abundance such as C hloroflexi, Acidobacteria, Deltaproteobacteria, Desulfovibronales, Myxococcales and Desulfobacterales. In general the treatment shaped the community to provide specific functional groups for nitrate, iron, sulfate, and uranium reduction. When those databases are compared to our previous clone libraries results, some differences in abundance are noted. Since we used two different systems in terms of primers, product size, and amplification conditions; some differences in abundance can be due to primer coverage (primer bias), sampling, and regions flanking the template region (21). Additionally, the lack of a cloning step could potentially avoid cloning bias. Cloning bias has been reported for F irmicutes rRNA genes cloned into a Gammaproteobacteria host (48). In our case, Desulfosporosinus contribution comprised 1% of clone libraries and 8% of the pyrosequencing produced, using the same DNA extract. This difference is more likely attributable to difference in bias and sampling. Beside avoiding cloning bias, our methodology also allowed us to explore more deeply the microbial diversity of the site, and to track specific populations that occur at low frequency (e. g. 0.5% for Anaeromyxobacter) and whose trends were previously difficult to follow. CONCLUSIONS The results from a much larger number of 16S rRNA gene sequences allowed us to more comprehensively study the microbial communities in response to treatment than was previously feasible, and to compare their patterns of diversity with a statistical 116 confidence. Clustering analysis showed differences in community structures that can be related to connectivity and pH among wells as well as organic carbon. In contrast with other sites where Geobacter populations are more abundant, our results suggest that SRB such as Desulfovibrio were more involved in uranium reduction at this particular location at the time of sampling, this is likely due to the choice of electron donor, its metabolism by SRB and F RB, and the use of a bioreactor that likely injected SRB into the aquifer. ACKNOWLEDGMENTS The microbial study was funded by the US. DOE under grant DE-FG02-97ER62469, and the field work was funded by the US. DOE Environmental Remediation Sciences Program under grant DOE C05-OOOR22725. Mary Beth Leigh who was a collaborator in this work was supported by a US. NSF postdoctoral fellowship. We thank Hui Yan, Kenneth Lowe, and Tonia Mehlhom for help with field and analytic work and Benli Chai and Qiong Wang from the Ribosomal Database Project for help in analyzing such large sequence dataset. 117 References Akob, D. M., H. J. Mills, T. M. Gihring, L. Kerkhof, J. W. Stucki, A. S. Anastacio, K. J. Chin, K. Kusel, A. V. Palumbo, D. B. Watson, and J. E. Kostka. 2008. Functional diversity and electron donor dependence of microbial populations capable of U(VI) reduction in radionuclide-contaminated subsurface sediments. Appl Environ Microbiol 74:3159-70. Anderson, R. T. 2006. DOE genomics: Applications to in situ subsurface bioremediation. Remediation Journal 17:23-38. Anderson, R. T., H. A. Vrionis, I. Ortiz-Bernad, C. T. Resch, P. E. Long, R. Dayvault, K. Karp, S. Marutzky, D. R. Metzler, A. Peacock, D. C. White, M. Lowe, and D. R. Lovley. 2003. Stimulating the in situ activity of Geobacter species to remove uranium from the groundwater of a uranium-contaminated aquifer. Appl Environ Microbiol 69:5884—91. Beller, H. R. 2005 . Anaerobic, nitrate-dependent oxidation of U(IV) oxide minerals by the chemolithoautotrophic bacterium Thiobacillus denitrificans. Appl Environ Microbiol 71:2170-4. Benjamini, Y., and Y. Hochberg. 1995. Controlling the False Discovery Rate: A Practical and Powerful Approach to Multiple Testing. J. R. Stat. Soc. B 57:289- 300. Brodie, E. L., T. Z. Desantis, D. C. Joyner, S. M. Back, J. T. Larsen, G. L. Andersen, T. C. Hazen, P. M. Richardson, D. J. Herman, T. K. Tokunaga, J. M. Wan, and M. K. Firestone. 2006. Application of a high-density oligonucleotide microarray approach to study bacterial population dynamics during uranium reduction and reoxidation. Appl Environ Microbiol 72:6288-98. Brodie, E. L., T. Z. Desantis, D. C. Joyner, S. M. Back, J. T. Larsen, G. L. Andersen, T. C. Hazen, P. M. Richardson, D. J. Herman, T. K. Tokunaga, J. M. Wan, and M. K. Firestone. 2006. Application of a high-density oligonucleotide microarray approach to study bacterial population dynamics during uranium reduction and reoxidation. Appl. Environ. Microbiol. 72:6288-98. 118 10. ll. 12. 13. 14. 15. 16. Brooks, S. C. 2001. Waste Characteristics of the Former S-3 ponds and outline of uranium chemistry relevant to NABIR Field Research Center studies ORNL/TM- 2001/27. NABIR Field Research Center. Cannone, J. J., S. Subramanian, M. N. Schnare, J. R. Collett, L. M. D'Souza, Y. Du, B. Feng, N. Lin, L. V. Madabusi, K. M. Muller, N. Pande, Z. Shang, N. Yu, and R. R. Gutell. 2002. The comparative RNA web (CRW) site: an online database of comparative sequence and structure information for ribosomal, intron, and other RNAs. BMC Bioinformatics 3:2. Cardenas, E., W. M. Wu, M. B. Leigh, J. Carley, S. Carroll, T. Gentry, J. Luo, D. Watson, B. Gu, M. Ginder-Vogel, P. K. Kitanidis, P. M. Jardine, J. Zhou, C. S. Criddle, T. L. Marsh, and J. M. Tiedje. 2008. Microbial communities in contaminated sediments, associated with bioremediation of uranium to submicromolar levels. Appl Environ Microbiol 74:3718-29. Chao, A., R. L. Chazdon, R. K. Colwell, and T. J. Shen. 2006. Abundance- based similarity indices and their estimation when there are unseen species in samples. Biometrics 62:361-71. Coates, J. D., D. J. Ellis, C. V. Gaw, and D. R. Lovley. 1999. Geothrix fermentans gen. nov., sp. nov., a novel Fe(III)-reducing bacterium from a hydrocarbon-contaminated aquifer. Int. J. Syst. Bacteriol. 49 Pt 4: 1615-22. Coe, R. K. a. R. 2005. Tree diversity analysis. A manual and software for common statistical methods for ecological and biodiversity studies. World Agroforestry Centre (ICRAF), Nairobi (Kenya). Cole, J. R., Q. Wang, E. Cardenas, J. Fish, B. Chai, R. J. Farris, A. S. Kulam-Syed-Mohideen, D. M. McGarrell, T. Marsh, G. M. Garrity, and J. M. Tiedje. 2009. The Ribosomal Database Project: improved alignments and new tools for rRNA analysis. Nucleic Acids Res 37:D141-5. Cord-Ruwisch, R., D. R. Lovley, and B. Schink. 1998. Growth of geobacter sulfurreducens with acetate in syntrophic cooperation with hydrogen-oxidizing anaerobic partners. Appl Environ Microbiol 64:2232-6. Dufrene, M., and P. Legendre. 1997. Species Assemblages and Indicator Species: The Need for a Flexible Asymmetrical Approach. Ecological Monographs 67 :345-366. 119 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. Dykhuizen, D. E. 1998. Santa Rosalia revisited: why are there so many species of bacteria? Antonie Van Leeuwenhoek 73:25-33. Finneran, K. T., M. E. Housewright, and D. R. Lovley. 2002. Multiple influences of nitrate on uranium solubility during bioremediation of uranium- contaminated subsurface sediments. Environ. Microbiol. 4:510-6. Garrity, G., J. Staley, D. Boone, R. Castenholz, D. Brenner, and N. Krieg (ed.). 