BACTERIA, MICROBIAL COMMUNITIES AND ENGINEERING: STUDIES ON THE MICROBIAL ECOLOGY OF SELECTED ENGINEERED SYSTEMS By Fan Yang A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Environmental Engineering – Doctor of Philosophy 2013 ABSTRACT BACTERIA, MICROBIAL COMMUNITIES AND ENGINEERING: STUDIES ON THE MICROBIAL ECOLOGY OF SELECTED ENGINEERED SYSTEMS By Fan Yang Environmental contaminants, such as soluble metal ions and agricultural wastes pose great risks for both human health and ecosystems. To reduce these risks, environmental engineers have developed remediation approaches that take the advantage of microbial communities and populations. Understanding these microbial resources is instrumental to manage and apply them in various engineered systems. In this dissertation, I study microbial communities and populations from three different approaches and demonstrate how basic microbial information can assist us in optimizing engineered systems. The first part of my dissertation focuses on understanding the genomic advantages of Ralstonia pickettii strains, which allows them to adapt to high copper environments. We have previously shown that these two strains were able to sequester a large amount of copper. Hence, these two bacterial strains have a great potential in for application to industrial wastewater treatment. Understanding the genomic evolution and adaptation behind the copper binding phenomenon could unveil the industrial potential of these bacterial strains. The second part of this dissertation focuses on understanding the role of anaerobic bacterial populations and communities in uranium immobilization. A large amount of research has been conducted on identifying the bacterial communities involved in in situ uranium immobilization. However, the extant of soil microbial diversity made it difficult to identify the most important specific populations. We employed enrichment culture methods to in- crease the abundance of potential important bacterial populations and to link the community functions. Finally, I present a study on microbial communities in methane producing agricultural waste co-bioreactors. Methane production has been reported as a highly cooperative reaction between bacteria and archaea. Linking bacterial populations to specific functions would help optimize agricultural waste degradation as well as alternative energy production. I chose these three topics to emphasize the importance of microbial populations in engineered systems. By understanding the roles of individual bacteria populations as well as their interactions with each other in a community, I hope to manage and utilize these microbial resources to improve our living environment. Copyright by FAN YANG 2013 This thesis is dedicated to my parents, Hanying Yang and Ning Shen. v ACKNOWLEDGEMENTS This dissertation marks the end of my journey as a Ph.D. student. Throughout the past five years, there was excitement, frustration, as well as a tremendous amount of support from people around me. I would like to begin the acknowledgement by thanking Dr. Terence L. Marsh, who graciously took me in as an undergraduate research assistant, from where I grew into loving microbial ecology. From 2004 to 2012, I have learned the importance of bacteria and their applicabilities in sustainably managing our environments. I thank Dr. Thomas C. Voice, who took the responsibility of being my major advisor in the department of Civil and Environmental Engineering. It was a drastic change switching from microbiology to engineering. I am grateful for Dr. Voice’s guidance on how to think and solve problems like an engineer, which sets me up for applied science. I would like to thank my research committee members, Dr. Alison Cupples and Dr. David T. Long, who have always been supportive and taught me many valuable lessons through the past five years. My thank goes to countless supportive friends. I’m especially thankful for Dr. Adina Howe and Mike Howe’s hospitality, great advice and the tremendous amount of support for the past one year and a half, which was definitely special and eventful. I would like to thank Dr. Juliana Bordowitz and Dr. Zarraz May-ping Lee for their great friendship, without whom, the transition to the Ph.D. program would not be easy. My thank goes to the entire Ribosomal Database Project members, who have been extremely helpful in preparing me for the big bioinformatic world. I