2005. Bergey's Manual of Systematic Bacteriology, Second Edition ed, vol. Volume Two: The Proteobacteria. Springer — Verlag. Gu, B., W. M. Wu, M. A. Ginder-Vogel, H. Yan, M. W. Fields, J. Zhou, S. Fendorf, C. S. Criddle, and P. M. Jardine. 2005. Bioreduction of uranium in a contaminated soil column. Environ. Sci. Technol. 39:4841-7. Hansen, M., T. Tolker-Nielsen, M. Givskov, and S. Molin. 1998. Biased 16S rDNA PCR amplification caused by interference from DNA flanking the template region. FEMS Microbiology Ecology 26:141-149. Hazen, T., L. Jiménez, G. Lopez de Victoria, and C. Fliermans. 1991. Comparison of bacteria from deep subsurface sediment and adjacent groundwater. Microbial Ecology 22:293-304. Hazen, T. C., and H. H. Tabak. 2005. Developments in Bioremediation of Soils and Sediments Polluted with Metals and Radionuclides: 2. Field Research on Bioremediation of Metals and Radionuclides. Reviews in Environmental Science and Biotechnology V4:157-183. Hemme, C. L., T. J. Gentry, L. Wu, M. W. Fields, C. Detter, K. Barry, D. Watson, C. W. Schadt, P. Richardson, T. Hazen, J. Tiedje, R. 8, E., and J. Zhou. 2006. Analysis of a Microbial Community Metagenome from Uranium- Contaminated Groundwater, Florida. Holmes, D. E., K. T. Finneran, R. A. O'Neil, and D. R. Lovley. 2002. Enrichment of members of the family Geobacteraceae associated with stimulation of dissimilatory metal reduction in uranium-contaminated aquifer sediments. Appl. Environ. Microbiol. 68:2300-6. Hong, S. H., J. Bunge, S. O. Jeon, and S. S. Epstein. 2006. Predicting microbial species richness. Proc Natl Acad Sci U S A 103:117-22. 120 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. Hwang, C., W. Wu, T. J. Gentry, J. Carley, G. A. Corbin, S. L. Carroll, D. B. Watson, P. M. Jardine, J. Zhou, C. S. Criddle, and M. W. Fields. 2009. Bacterial community succession during in situ uranium bioremediation: spatial similarities along controlled flow paths. ISME J 3:47-64. Hwang, C., W. M. Wu, T. J. Gentry, J. Carley, S. L. Carroll, C. Schadt, D. Watson, P. M. Jardine, J. Zhou, R. F. Hickey, C. S. Criddle, and M. W. Fields. 2006. Changes in bacterial community structure correlate with initial operating conditions of a field-scale denitrifying fluidized bed reactor. Appl. Microbiol. Biotechnol. 71 :748-60. Istok, J. D., J. M. Senko, L. R. Krumholz, D. Watson, M. A. Bogle, A. Peacock, Y. J. Chang, and D. C. White. 2004. In situ bioreduction of technetium and uranium in a nitrate-contaminated aquifer. Environ Sci Technol 38:468-75. Kodama, Y., and K. Watanabe. 2004. Sulfuricurvum kujiense gen. nov., sp. nov., a facultatively anaerobic, chemolithoautotrophic, sulfur-oxidizing bacterium isolated from an underground crude-oil storage cavity. Int J Syst Evol Microbiol 54:2297-2300. Krieg, N. R., and J. G. Holt. 1984. Bergey's manual of systematic bacteriology. Williams & Wilkins, Baltimore. Legendre, P., and E. Gallagher. 2001. Ecologically meaningful transformations for ordination of species data. Oecologia 129:271-280. Lovley, D. R., and E. J. Phillips. 1992. Reduction of uranium by Desulfovibrio desulfuricans. Appl. Environ. Microbiol. 58:850-856. Lovley, D. R., E. J. P. Phillips, Y. A. Gorby, and E. R. Landa. 1991. Microbial reduction of uranium. Nature 350:413-416. Luo, J., W.-M. Wu, J. Carley, C. Ruan, B. Gu, P. M. Jardine, C. S. Criddle, and P. K. Kitanidis. 2007. Hydraulic performance analysis of a multiple injection-extraction well system. Journal of Hydrology 336:294-302. Marciano-Cabral, F., and G. Cabral. 2003. Acanthamoeba spp. as Agents of Disease in Humans. Clin. Microbiol. Rev. 16:273-307. 121 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. McDonald, J. E., R. J. Lockhart, M. J. Cox, H. E. Allison, and A. J. McCarthy. 2008. Detection of novel Fibrobacter populations in landfill sites and determination of their relative abundance via quantitative PCR. Environ Microbiol 10:1310-9. Michalsen, M. M., B. A. Goodman, S. D. Kelly, K. M. Kemner, J. P. McKinley, J. W. Stucki, and J. D. Istok. 2006. Uranium and technetium bio- immobilization in intermediate-scale physical models of an in situ bio-barrier. Environ Sci Technol 40:7048-53. Mohanty, S. R., B. Kollah, D. B. Hedrick, A. D. Peacock, R. K. Kukkadapu, and E. E. Roden. 2008. Biogeochemical processes in ethanol stimulated uranium-contaminated subsurface sediments. Environ Sci Technol 42:43 84-90. N'Guessan, A. L., H. A. Vrionis, C. T. Resch, P. E. Long, and D. R. Lovley. 2008. Sustained removal of uranium from contaminated groundwater following stimulation of dissimilatory metal reduction. Environ Sci Technol 42:2999-3004. Nevin, K. P., and D. R. Lovley. 2000. Potential for Nonenzymatic Reduction of Fe(III) via Electron Shuttling in Subsurface Sediments. Environ. Sci. Technol. 34:2472-2478. North, N. N., S. L. Dollhopf, L. Petrie, J. D. Istok, D. L. Balkwill, and J. E. Kostka. 2004. Change in bacterial community structure during in situ biostimulation of subsurface sediment cocontaminated with uranium and nitrate. Appl. Environ. Microbiol. 70:4911-20. Nyman, J. L., T. L. Marsh, M. A. Ginder—Vogel, M. Gentile, S. Fendorf, and C. Criddle. 2006. Heterogeneous response to biostimulation for U(VI) reduction in replicated sediment microcosms. Biodegradation V17:303-316. O'Loughlin, E. J., S. D. Kelly, R. E. Cook, R. Csencsits, and K. M. Kemner. 2003. Reduction of uranium(VI) by mixed iron(II)/iron(III) hydroxide (green rust): formation of UO2 nanoparticles. Environ Sci Technol 37:721-7. Riley, R. G., and J. M. Zachara. 1992. Chemical contaminants on DOE lands and selection of contaminant mixtures for subsurface science research. US. Department of Energy, Washington, DC. Roberts, D. W. 2007. Labdsv: Ordination and Multivariate Analysis for Ecology. 122 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. Robertson, W. J., J. P. Bowman, P. D. Franzmann, and B. J. Mee. 2001. Desulfosporosinus meridiei sp. nov., a spore-forming sulfate-reducing bacterium isolated from gasolene-contaminated groundwater. Int. J. Syst. Evol. Microbiol. 51:133-40. Sorek, R., Y. Zhu, C. J. Creevey, M. P. Francino, P. Bork, and E. M. Rubin. 2007. Genome-wide experimental determination of barriers to horizontal gene transfer. Science 318:1449-52. Spain, A. M., A. D. Peacock, J. D. Istok, M. S. Elshahed, F. Z. Najar, B. A. Roe, D. C. White, and L. R. Krumholz. 2007. Identification and isolation of a Castellaniella species important during biostimulation of an acidic nitrate- and uranium-contaminated aquifer. Appl Environ Microbiol 73:4892-904. Stevens, J. O. a. R. K. a. P. L. a. B. O. H. a. M. H. H. 2007. Vegan: Community Ecology Package. Storey, J., J. Taylor, and D. Siegmund. 2004. Strong control, conservative point estimation and simultaneous conservative consistency of false discovery rates: a unified approach. Journal of the Royal Statistical Society Series B 66: 187-205. Storey, J., and R. Tibshirani. 2003. Statistical significance for genomewide studies. Proceedings of the National Academy of Sciences of the United States of America 100:9440-9445. Suzuki, Y., S. D. Kelly, K. M. Kemner, and J. F. Banfield. 2003. Microbial populations stimulated for hexavalent uranium reduction in uranium mine sediment. Appl Environ Microbiol 69:1337-46. Vrionis, H. A., R. T. Anderson, I. Ortiz-Bernad, K. R. O'Neill, C. T. Resch, A. D. Peacock, R. Dayvault, D. C. White, P. E. Long, and D. R. Lovley. 2005. Microbiological and geochemical heterogeneity in an in situ uranium bioremediation field site. Appl Environ Microbiol 71:6308-18. Wall, J. D., and L. R. Krumholz. 2006. Uranium reduction. Annu Rev Microbiol 60: 149-66. Wan, J., T. K. Tokunaga, E. Brodie, Z. Wang, Z. Zheng, D. Herman, T. C. Hazen, M. K. Firestone, and S. R. Sutton. 2005. Reoxidation of bioreduced uranium under reducing conditions. Environ. Sci. Technol. 39:6162-9. 123 57. 58. 59. 60. 61. 62. Wang, Q., G. M. Garrity, J. M. Tiedje, and J. R. Cole. 2007. Naive Bayesian classifier for rapid assignment of rRNA sequences into the new bacterial taxonomy. Appl Environ Microbiol 73:5261-7. Wu, W. M., J. Carley, M. Fienen, T. Mehlhorn, K. Lowe, J. Nyman, J. Luo, M. E. Gentile, R. Rajan, D. Wagner, R. F. Hickey, B. Gu, D. Watson, 0. A. Cirpka, P. K. Kitanidis, P. M. Jardine, and C. S. Criddle. 2006. Pilot-scale in situ bioremediation of uranium in a highly contaminated aquifer. 1. Conditioning of a treatment zone. Environ Sci Technol 40:3978-85. Wu, W. M., J. Carley, T. Gentry, M. A. Ginder-Vogel, M. Fienen, T. Mehlhorn, H. Yan, S. Caroll, M. N. Pace, J. Nyman, J. Luo, M. E. Gentile, M. W. Fields, R. F. Hickey, B. Gu, D. Watson, 0. A. Cirpka, J. Zhou, S. F endorf, P. K. Kitanidis, P. M. Jardine, and C. S. Criddle. 2006. Pilot-scale in situ bioremediation of uranium in a highly contaminated aquifer. 2. Reduction of u(VI) and geochemical control of u(VI) bioavailability. Environ Sci Technol 40:3986-95. Wu, W. M., J. Carley, J. Luo, M. Ginder-Volgel, E. Cardenas, M. B. Leigh, C. Hwang, S. D. Kelly, C. Ruan, L. Wu, J. van Nostand, T. Gentry, K. Lowe, T. Melhourne, S. Carroll, M. W. Fields, B. Gu, D. Watson, K. M. Kemner, T. L. Marsh, J. M. Tiedje, J. Zhou, P. K. Fendorf, P. K. Kitanidis, P. M. Jardine, and C. S. Criddle. 2007. In situ bioremediation of uranium (VI) to submicromolecular levels and reoxidation by dissolved oxygen. Environ. Sci. Technol. 41:5716-5723. Wu, W. M., B. Gu, M. W. Fields, M. Gentile, Y. K. Ku, H. Yan, S. Tiquias, T. Yan, J. Nyman, J. Zhou, P. M. Jardine, and C. S. Criddle. 2005. Uranium (VI) Reduction by Denitrifying Biomass. Bioremediation 9:46-61. Yamada, T., Y. Sekiguchi, S. Hanada, H. Imachi, A. Ohashi, H. Harada, and Y. Kamagata. 2006. Anaerolinea thermolimosa sp. nov., Levilinea saccharolytica gen. nov., sp. nov. and Leptolinea tardivitalis gen. nov., sp. nov., novel filamentous anaerobes, and description of the new classes Anaerolineae classis nov. and Caldilineae classis nov. in the bacterial phylum Chloroflexi. Int J Syst Evol Microbiol 56: 133 1-40. 124 CHAPTER FOUR Cardenas, E., and J. M. Tiedje. 2008. New tools for discovering and characterizing microbial diversity. Curr Opin Biotechnol 19:544-9. 125 New tools for discovering and characterizing microbial diversity Summary To discover and characterize microbial diversity, approaches based on new sequencing technologies, novel isolation techniques, microfluidics and metagenomics among others are being used. These approaches have contributed to discovery of novel genes from environmental samples, to massive characterization of functional and phylogenetic genes and to isolation of members of formerly uncultured yet ubiquitous groups like Verrucomicrobia, Acidobacteria, OP10 and methanogenic Archaea. Cheaper sequencing is key in this process by making available applications that were previously restricted to big research centers, complementing previously available methodologies and potentially replacing some of them. The new tools are reshaping the way we study the environment, increasing the resolution at which microbial communities, their complexities and dynamics, can be studied to reveal their genetic potential and their functional diversity. Introduction Discovering new microbes and characterizing their functions are major goals in the study of microbial diversity. Historically, this was achieved through cultivation and subsequent characterization of strains. Decades later, culture-independent methods have provided new tools to study the microbial world. Recently, metagenomics is providing a gene-based exploration of the community as a whole while isolation and cultivation 126 continue as the main-stay for testing metabolic abilities and more detailed genomic studies of individual members of this community. As more and more environments were sampled, it became evident that the majority of the microbes have yet to be cultured (30), driving the development of new methods to unveil the unknown (Figure 4.1). These methods benefit from advances in sequencing, isolation and microfluidics and allow researchers to study microbial communities in their natural environments and with higher resolution. 127 I ‘_ . _ I 1 ,3 “*"-m“‘*"“‘ 'A ‘ * ' «g. .g..mu‘:'m W Mao” await-mu. ,.. w. . o W ’\.~ P ‘I w a, r . I '~ - . r" , 1 Community Structure 1 Genetic Potential ' (Who is there?) FE (What can they do?) ' i . - Phylogenetic arrays i - Metagenomics " ' 16S rRNAtag sequencing l E - Functional gene arrays - Community Image analysis 1 ‘ - Neural Gene Networks l. (CMEIAS) - E 3*- ~ . a} .gA‘Ifi'éh.:A-utouur, M. «New; .23; 5.1-...)"13'Jui‘; .. -33... g; ’2’;.“Qmsru .-.-~.u.-...x-...-e.. fir K; .‘ 6' Gene expression Functional characterization (What are they doing?) '3 4‘1. — Novel isolation methods - Metatranscriptomics ~ - Stable isotope probing - Metaproteom . ' NaHOSIMS ; , . CARD-FISH Fig: 4.1. Novel methods to discover 21:1 ficharfiacterize microbial diversity. In boldjr: the methods reviewed in this article. The image showing the morphological diversity from an anaerobic bioreactor community is courtesy of Frank Dazzo. Impact of the new sequencing technologies in microbial ecology As we write this report, new sequencing technologies are changing the way we study microbial communities. Traditionally, the study of genes from natural environments included cloning DNA into a vector, inserting that vector into a host, screening, and Sanger sequencing. Sequence-by-synthesis methods provide faster, cheaper and simpler methods for (meta)genome sequence that bypass the cloning bias and labor-intensive Sanger method. New sequencing technologies differ in type of output, read length, speed, and error rate distribution (Table 4.1); for a more detailed review see ref. (34). Single molecule sequencing, expected soon by Helicos and Pacific Biosciences, will provide another leap in sequence capacity with the latter company 128 projecting reads as long as 10 kilobases. The different sequencers are being used for different applications: longer reads being preferred for genome sequencing and gene surveys while shorter reads being the choice for transcriptomes, small RNAs analysis and comparison of closely related strains or strain variants. These methods compensate their shortcomings in read length and error bias with massive output (and thus coverage), speed and simplicity. The new methodologies create a situation where the limitation is not the ability to produce sequence data but the ability to store and analyze it in new revealing ways. Table 4.1. Currently available sequencing technologies and their features. Not included in the table are technologies under development by Visigen Biotechnologies, Polonator, Pacific Biosciences, Intelligent Biosystems, Complete Genomics. N/A, not available. Method Output per run Read length System described (per day) (bases) 52-86 kB ABI 3730xl DNA sang“ (0.69-2.1MB) 550'900 Analyzer. 