truly appreciate the great relationships I have with the past and current Tiedje group members. My special thank goes to the entire vi ROME lab, especially, Chirstina Harzman, Natasha Issacs-Cosgrove, Joe Duris, and Lisa Reynolds Fogarty, without whom, I would not think graduate school is fun. I also would like to acknowledge all of the hard work Katherine Ivens, Rachael Rossi, and Brian Lovett have done as undergraduate research assistants. My thank also goes to all of the secretaries in both the Department of Microbiology and Molecular Genetics and the Department of Civil and Environmental Engineering, especially Lori Larner, without whom, I would not be able to make it through the first year. Finally, I would like to thank my parents, Hanying Yang and Ning Shen, for the wonderful opportunity they gave me to study abroad at Michigan State University. It has been a long journey and I am thankful that I was not doing this alone. vii TABLE OF CONTENTS LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii CHAPTER 1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . 1.1 Bioremediaton of Metals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1.1 Biosequestration of Metals . . . . . . . . . . . . . . . . . . . . . . . . 1.1.1.1 Metal Contamination at Torch Lake . . . . . . . . . . . . . 1.1.1.2 Adaptation of Raltonisa pickettii in High Copper . . . . . . 1.1.2 Bioimmobilization of Metals . . . . . . . . . . . . . . . . . . . . . . . 1.1.2.1 Metal Contamination at Oak Ridge Reservation . . . . . . . 1.1.2.2 Bacterial Community Profiling . . . . . . . . . . . . . . . . 1.1.2.3 Reconstruct Uranium Immobilizing Bacterial Communities . 1.2 Bioremediation of Agriculture Wastes . . . . . . . . . . . . . . . . . . . . . . 1.2.1 The Role of Class Clostridia in Methane Producing Anaerobic Cobioreactors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3 Managing Complex Microbial Communities for Sustainable Environment . . 1 1 3 6 6 11 12 14 14 16 19 20 CHAPTER 2 CO-EVOLUTION OF RALSTONIA PICKETTII STRAIN 12D AND 12J: GENOME REARRANGEMENTS AND THEIR INFLUENCES ON THE EVOLUTION OF METAL RESISTANCE . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 2.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 2.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 2.3 Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 2.3.1 Genome Retrieval and Gene Annotations . . . . . . . . . . . . . . . . 26 2.3.2 Genome Comparisons . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 2.3.3 Identification of the Metal Resistant Genes . . . . . . . . . . . . . . . 27 2.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 2.4.1 Genome Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 2.4.2 Co-evolution of R. pickettii 12D and 12J . . . . . . . . . . . . . . . . 39 2.4.2.1 16S rRNA Gene Sequence Variations . . . . . . . . . . . . . 39 2.4.2.2 dif Site Sequence Variations . . . . . . . . . . . . . . . . . . 40 2.4.2.3 Filamentous Phage Elements in R. pickettii 12D and 12J . . 42 2.4.3 Genome Syntenies of R. pickettii 12D and 12J . . . . . . . . . . . . . 43 2.4.4 The Acquisition and Duplication of Resistance Determinants in R. pickettii 12D and 12J . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 2.4.4.1 Metal Resistant Genes in R. pickettii 12D and 12J . . . . . 49 2.4.4.2 The Phylogenetic Lineage of czcA Copies in R. pickettii 12D and 12J . . . . . . . . . . . . . . . . . . . . . . . . . . 53 viii 2.4.4.3 2.5 2.6 The Phylogenetic 12D and 12J . . Discussion . . . . . . . . . . . . . Conclusion . . . . . . . . . . . . . Lineage . . . . . . . . . . . . . . . of copA . . . . . . . . . . . . . . . Copies . . . . . . . . . . . . in R. pickettii . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 60 65 CHAPTER 3 RESOURCE-INDUCED BACTERIAL STRUCTURE SHIFTS AND THEIR INFLUENCE ON COMMUNITY FUNCTIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 3.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 3.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68 3.3 Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 3.3.1 Anaerobic Media and Stocks . . . . . . . . . . . . . . . . . . . . . . . 71 3.3.2 Bacterial Community Enrichments . . . . . . . . . . . . . . . . . . . 71 3.3.3 Quantitative Analyses . . . . . . . . . . . . . . . . . . . . . . . . . . 72 3.3.4 DNA Extraction and Community Profiling . . . . . . . . . . . . . . . 73 3.3.5 Community Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 3.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 3.4.1 Monitoring the Enrichment Cultures . . . . . . . . . . . . . . . . . . 75 3.4.2 Overview of the Community Dynamics . . . . . . . . . . . . . . . . . 76 3.4.3 Shifts in the Enrichment Bacterial Community Structures . . . . . . 77 3.4.4 Shared Genera . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 3.4.5 Functional Gene Profiles . . . . . . . . . . . . . . . . . . . . . . . . . 83 3.4.6 Non-metric multidimensional scaling (NMDS) analysis . . . . . . . . 86 3.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 3.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 CHAPTER 4 THE ROLE OF FIRMICUTES IN URANIUM IMMOBILIZATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 4.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 4.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97 4.3 Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 4.3.1 Uranium Immobilization by FW107 and FW102-2 Enrichment Cultures 98 4.3.2 Quantitative Analyses . . . . . . . . . . . . . . . . . . . . . . . . . . 98 4.3.3 DNA Extraction and Sequecing Preparation . . . . . . . . . . . . . . 99 4.3.4 Community Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . 100 4.3.5 Isolation of Potential Uranium Immobilizing Bacteria . . . . . . . . . 101 4.3.6 Growth Curves of Isolates . . . . . . . . . . . . . . . . . . . . . . . . 101 4.3.7 Isolate Identifications . . . . . . . . . . . . . . . . . . . . . . . . . . . 102 4.3.8 Cell Suspension Assay . . . . . . . . . . . . . . . . . . . . . . . . . . 103 4.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103 4.4.1 Soluble U6+ Immobilization by FW107 and FW102-2 Enrichments . 103 4.4.2 Bacterial Community Compositions of FW107 and FW102-2 Enrichments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104 4.4.3 Potential U6+ Immobilizers . . . . . . . . . . . . . . . . . . . . . . . 111 4.4.4 U6+ Immobilization By Isolates . . . . . . . . . . . . . . . . . . . . . 117 ix 4.5 4.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 CHAPTER 5 PHYLOGENETIC COMPOSITION OF BACTERIA IN ANAEROBIC CO-DIGESTION OF DAIRY MANURE AND CORN STOVER . . . . . . . . . . . . . . . . . . . . . . 5.1 Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3 Materials and Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.1 Anaerobic Bioreactors . . . . . . . . . . . . . . . . . . . . . . . . . . 5.3.2 DNA Extraction and Targeted Gene Amplifications . . . . . . . . . . 5.3.3 Preparation for Pyrosequencing and Construction of Clone Libraries . 5.3.4 Sequence and Community Analyses . . . . . . . . . . . . . . . . . . . 5.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.1 The Influence of HRT and Corn Stover on System Performance . . . 5.4.2 Phylogenetic Comparisons of Anaerobic Bioreactor Bacterial Communities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.3 Bacteroidetes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.4.4 Hydrogen Producing Population Profiles . . . . . . . . . . . . . . . . 5.4.5 Non-metric multidimensional scaling (NMDS) analysis . . . . . . . . 5.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 122 123 124 124 125 126 127 128 128 130 140 143 148 150 159 APPENDICES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161 APPENDIX A COMPARATIVE GENOMICS OF RALSTONIA PICKETTII 12D AND 12J . . . . . . . . . . . . . . . . . . . . . A.1 Metabolic Pathways of R. pickettii 12D and 12J . . . . . . . . . . . . . . . . A.2 Clustering of R. pickettii 12D and 12J and other 23 Proteobacteria . . . . . A.3 Genetic Elements in R. pickettii 12D and 12J . . . . . . . . . . . . . . . . . A.4 Variation of R. pickettii 12D and 12J dif Sites . . . . . . . . . . . . . . . . . A.5 Genome Syntenies of R. pickettii 12D and 12J . . . . . . . . . . . . . . . . . 162 162 163 164 165 167 APPENDIX B ANAEROBIC ENRICHMENTS . . . . . . . . . . . . . . 169 B.1 Media Preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169 B.2 Functional Gene Profiles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171 APPENDIX C URANIUM IMMOBILIZATION BY C.1 Uranium-azide Assay . . . . . . . . . . . . . . . . . . . C.2 5X Gitschier Buffer . . . . . . . . . . . . . . . . . . . . C.3 Growth Curves of Isolates in Minimal Media . . . . . . FIRMICUTES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172 172 173 173 APPENDIX D BACTERIAL COMMUNITIES IN ANAEROBIC COBIOREACTORS . . . . . . . . . . . . . . . . . . . . . . . . 