96 Well Roche 454 Life Sciences Pyrosequencing 400-600 MB 400 GS FLX Titanium Series Reagents . N/A The HeliScope Single Helicos (600-2160 MB) 30 Molecule Sequencer SOLiD 6 GB (600MB) 35 SOLiD system Illumina Illumina Genome (Solexa) 3GB (600MB) 36 Analyzer Currently pyrosequencing is favored for microbial genome sequencing and metagenomic studies. It is valuable for sequencing regions that are technically difficult (e. g. due to strong secondary structure or high GC content) and for covering regions recalcitrant to cloning in Escherichia coli (36). In one example, a Sanger/pyrosequencing hybrid approach completely eliminated gaps in two genomes and reduced gaps by 86% in 129 four other genomes (12). In the large sequencing centers, the Sanger method is being replaced as the main sequencing method for bacterial genomes though it remains an important tool for closing them. See also (32) for a more historical perspective of 454 sequencing. A major advantage of the new sequencing technologies is the substantial reduction in cost, democratizing genomics and making massive gene surveys feasible as a way to discover new microbes and genes. This is especially useful for the study of microbial communities by 16S rRNA gene analysis, the most commonly used species proxy. By combining primers that target hypervariable regions of the 168 rRN A gene with nucleotide barcodes, it is now more feasible to survey environments with thousands of sequences at the time. The hypervariable regions targeted are short enough (100-3 50 bases) to be covered by some of the new sequencing technologies yet long enough to be informative for classification by the current rRNA databases (e.g. Ribosomal database project (RDP), Greengenes, Silva), though they are poor for phylogeny in comparison to the full length 168 rRN A gene (~1500 bases). The RDP has reported accuracy in classification of partial sequences to the genus level for 400-base reads and to the family level for ZOO-base reads (44). The first in the new wave of massive 16S rRNA gene studies was of deep sea sediments using the V6 hypervariable region (35). Since then, other regions have been used to study temperate and tropical soils (31). Highlighting the magnitude of the species diversity question, in one case 750,000 sequences were not enough to recover all the microbial diversity present in deep sea communities near hydrothermal vents (15). Longer reads will allow targeting of regions that provide a 130 better classification (44) and to use the sequences for more accurate phylogenetic studies. A complementary approach to massive sequencing is the adoption of nucleotide barcodes in the amplification primers. In this method, samples from different origins can be mixed in one run and after sequencing their data can be separated according to their barcode (22, 25). This approach decreases the cost per sample since more samples can be pooled in the sequencing run rather than sequencing fewer samples to greater depth. For example, using 454 F LX pyrosequencing 400,000 reads are usually generated; by using 80 barcodes, 5000 reads per sample can be obtained. This number of sequences is at least one order of magnitude higher than those from traditional clone libraries. Transcriptomics is also being transformed by the new sequencing technologies. Gene prediction algorithms, probe design, cloning, and hybridization steps involved in traditional transcriptomic analyses can all be bypassed by massively parallel direct sequencing of cDNA and subsequently mapping the reads to a reference genome to quantify known transcripts and even detect putative new genes. For example, Illumina (Solexa) sequencing has been used to detect genes, intron usage and alternative initiation codons in yeast (23). In a more recent article, 512 new genes were predicted for the nitrogen-fixing plant symbiont Sinorhizobium meliloti after a transcriptome analysis that used pyrosequencing (19). A new look at metagenomics Metagenomics, the study of microbial communities based on its genetic material (6), has paved the way for the discovery of new genes, proteins and biochemical 131 pathways. Genes involved in anaerobic methane oxidation were discovered with a metagenomics effort that targeted an association of Archaea and sulfate-reducing bacteria (13). Metagenomics has proved more valuable in recovering complete genomes in systems with low diversity, such as the acid drainage from an abandoned mine (43) or in highly enriched cultures such as the first genome of the annamox (anaerobic ammonia oxidation) group, Kuenenia stuttgartiensis (41). When applied to more diverse environments, such as soils, the yield of assembled genomes has been poor (3 8). However, assembly is not necessary to make some important inferences. One particularly valuable use of a metagenome can be as the basis for metranscriptomics (9) and metaproteomics (28). Array-based approaches for studying microbial diversity The study of the microbial community structure and its functional diversity has benefited from the use of microarrays that target phylogenetic markers and functional genes. Arrays provide high throughput, specificity and relative quantification free of cloning bias. A functional gene array, Geochip (14), has been helpful in the study of nitrogen and carbon cycles in Antarctic soils (47) and PCB contaminated soils (18). Arrays targeting the 16S rRN A gene have been used to study microbial communities involved in uranium reduction and reoxidation (4). Phylogenetic arrays easily detect more microbial diversity than traditional clone libraries (7). Nevertheless, probe design limits detection to what is already known, or what can be placed on a chip, and hence microarrays are tools better suited to generally characterize ecosystems rather than to discover new genes and functions. 132 Microfluidics, manipulation at micro(be) scale Microfluidics, the manipulation of small volumes, is becoming a powerful tool in microbial ecology. By working at dimensions relevant to microbes it has been possible to study specific microbes via isolation, gene amplification or genome reconstruction. Microfluidics allowed for individual members of termite hindgut flora to be sorted and screened by multiplex PCR (24). By targeting the 16S rRNA gene and functional genes it was possible to unambiguously link phylogeny and function (no cultivation required), and to detect horizontal gene transfer events (24). This approach has been coupled with whole-genome amplification steps that target more genes and to sequence much of an organism (3 7). A genome from the candidate phylum TM7 has been reconstructed from human mouth samples with the help of a microfluidic device that selected for rod-like morphology, amplified individual genomes and screened for TM7 signatures (20). The formerly unculturables — New approaches for isolation Renewed efforts in isolation are being taken to capture more of the microbial functional diversity. These novel methods demonstrate that culturing is often possible if niche conditions are mimicked. In contrast with traditional cultivation, the novel isolation conditions try to better reflect natural environments in terms of nutrients (composition and concentration), oxygen levels, pH and natural associations with other microorganisms. These novel isolation methods also benefit from advances in microfluidics. 