176 D.1 Deep Sequencing of Bacterial Communities in Anaerobic Co-bioreactors . . . 176 D.2 Anaerobic Co-bioreactor Bacterial Community Compositions at Phylum Level 178 x BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182 xi LIST OF TABLES Table 2.1 Table 2.2 Table 3.1 Gene summaries of R. solanacearum GMI1000, C. metallidurans CH34, R. pickettii 12D and 12J. . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 The abundance of Co-Zn-Cd, Cu, and Hg resistant genes in R. pickettii 12D and 12J, R. solanacearum GMI1000, and C. metallidurans CH34. . . 51 Ecological indices of iFRC enrichment bacterial communities. O represents original soil sediment samples. Fe, NO3, and U supplemented enrichment cultures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78 Table A.1 Side-by-side genome comparison of R. pickettii 12D and 12J genetic elements by genomic scaffolds. . . . . . . . . . . . . . . . . . . . . . . . . . 164 Table B.1 The composition of the modified FWM. . . . . . . . . . . . . . . . . . . . 169 Table B.2 The composition of the mineral mix (100X). . . . . . . . . . . . . . . . . 169 Table B.3 The composition of the vitamin mix (1000X). . . . . . . . . . . . . . . . 170 Table C.1 Ingredients for 200 ml of 5X Gitschier Buffer. . . . . . . . . . . . . . . . 173 Table D.1 Measurements of sampling size and estimation of ecological indices of bacterial community in six anaerobic bioreactors. Calculations were based on the complete linkage clustering of chimera removed and trimmed bacterial 16S rRNA gene sequence alignments at 97% similarity. N represents the total number of non-chimera sequences that passed quality trimming. Shannon index (H’) and Simpson’s index (D) were expressed in eH and 1-D, respectively to emphasize the differences. C represents the sampling coverage via gene targeted pyrosequencing. . . . . . . . . . . 176 xii LIST OF FIGURES Figure 2.1 Figure 2.2 Figure 2.3 Figure 2.4 Figure 2.5 Figure 2.6 Figure 2.7 Figure 2.8 R. pickettii 12D genome scaffolds. The distribution of putative horizontally transferred genes (HGT genes) identified by JGI-IMG are marked as black dashes in rings 4-10. The scaffold sizes are in units of Mb. For interpretation of the references to color in this and all other figures, the reader is referred to the electronic version of this dissertation. . . . . . . 31 R. pickettii 12J genome scaffolds. The distribution of putative horizontally transferred genes (HGT genes) identified by JGI-IMG are marked as black dashes in rings 4-9. The scaffold sizes are in units of Mb. . . . 35 Minimal evolution phylogenetic tree of β-Proteobacteria dif sequences. The numbers on branches are bootstrap values. The scale at the left hand lower corner represents the distance. . . . . . . . . . . . . . . . . . 41 Filamentous phage p12J gene compositions and their homologs on R. pickettii 12D and 12J chromosomes. Fragments of the same fragments represent homologous genes. Genes colored yellow with a red border represent genes that share no homology to p12J Orf7 but are orthologs to each other. Grey colored fragments represent other phage related genes (not p12J related). White colored blocks are other non-phage related genes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 Genome syntenies between R. pickettii 12D and 12J scaffolds. Panel A, 12D_C1, 12J_C1 and 12D_P2; Panel B, 12D_C2 and 12J_C2; Panel C, 12D_P3 and 12J_P1. Arrows indicate location of ori C on the chromosomes. Blocks with the same color represent the regions that are similar (connected by lines). The amount of color indicates the percentage of identity. White areas within the blocks indicate absence of a homologous sequence. . . . . . . . . . . . . . . . . . . . . . . . . . 44 BLAST results mapped onto the 1.72 Mb-1.94 Mb region of R. pickettii 12J_C1. The top scale indicates the length of the query region. Similar genes are shown underneath the query sequence. The short thin black bars indicate that the next stretch of sequence is on the opposite strand of DNA. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47 Phylogenetic position of R. pickettii 12D and 12J czcA genes. The numbers on the branches represent bootstrap values. . . . . . . . . . . . 54 Phylogenetic position of R. pickettii 12D and 12J copA genes. The numbers on the branches represent bootstrap values. . . . . . . . . . . . 59 xiii Figure 3.1 Figure 3.2 Figure 3.3 Figure 3.4 Figure 3.5 Figure 3.6 Figure 3.7 Figure 3.8 The concentration of Fe2+ (panels A and B), lactate and acetate (panels C and D) in each ferric citrate enrichment cultures. The first generation enrichments are presented in panel A and C and the second generation enrichments are presented in panel B and D. Error bars represent the differences between two replicates. . . . . . . . . . . . . . . . . . . . . . 79 The top ten (summed across all samples) most abundant genera identified via RDP multiclassifier. Sample replicates were labeled as "_1" and "_2". Original soil sediment samples were denoted with "_O." . . 80 The shared genera among FW107 samples. The color of sample font is the same as the cognate ellipse. . . . . . . . . . . . . . . . . . . . . . . . 81 The shared genera among FW102-2 samples. The color of sample font is the same as the cognate ellipse. . . . . . . . . . . . . . . . . . . . . . 82 The shared genera among FW102-3 samples. The color of sample font is the same as the cognate ellipse. . . . . . . . . . . . . . . . . . . . . . 83 The functional gene profiles of the original soil sediment and enrichment samples. The functional gene signals of all samples (columns) were standardized to the same amount. Each functional gene (row) was then scaled among all 21 samples to show the differences in abundance. Green indicates high abundance and red indicates low abundance. . . . . . . . 85 NMDS analysis of NO− enrichments (grey), Fe3+ enrichments (pink), 3 U6+ enrichments (yellow), and original samples (green). Each ellipse was drawn around each group centroid based on a group distance standard error at 95% confidence level. The changes in genera abundances that drive comparative community differences are indicated with red arrows. The arrows point to the areas of high abundance. The red arrows indicate the abundance of genera that changed significantly (P ≤ 0.001). The specific genera are shown in red according to the group labeling. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 NMDS analysis of NO− enrichments (grey), Fe3+ enrichments (pink), 3 U6+ enrichments (yellow), and original samples (green). Each ellipse was drawn around each group centroid based on group distance standard error at 95% confidence level. The changes in functional gene (with the specific microorganism from which the probes were designed) abundance were correlated to the shifts observed in community structures. The red arrows indicate the abundance of functional genes that changed significantly (P ≤ 0.001). The specific genera are shown in red according to the group labeling. The arrows point to the areas of high abundance. 89 xiv NMDS analysis of NO− enrichments (grey), Fe3+ enrichments (pink), 3 and U6+ enrichments (yellow). Each ellipse was drawn around each group centroid based on group distance standard error at 95% confidence level. The changes in genera abundance were correlated to the shifts observed in community structures. The colored arrows indicate the abundance of genera that changed significantly (red: P ≤ 0.001, blue: P ≤ 0.05, and green: P ≤ 0.01). The genera names are shown in the same color according to the group labeling. The arrows point to the areas of high abundance. . . . . . . . . . . . . . . . . . . . . . . . . . . 90 Figure 3.10 NMDS analysis of NO− enrichments (grey), Fe3+ enrichments (pink), 3 and U6+ enrichments (yellow). Each ellipse was drawn around each group centroid based on group distance standard error at 95% confidence level. The changes in functional gene abundance were correlated to the shifts observed in community structures. The red arrows indicate the abundance of functional genes that changed significantly (P ≤ 0.001). The functional genes and the specific microorganisms from which the probes were designed are shown in the red according to the group labeling on the side. The arrows point to the areas of high abundance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91 Figure 3.9 Figure 4.1 The removal of soluble U6+ in FW107 and FW102-2 enrichments. The high soluble U(VI) concentration in the control culture (No Cells) was due to the excess amount of uranyl acetate carried over from the syringe needle during inoculation. . . . . . . . . . . . . . . . . . . . . . . . . . . 106 Figure 4.2 Bacterial community shifts with the increasing of incubation time in FW107 U-PBAGW microcosm. . . . . . . . . . . . . . . . . . . . . . . . 