133 Novel lineages of Actinobacteria, Acidobacteria, Proteobacteria and Verrucomicrobia were isolated using diluted nutrient broth solidified with gellan (16). Gellan is a good alternative solidifying reagent producing gels that are clearer (easier screening) and more stable than agar under a wider range of temperatures and pHs. Gellan also worked as a carbon source when isolating the first cultured member of the OP10 division from soils (40). In one study, colony counts increased up to 21.3 times when replacing agar with gellan in culture medium (42). In terms of oxygen concentration, microaerophillic conditions have led to increased recovery of termite gut bacteria (46) and soil Acidobacteria (8). In situ and trap cultivation introduce sterile media to natural environments to use local conditions for enrichment. Using in situ cultivation with diffusion chambers, the diversity of isolates recovered from freshwater sediments increased and isolates from some rarely cultivated groups, such as Acidobacteria and Verrucomicrobia, were recovered (3). Trap cultivation has been used to capture novel soil Actinobacteria, known secondary metabolite producers, whose filaments penetrate the medium when growing (10). Because organisms rarely live in pure cultures, cocultivation is another strategy for isolation. This is well established for hydrogenotrophic methanogenic Archaea and their partners. In these associations Archaea consume hydrogen, a product of incomplete substrate oxidation in anaerobic environments and make the oxidation reaction thermodynamically feasible (17). Through cocultivation, a globally abundant methanogen 134 (with a propionate-oxidizing, H2 producing partner) (33), and an anaerobic phenol degrader (and its hydrogen consuming Archaeal syntroph) (27) have been isolated. Carbon exchange occurring in natural associations can also be used for isolation purposes. Novel heterotrophic partners of algae have been isolated by adding algae- excreted organic material to culture media (21) or by creating defined media to simulate the exudates’ composition (45). Another novel isolation approach uses of a combination of high throughput screening, microfluidics and dilution-to-extinction techniques originally used by Casida (5). In a current version of this approach(39), small volumes of inocula, sometimes picoliters, are added to sterile medium, typically in a 96-well plate format. By having few cells per well, the competition with fast growers is greatly reduced. After incubation different preparations can be screened for growth using fluorescent staining or group- specific probes. The most dominant group in marine ecosystems (SARll clade) (29) and novel isolates of Proteobacteria, Actinobacteria, F irmicutes and F lavobacteria- Cytophaga (11) have been obtained using this approach. There appears to be no reason why most of the formerly unculturable microbes cannot be made culturable with sufficient study, cleverness and patience. Study of cultured isolates remains the most straightforward method to characterize an organism’s functions and evaluate its biotechnological potential. Metatranscriptomics and metaproteomics, fields in their infancy 135 The study of gene expression and translation into proteins within natural environments are two emerging fields in microbial ecology which hold special promise for studying function. Recovering mRN A and proteins from natural environments is challenging, yet several groups have pace-setting examples. The first metatranscriptome study recovered transcripts encoding Archaea and Bacteria genes involved in carbon, sulfur and nitrogen cycling from marine and freshwater bacterioplankton communities (26). A marine metatranscriptome study found that hypothetical genes were among the most highly expressed in Prochlorococcus spp. and that a large fraction of the transcripts recovered were novel, indicating the immense metabolic diversity hidden in the oceans (9). In a soil transcriptome, even though most of the transcripts coded for housekeeping proteins, genes linked to soil processes were also found (1). In addition, 32% of eukaryote transcripts recovered carried novel genes (1), again highlighting the vast unknown within hypothetical genes. In the case of metaproteomics, proteins identified as chlorocatechol dioxygenases were recovered from a chlorobenzene-contaminated aquifer, a finding that agrees with the predicted degradation pathway of chlorobenzene (2). Many challenges remain in metagenomic, metatranscriptomic and metaproteomic approaches, including DNA, RNA and protein extraction, mRN A instability, low abundance, and low proportion of mRNA in total RNA. As these are still novel approaches, new solutions to these technical problems are expected to arise, which would allow them to reach their full potential of characterizing functional diversity. Databases and software tools (Table 4.2) are essential and their further development is needed to 136 deal with the growing bottleneck in the analysis of metagenomic, metatranscriptomic and metaproteomic data. Table 4.2. Selected resources available for metagenomics, metatranscriptomics and mass spectrometry data to identify proteins from primary sequence databases metaproteomics. Resource Features Website Data repository and CAMERA bioinformatics tools httgflcameracal illnet resource Integrated Microbial . Genomes with Comparative . . . . . . . . . . s , r ’ / Microbiome metagenome analysrs http/hmgjgdoegov, CQl-bm/m.» 1113”].ng and annotation tools samples (IMG/M) Human Oral Database on bacteria Microbiome from the human oral h_ttp://www.h0md.org Database cavity SEED Comparative genomics http://wwwtheseedoru envuonment ' Metagenome Analysis http://www-ab.infomiatikuni- MEGAN . , - . ,. , Software tuebmgen .de/sottwai e/ m «3221 n/ Automated 33%c;mfs ifs‘intagggllliciing analysrs http://metagenomics.nmpdr.org/ environment PFAM Collection of protein http://pfamjaneliaorgj families using multiple sequence alignments and hidden Markov models Phylofacts Structural http://phylogenomics.berkelevedu/phvlofacts/indexphp phylogenomic encyclopedia for protein functional and structural classification Uniprot Protein sequence and http://wwwuniprot.org1 fimctional information Mascot Search engine that uses http://www.matrixscience.comKhome.html 137 Future challenges and directions Massively parallel sequencing, metagenomics, metatranscriptomics, metaproteomics, and isolation will allow us to study