107 Figure 4.3 Bacterial community shifts with the increase of incubation time in FW1022 U-PBAGW microcosm. . . . . . . . . . . . . . . . . . . . . . . . . . . 108 Figure 4.4 Distribution of [FeFe]-hydrogenase containing bacteria in FW107 and FW102-2 grown in 1/2X R2B media (U 0 days) and in uranyl acetate enrichments (U 64 days). The bacterial populations represented by the small fractions of each pie chart are listed on the right hand side with color blocks corresponding to the colors found in each pie chart. The distribution of genera Clostridium (cyan color) and Desulfotomaculum (purple color) are shown as the outer ring of each pie chart. . . . . . . . 109 Figure 4.5 The phylogenetic distribution of the anaerobic isolates based on their 16S rRNA gene sequences. The numbers on the branches represent the bootstrap values (1000 permutations). . . . . . . . . . . . . . . . . . . . 112 Figure 4.6 The growth curves of FRC isolates in 1/2X R2B under anaerobic condition (H2 :CO2 :N2 3:10:87). See next page for legends. . . . . . . . . . 114 xv Figure 4.7 REP-BOX-(GTG)5 -PCR profiles of potential uranium immobilizing bacteria. Branches labeled Dhaf represent control samples. The rest represents enrichment isolates. . . . . . . . . . . . . . . . . . . . . . . . . . . 116 Figure 4.8 Soluble U6+ removal by Firmicutes isolates (from FW107) and FW107 uranium enriched community after 70 hours (T70) of incubation at room temperature in an anaerobic chamber (H2 :CO2 :N2 3:10:87). Sample "neg" represents a control culture without the addition of any bacterial cells. The error bars represent measurements of two replicates. . . . . . 118 Figure 5.1 The total solids (TS) reduction efficiency (diagonal and grid lined columns) and the production rate of biogas (grey and white columns) in the DM and DM+CS bioreactors at four HRTs: a HRT of 30 days (HRT30d), a HRT of 40 days (HRT40d) and a HRT of 50 days (HRT50d). Solid diamonds and solid triangles represent the amount of methane produced in the DM and DM+CS bioreactors, respectively. . . . . . . . . . . . . . 129 Figure 5.2 The abundance and diversity (OTU at 97% sequence similarity) of phyla Bacteroidetes (panel A1), class Clostridia (panel B1), and Escherichia / Shigella (panel C1) in each bioreactor bacterial community. The distribution of OTUs among samples are presented as Venn diagrams in panel A2 (Bacteroidetes), B2 (Clostridia) and C2 (Escherichia/Shigella). The colors of the bioreactor/HRT font corresponds to an area bounded by the same color. Labels HRT30d, HRT40d and HRT50d represent bioreactors with a HRT of 30, 40, and 50 days, respectively. . . . . . . . . . . 131 Figure 5.3 The abundance of phyla Bacteroidetes (vertical stripes) in each bioreactor bacterial communities. The distribution of classified and unclassified Bacteroidetes within phylum Bacteroidetes were calculated and displayed in left panel (white and diagonal striped bars) and right panel (black and gridded bars) respectively. Labels HRT30d, HRT40d and HRT50d represent bioreactors with a HRT of 30, 40, and 50 days, respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142 Figure 5.4 The comparisons of Bacteroidetes clusters (97% sequence similarity) between the DM+CS bioreactor with a HRT of 40 days and other bioreactor samples. Labels HRT30d, HRT40d and HRT50d represent bioreactors with a HRT of 30, 40, and 50 days, respectively. Symbol ∗ denote peaks that are clusters of unclassified Bacteroidetes that are highly abundant in all DM+CS bioreators (#11 and #14), in DM+CS bioreactors at long HRTs (#43), and in the DM+CS bioreactor with a HRT of 40 days (#69). . . . . . . . . . . . . . . . . . . . . . . . . . . . 144 xvi Figure 5.5 The shifts of abundance of the three most abundant [FeFe]-hydrogenases containing bacterial groups in anaerobic bioreactors fed with DM and DM+CS. Labels HRT30d, HRT40d and HRT50d represent bioreactors with a HRT of 30, 40, and 50 days, respectively. . . . . . . . . . . . . . . 146 Figure 5.6 The abundance of [FeFe]-hydrogenase containing bacteria in anaerobic bioreactor samples compared to the abundance pattern of unclassified Bacteroidetes (background) determined by 16S rRNA gene sequencing. The distribution of clusters #11, #14, #43 and #69 were also plotted as part of unclassified Bacteroidetes within each sample. Labels HRT30d, HRT40d and HRT50d represent bioreactors with a HRT of 30, 40, and 50 days, respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147 Figure 5.7 NMDS analysis of the bacterial community in anaerobic bioreactors (based on 16S rRNA gene sequences at 97% similarity cutoff). The green arrows indicate the increase of biogas production (Biogas), HRT (HRT) and dairy manure degradation rate (Degradation). The red and blue arrows indicate bacterial genera that shifted significantly, where P≤0.001 and P≤0.01, respectively. The black, yellow and purple arrows represent the [FeFe]-hydrogenase phylogenetic affiliations that shifted significantly (P≤0.001, P≤0.05, and P≤0.01, respectively). All arrows point towards the abundant range of the gradients. Labels HRT30d, HRT40d and HRT50d represent bioreactors with a HRT of 30, 40, and 50 days, respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149 Figure 5.8 Diagram (adapted from Schink et al. [140]) of microbial corporation during methane generation in DM and DM+CS bioreactors operated at HRT 40 days. The thickness of the arrow represent the relative abundance of microbial populations possibly associated with the methanogensis step. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151 Figure A.1 Distance tree of R. pickettii 12D and 12J in relationship to 23 other Proteobacteria based on their 16S rRNA gene sequences. The bar in the bottom left corner represents the distance scale. . . . . . . . . . . . . 163 Figure A.2 The sequence variations of dif sites of R. pickettii 12D and 12J. The consensus reveals the conserved the regions. . . . . . . . . . . . . . . . . 165 Figure A.3 Genome syntenies of R. pickettii 12D and 12J revealed by a DNA sequence dotplot. Blue dots represent matching genes on the same direction strands. Red dots represent matching genes on the opposite direction strands. The scaffold nucleotide coordinates are ranked from left to right on the x-axis and bottom to top on the y-axis. . . . . . . . 168 xvii Figure B.1 Functional gene profiles of original soil communities and the enriched communities (replicate 1). Symbol "O" indicates original communities and "Fe", "NO3" and "U" indicate supplemented electron acceptors in enrichments. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171 Figure C.1 Standard curve of the uranium-azide assay. The error bars represent data from two replicates. . . . . . . . . . . . . . . . . . . . . . . . . . . 172 Figure C.2 The growth curves of FRC isolates (first replicate) in minimal media supplemented with sodium pyruvate alone (blue) and minimal media supplemented with sodium pyruvate and uranyl acetate (red) under anaerobic condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174 Figure C.3 The growth curves of FRC isolates (second replicate) in minimal media supplemented with sodium pyruvate alone (blue) and minimal media supplemented with sodium pyruvate and uranyl acetate (red) under anaerobic condition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175 Figure D.1 Estimation of community coverage by rarefaction analysis. Rarefaction curves of anaerobic bioreactor 16S rRNA gene sequences were constructed based on the estimation of OTUs and ChaoI indices at 97% similarity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177 Figure D.2 The distribution of the top three most abundant phyla (wide columns) in bioreactor samples. The narrow columns represent the abundance of genera Escherichia/Shigella (checker box), group unclassified Bacteroidetes (diagonal stripes) and class Clostridia (blacks) in each community. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180 xviii CHAPTER 1 INTRODUCTION Environmental engineering technologies frequently take advantage of microbial communities and populations. My dissertation is focused on the investigations of microbial communities involved with bioremediation of metal contaminated sites and the anaerobic digestion of farm waste. In both cases, increasing our knowledge base regarding the microbial communities that are instrumental in each of these activities is critical to efficient remediation and longterm management. In this background section I will first discuss bioremediation of metals followed by anaerobic digestion of farm wastes. 