deeper and deeper layers of the microbial diversity, including genes behind novel or valued functions. These methods promise to broaden the use of (meta)genomics and offer the chance to study specific groups that were previously unknown and uncharacterized. We can now ask more complex questions such as what is the role of low abundant organisms?, can we estimate the total diversity of the system and assess its value?, how important is it to know in what organism particular genes reside?, and more importantly, how can we use this information to better characterize, manage and/or harvest even greater functional diversity. The ball is in our court. Reference annotations ° of special interest ” of outstanding interest Sogin ML, Morrison HG, Huber JA, Mark Welch D, Huse SM, Neal PR, Arrieta JM, Hemdl GJ: Microbial diversity in the deep sea and the underexplored. Proc Natl AcadSci USA 2006, 103:12115-12120. "First paper on massive parallel l6S rRNA gene surveys. Deep sea samples were studied by targeting the V6 hypervariable region. By comparing against a custom V6 database, it 138 was found that 80% of the sequences were divergent by 10% or more, suggesting an immense diversity present in rare members of the community. Nagalakshmi U, Wang Z, Waern K, Shou C, Raha D, Gerstein M, Snyder M: The transcriptional landscape of the yeast genome defined by RNA sequencing. Science 2008, 320:1344-1349. "Transcriptomic study using massive parallel sequence of cDNA. By comparing the transcripts against the yeast genome it was possible to detect alternative splicing, initiation sites and delimit the boundaries of genes. Mao C, Evans C, Jensen RV, Sobral BW: Identification of new genes in Sinorhizobium meliloti using the Genome Sequencer F LX system. BMC Microbiol 2008, 8:72. - Massive sequencing of cDNA using pyrosequencing. Putative new genes were found, these genes were likely missed by prediction algorithm because they were short or had low scores against known databases. F rias-Lopez J, Shi Y, Tyson GW, Coleman ML, Schuster SC, Chisholm SW, DeLong EF: Microbial community gene expression in ocean surface waters. Proc Natl Acad Sci USA 2008, 105:3805-3810. "A study of DNA and cDNA from the same marine community. Proof that gene expression in complex systems if possible by mapping the transcripts to the metagenome. Interestingly, some of the more highly expressed genes were hypothetical and a large 139 fraction of cDNA reads were not present in databases, suggesting that large part of the microbial diversity remains hidden in the oceans. He Z, Gentry TJ, Schadt CW, Wu L, Liebich J, Chong SC, Huang Z, Wu W, Gu B, Jardine P, et al.: GeoChip: a comprehensive microarray for investigating biogeochemical, ecological and environmental processes. ISME J 2007, 1:67-77. °Description of the functional gene array Geochip. Brodie EL, Desantis TZ, Joyner DC, Baek SM, Larsen JT, Andersen GL, Hazen TC, Richardson PM, Herman DJ, Tokunaga TK, et al.: Application of a high-density oligonucleotide microarray approach to study bacterial population dynamics during uranium reduction and reoxidation. Appl Environ Microbiol 2006, 72:6288-6298. °First example of the use of high-density phylogenetic arrays to profile microbial communities in an uranium contaminated aquifer. The results suggest a role of member of the Geobacter spp. Geothrixfermentans in uranium reduction. Stepanauskas R, Sieracki ME: Matching phylogeny and metabolism in the uncultured marine bacteria, one cell at a time. Proc Natl Acad Sci U S A 2007, 104:9052-9057. °An approach that combines flow cytometry, whole genome amplification and screening to allow single cells to be separated and their genome amplified. The method allows targeting metabolic and phylogenetic markers in the same uncultured microbe. 140 Marcy Y, Ouvemey C, Bik EM, Losekann T, Ivanova N, Martin HG, Szeto E, Platt D, Hugenholtz P, Relman DA, et al.: Dissecting biological "dark matter" with single-cell genetic analysis of rare and uncultivated TM7 microbes from the human mouth. Proc Natl Acad Sci USA 2007, 104:11889-11894. "A microfluidic device that allows selection of specific uncultured bacterium and amplification and recovery of their genome. Stott MB, Crowe MA, Mountain BW, Smirnova AV, Hou S, Alam M, Dunfield PF: Isolation of novel bacteria, including a candidate division, from geothermal soils in New Zealand. Environ Microbiol 2008. - A fine example of novel isolation methods that combine low nutrient conditions, and alternative gelling agents to recover novel bacteria from soils. Bollmann A, Lewis K, Epstein SS: Incubation of environmental samples in a diffusion chamber increases the diversity of recovered isolates. Appl Environ Microbiol 2007, 73:63 86-6390. 0 By using a diffusion chamber incubated in situ, rarely recovered groups from Deltaproteobacteria, Verrucomicrobia, Spirochaetes, and Acidobacteria were isolated. 141 Stingl U, Tripp HJ, Giovannoni SJ: Improvements of high-throughput culturing yielded novel SARll strains and other abundant marine bacteria from the Oregon coast and the Bermuda Atlantic Time Series study site. ISME J 2007, 1:361-371. 0- Excellent example of high-throughput dilution-to-extinction culturing yielding novel isolates from the globally abundant SARll group. Acknowledgments This work was supported by the US. DOE under grants DE-FG02-97ER62469 and DE- FG02-99ER62848, and NIEHS grant 5P42 ESOO491 1-18/19. The authors thank Brian Campbell, Dieter Tourlousse and Patrick Chain for useful comments and discussion. 142 REFERENCES Bailly, J., L. F raissinet-Tachet, M. C. Vemer, J. C. Debaud, M. Lemaire, M. Wesolowski-Louvel, and R. Marmeisse. 2007. Soil eukaryotic functional diversity, a metatranscriptomic approach. ISME J 1:632-42. Benndorf, D., G. U. Balcke, H. Harms, and M. van Bergen. 2007. Functional metaproteome analysis of protein extracts from contaminated soil and groundwater. ISME J 1:224-34. Bollmann, A., K. Lewis, and S. S. Epstein. 2007. Incubation of environmental samples in a diffusion chamber increases the diversity of recovered isolates. Appl Environ Microbiol 73:63 86-90. Brodie, E. L., T. Z. Desantis, D. C. Joyner, S. M. Baek, J. T. Larsen, G. L. Andersen, T. C. Hazen, P. M. Richardson, D. J. Herman, T. K. Tokunaga, J. M. Wan, and M. K. Firestone. 2006. Application of a high-density oligonucleotide microarray approach to study bacterial population dynamics during uranium reduction and reoxidation. Appl Environ Microbiol 72:6288-98. Casida, L. E., Jr. 1965. Abundant Microorganism in Soil. Appl Microbiol 13:327-34. Committee on Metagenomics: Challenges and Functional Applications, N. R. C. 2007. The New Science of Metagenomics: Revealing the Secrets of our Microbial Planet. National Academies Press, Washington, DC. DeSantis, T. 2., E. L. Brodie, J. P. Moberg, I. X. Zubieta, Y. M. Piceno, and G. L. Andersen. 2007. High-density universal 16S rRNA microarray analysis reveals broader diversity than typical clone library when sampling the environment. Microb Ecol 53:371-83. Eichorst, S. A., J. A. Breznak, and T. M. Schmidt. 2007. Isolation and characterization of soil bacteria that define Terriglobus gen. nov., in the phylum Acidobacteria. Appl Environ Microbiol 73:2708-17. 143 10. 11. 12. 13. 14. 15. 16. 17. Frias-Lapez, J., Y. Shi, G. W. Tyson, M. L. Coleman, S. C. Schuster, S. W. Chisholm, and E. F. Delong. 2008. Microbial community gene expression in ocean surface waters. Proc Natl Acad Sci U S A 105:3805-10. Gavrish, E., A. Bollmann, S. Epstein, and K. Lewis. 2008. A trap for in situ cultivation of filamentous actinobacteria. J Microbiol Methods 72:257-62. Gich, F., K. Schubert, A. Bruns, H. Hoffelner, and J. Overmann. 2005. Specific detection, isolation, and characterization of selected, previously uncultured members of the freshwater bacterioplankton community. Appl Environ Microbiol 71:5908-19. Goldberg, S. M., J. Johnson, D. Busam, T. Feldblyum, S. Ferriera, R. Friedman, A. Halpern, H. Khauri, S. A. Kravitz, F. M. Laura, K. Li, Y. H. Rogers, R. Strausberg, G. Sutton, L. Tallan, T. Thomas, E. Venter, M. Frazier, and J. C. Venter. 2006. A Sanger/pyrosequencing hybrid approach for the generation of high-quality draft assemblies of marine microbial genomes. Proc Natl Acad Sci U S A 103:11240-5. Hallam, S. J., N. Putnam, C. M. Preston, J. C. Detter, D. Rokhsar, P. M. Richardson, and E. F. DeLong. 2004. Reverse methanogenesis: testing the hypothesis with environmental genomics. Science 305: 1457-62. He, Z., T. J. Gentry, C. W. Schadt, L. Wu, J. Liebich, S. C. Chang, Z. Huang, W. Wu, B. Gu, P. Jardine, C. Criddle, and J. Zhou. 2007. GeoChip: a comprehensive microarray for investigating biogeochemical, ecological and environmental processes. ISME J 1:67-77. Huber, J. A., D. B. Mark Welch, H. G. Morrison, S. M. Huse, P. R. Neal, D. A. Butterfield, and M. L. Sogin. 2007. Microbial population structures in the deep marine biosphere. Science 318:97-100. Janssen, P. H., P. S. Yates, B. E. Grintan, P. M. Taylor, and M. Sait. 2002. Improved culturability of soil bacteria and isolation in pure culture of novel members of the divisions Acidobacteria, Actinobacteria, Proteobacteria, and Verrucomicrobia. Appl Environ Microbiol 68:2391-6. Kamagata, Y., and H. Tamaki. 2005. Cultivation of Uncultured Fastidiaus Microbes. Microbes and Environments 20:85-91. 144 18. 19. 20. 21. 22. 23. 24. 25. 26. Leigh, M. B., V. H. Pellizari, O. Uhlik, R. Sutka, J. Rodrigues, N. E. Ostrom, J. Zhou, and J. M. Tiedje. 2007. Biphenyl-utilizing bacteria and their fimctional genes in a pine root zone contaminated with polychlorinated biphenyls (PCBs). ISME J 1:134-48. Mao, C., C. Evans, R. V. Jensen, and B. W. Sobral. 2008. Identification of new genes in Sinorhizobium meliloti using the Genome Sequencer F LX system. BMC Microbial 8:72. Marcy, Y., C. Ouverney, E. M. Bik, T. Losekann, N. Ivanova, H. G. Martin, E. Szeto, D. Platt, P. Hugenholtz, D. A. Relman, and S. R. Quake. 2007. Dissecting biological "dark matter" with single-cell genetic analysis of rare and uncultivated TM7 microbes from the human mouth. Proc Natl Acad Sci U S A 104:11889-94. Mayali, X., P. J. Franks, and F. Azam. 2008. Cultivation and ecosystem role of a marine raseabacter clade-affiliated cluster bacterium. Appl Environ Microbial 74:2595-603. Meyer, M., U. Stenzel, S. Myles, K. Prufer, and M. Hofreiter. 2007. Targeted high-throughput sequencing of tagged nucleic acid samples. Nucleic Acids Res 35:e97. Nagalakshmi, U., Z. Wang, K. Waern, C. Shou, D. Raha, M. Gerstein, and M. Snyder. 2008. The transcriptional landscape of the yeast genome defined by RNA sequencing. Science 320:1344-9. Ottesen, E. A., J. W. Hong, S. R. Quake, and J. R. Leadbetter. 2006. Microfluidic digital PCR enables multigene analysis of individual environmental bacteria. Science 314: 1464-7. Parameswaran, P., R. Jalili, L. Tao, S. Shokralla, B. Gharizadeh, M. Ronaghi, and A. Z. Fire. 2007. A pyrosequencing-tailored nucleotide barcode design unveils opportunities for large-scale sample multiplexing. Nucleic Acids Res 35:e130. Poretsky, R. S., N. Bano, A. Buchan, G. LeCleir, J. Kleikemper, M. Pickering, W. M. Pate, M. A. Moran, and J. T. Hollibaugh. 2005. Analysis of microbial gene transcripts in environmental samples. Appl Environ Microbiol 71:4121-6. 145 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. Qiu, Y. L., S. Hanada, A. Ohashi, H. Harada, Y. Kamagata, and Y. Sekiguchi. 2008. Syntrophorhabdus aromaticivorans gen. nov., sp. nov., the first cultured anaerobe capable of degrading phenol to acetate in obligate syntrophic associations with a hydrogenotrophic methanogen. Appl Environ Microbiol 74:2051-8. Ram, R. J., N. C. Verberkmoes, M. P. Thelen, G. W. Tyson, B. J. Baker, R. C. Blake, 2nd, M. Shah, R. L. Hettich, and J. F. Banfield. 2005. Community proteomics of a natural microbial biofilm. Science 308:1915-20. Rappe, M. S., S. A. Cannon, K. L. Vergin, and S. J. Giovannoni. 2002. Cultivation of the ubiquitous SARll marine bacterioplankton clade. Nature 418:630-3. Rappe, M. S., and S. J. Giovannoni. 2003. The uncultured microbial majority. Annu Rev Microbiol 57:369-94. Roesch, L. F., R. R. Fulthorpe, A. Riva, G. Casella, A. K. Hadwin, A. D. Kent, S. H. Daroub, F. A. Camargo, W. G. Farmerie, and E. W. Triplett. 2007. Pyrosequencing enumerates and contrasts soil microbial diversity. ISME J 1:283-90. Rothherg, J. M., and J. H. Leamon. 2008. The development and impact of 454 sequencing. Nat Biotechnol 26:1 1 17-24. Sakai, 8., H. Imachi, Y. Sekiguchi, A. Ohashi, H. Harada, and Y. Kamagata. 2007. Isolation of key methanogens for global methane emission from rice paddy fields: a novel isolate affiliated with the clone cluster rice cluster 1. Appl Environ Microbiol 73:4326-31 . Shendure, J., and H. Ji. 2008. Next-generation DNA sequencing. Nat Biotechnol 26:1 135-45. Sogin, M. L., H. G. Morrison, J. A. Huber, D. Mark Welch, S. M. Huse, P. R. Neal, J. M. Arrieta, and G. J. Herndl. 2006. Microbial diversity in the deep sea and the underexplored. Proc Natl Acad Sci U S A 103:12115-20. Sorek, R., Y. Zhu, C. J. Creevey, M. P. Francino, P. Bork, and E. M. Rubin. 2007. Genome-wide experimental determination of barriers to horizontal gene transfer. Science 318:1449-52. 146 37. 38. 39. 40. 41. 42. 43. 44. Stepanauskas, R., and M. E. Sieracki. 2007. Matching phylogeny and metabolism in the uncultured marine bacteria, one cell at a time. Proc Natl Acad Sci U S A 104:9052-7. Steward, G. F., and M. S. Rappe. 2007. What's the 'meta' with metagenomics? ISME J 1:100-2. Stingl, U., H. J. Tripp, and S. J. Giovannoni. 2007. Improvements of high- throughput culturing yielded novel SAR11 strains and other abundant marine bacteria from the Oregon coast and the Bermuda Atlantic Time Series study site. ISME J 1:361-71. Stott, M. B., M. A. Crowe, B. W. Mountain, A. V. Smirnova, S. Hou, M. Alam, and P. F. Dunfield. 2008. Isolation of novel bacteria, including a candidate division, from geothermal soils in New Zealand. Environ Microbiol. Strous, M., E. Pelletier, S. Mangenot, T. Rattei, A. Lehner, M. W. Taylor, M. Horn, H. Daims, D. Bartol-Mavel, P. Wincker, V. Barbe, N. Fonknechten, D. Vallenet, B. Segurens, C. Schenowitz-Truong, C. Medigue, A. Collingro, B. Snel, B. E. Dutilh, H. J. Op den Camp, C. van der Drift, I. Cirpus, K. T. van de Pas-Schoonen, H. R. Harhangi, L. van Niftrik, M. Schmid, J. Keltjens, J. van de Vossenberg, B. Kartal, H. Meier, D. Frishman, M. A. Huynen, H. W. Mewes, J. Weissenhach, M. S. Jetten, M. Wagner, and D. Le Paslier. 2006. Deciphering the evolution and metabolism of an anammox bacterium from a community genome. Nature 440:790-4. Tamaki, H., Y. Sekiguchi, S. Hanada, K. Nakamura, N. Nomura, M. Matsumura, and Y. Kamagata. 2005. Comparative analysis of bacterial diversity in freshwater sediment of a shallow eutrophic lake by molecular and improved cultivation-based techniques. Appl Environ Microbiol 71:2162-9. Tyson, G. W., J. Chapman, P. Hugenholtz, E. E. Allen, R. J. Ram, P. M. Richardson, V. V. Solovyev, E. M. Rubin, D. S. Rokhsar, and J. F. Banfield. 2004. Community structure and metabolism through reconstruction of microbial genomes from the environment. Nature 428:37-43. Wang, Q., G. M. Garrity, J. M. Tiedje, and J. R. Cole. 2007. Naive Bayesian classifier for rapid assignment of rRNA sequences into the new bacterial taxonomy. Appl Environ Microbiol 73:5261-7. 147 45. 46. 47. Watanabe, K., M. Imase, H. Aoyagi, N. Ohmura, H. Saiki, and H. Tanaka. 2008. Development of a novel artificial medium based on utilization of algal photosynthetic metabolites by symbiotic heterotrophs. J Appl Microbiol. Wertz, J. T., and J. A. Breznak. 2007. Physiological ecology of Stenoxybacter acetivorans, an obligate microaerophile in termite guts. Appl Environ Microbiol 73:6829-41. Yergeau, E., S. Kang, Z. He, J. Zhou, and G. A. Kowalchuk. 2007. Functional microarray analysis of nitrogen and carbon cycling genes across an Antarctic latitudinal transect. ISME J 1:163-79. 148 CHAPTER FIVE CONCLUSIONS Comparative l6S rRNA gene analyses have allowed us to explore the diversity of the microbial world present in the Field Research Center Area 3 where uranium remediation is being applied. The results from both studies have suggested that bacteria related to sulfate-reducing bacteria, Desulfovibrio and Desuljbsporosinus and iron reducing bacteria are directly involved in uranium reduction in situ. Desulfovibrio seems to be the most important group given its relative abundance in wells where uranium reduction occurs, and also because the rapid conversion of ethanol to acetate observed in this site. In addition to the uranium reducing groups, other bacteria seems to contribute important functions to the community such as the ability to eliminate nitrate, a competing electron (e.g. Thauera); create resistance structure such as biolfilms (e.g. Chloroflexi), recycle nutrients (e. g. Chloroflexi, Acanthoamoeba) and provide uranium reduction indirectly through reduced iron (F RB, e. g. F erribacterium), sulfur (SRB, e. g. Desulfosporosinus) and nitrogen compound (Denitrifiers e. g Acidovorax ) as well as reduced humic acids (Geothrix). One surprising find was the abundance of iron-oxidizing populations such as T hiobacillus and Gallionella. Thiobacillus can oxidize uranium while reducing nitrate, Gallionella can oxidize iron under microaerophillic conditions, and iron oxides can directly reoxidize uranium. The effect of iron oxidizers on uranium stability needs further attention. 149 The ability of Acidobacteria iron reducers such as Geothrix, to reduce uranium also needs further study. In our experiments, Geothrix-related sequences were associated with the wells of high uranium reducing activity. These bacteria can reduced humic acids and through them potentially reduced uranium, yet the ability to reduce uranium directly is unknown. The use of Acidobacteria in uranium reduction would be a useful application given their ability to grow under a wide range of pH, resist environmental stresses and grow on oligotrophic conditions. In terms of diversity, what started as a low diversity community of a heavily contaminated site has changed to a diverse community that can perform the task we stimulated it for, uranium reduction. Even when the response was heterogeneous i.e. different uranium reducing population responded in different sites, the diverse response was partially explained by hydrology parameters. Using the hydrology information helped us to understand the gradients present in the site and better understand how the communities respond to the stimulation. We have demonstrated that hydrology information is no longer relevant only for engineering purposes but can also be used to explain the natural variability in the microbial communities. Method wise, massive parallel sequencing has shown to be powerful in detecting minor members of the community that seemed to respond to stimulation such as Anaeromyxobacter and Desulfosporosinus, it also potentially avoided cloning bias. Additionally, massive sequencing provides the opportunity to perform multivariate analyses that are powered by high sequence counts. It is expected that in the future this technology will be widely adopted for the analysis of microbial communities, this works uses for the first time a combination of massive sequencing and indicator species 150 approach to provide a consistent and statistically sound method to detect candidate species for a given groups. We expect this approach to be more adopted to deal with the high complexity of massive parallel sequencing databases. Overall this work has answered some questions regarding the identity of the populations involved in uranium reduction but has also raised new question regarding the role of non-uranium reducers on the stability of reduced uranium, and their role in the community. The massive aspect has potentially shown areas of the microbial diversity where isolation efforts should be made in order to obtain microorganism relevant to the cleanup of this site e. g. in the isolation of new Acidobacteria associated with uranium reduction. 151 APPENDIX My research has greatly benefited from a year of work in the Ribosomal Database Project where I developed and produced video tutorials to increase the adoption of the Functional Gene Database / Repository as part of the NIEHS Research translation core of the Michigan State University Superfund group. Additionally I led in the development of a new feature of the Ribosomal Database Project, Assignment Generator. This application provides instructors the chance to create a realistic assignment on comparative 16S rRNA gene analysis. These two projecta, the tutorials and the Assignment Generator, have been published as part of the Ribosomal Database Project’s papers in the journal Nucleic acid research in 2007(1) and 2009 (2), respectively. 152 REFERENCES Cole, J. R., B. Chai, R. J. Farris, Q. Wang, A. S. Kulam-Syed-Mohideen, D. M. McGarrell, A. M. Bandela, E. Cardenas, G. M. Garrity, and J. M. Tiedje. 2007. The ribosomal database project (RDP-II): introducing myRDP space and quality controlled public data. Nucleic Acids Res. 35:D169-72. Cole, J. R., Q. Wang, E. Cardenas, J. Fish, B. Chai, R. J. Farris, A. S. Kulam-Syed-Mohideen, D. M. McGarrell, T. Marsh, G. M. Garrity, and J. M. Tiedje. 2009. The Ribosomal Database Project: improved alignments and new tools for rRNA analysis. Nucleic Acids Res 37:D141-5. 153 IIIIIII STATE NI ‘6 VERSITY LIBRARIES 11111311111er Mlflljllflllflljlmblflll 1293 0