1.1 Bioremediaton of Metals Metals are naturally occurring elements that are important components of the earth’s crust and have played significant roles in the development of human civilizations. The discovery and utilization of metals signify the development of different culture eras associated with the history of human civilizations, the Bronze Age, the Iron Age, the Steel Age, and finally the Nuclear Age. Each era is represented by the mining and smelting of a particular metal ore, e.g., Cu, Fe, iron alloys, and U. As the primary method to expose and collect metals in the earth crust, mining became one of the most important industries in human history [55]. Although the development of mining was instrumental in the evolution of human civilizations, mining does massive damage to the environment. Commonly known issues such as land erosion and formation of sinkholes may accompany mining processes. The toxic 1 chemicals (e.g., sulfuric acid, cyanide, alkyl xanthate salt) used during mining to enrich specific metals (e.g., Cu, Au) can contaminate surrounding soil, surface and subsurface water bodies. The metal-containing tailing solutions along with naturally occurring oxidation on exposed metals, can mobilize metals from the earth’s crust and cause further contamination of the surrounding environment via river and aquifers. Although the toxicity of metals varies, ingestion of soluble metals, especially heavy metals, can be life threatening. Hence, the immense impact of mining on biodiversity and ecosystems cannot be ignored. The mobility and toxicity of metals in the environment are controlled by local acidbase and redox reactions. Shifts in pH can largely influence metal mobilities. Take the most abundant transition metal, iron, as an example. Oxidized iron (Fe3+ ) precipitates as Fe(OH)3 or other Fe3+ minerals at near-neutral pH but it becomes more soluble when pH decreases. Therefore, Fe3+ is mobile in acidic environments. Environmental redox reactions control the formation and dissolution of minerals. For example, iron can be oxidized to Fe3+ by exposure to oxygen. In the presence of sulfide or reduced humics, Fe3+ can be readily reduced to Fe2+ . At neutral pH, Fe2+ is much more soluble than Fe3+ . Hence, in the form of Fe2+ , iron can be readily absorbed by living organisms. For this reason, Fe2+ is more toxic than Fe3+ at near-neutral pH [8]. Bacteria contribute substantially to metal mobility and toxicity. The growth of bacteria can change local pH. Through metabolic processes, such as denitrification, bacteria can increase the pH of local environments. When this pH shift occurs in an acidic habitat, the precipitation of metal ions would occur due to the increase of pH [187]. Also through metabolic processes, such as oxidation of pyrite by Acidithiobacillus species, bacteria can decrease the pH of the local environment. This decline of pH may result in metal leaching from 2 rocks increasing the mobility and toxicity of heavy metals [54]. Bacteria can also physically reduce the concentration of mobile metal ions via biosequestration. Due to their large surface to volume ratios and the composition of the cell envelopes, bacteria (alive or dead) can naturally absorb mobile metal ions [178]. Studies have also shown that some bacteria can sequester a large amount of metals in the outer envelope of the cell and thus remove metal ions from aqueous solution [90, 136, 191]. Finally, bacteria play an essential role in environmental redox processes. Some bacteria can derive energy by transferring electrons from organic carbon or inorganic compounds to electron acceptors, including transition metals, metalloids and actinides. Studies have demonstrated that anaerobic bacteria, such as Geobacter metallireducens, Shewanella oneidensis, and Desulfuromonas acetoxidans, can dissimilatorily reduce metal ions (e.g., from Fe3+ to Fe2+ ) under anoxic conditions [65]. These indirect and direct interactions between bacteria and metals have been observed across a wide spectrum of heavy metals, including Mn, Cu, Cr, Co, Hg, and U. The unique attributes of bacterial metal sequestration and immobilization have been cultivated by environmental engineers for remediation of metal contamination. 1.1.1 Biosequestration of Metals Biosequestration of metals herein refers to metal adsorption and accumulation by either live or treated bacteria. Biosorption usually refers to the passive binding of metals to biological materials. The sorptivity of metals by the cell envelope is independent of active metabolism but directly related to sorbent surface area. The composition of the bacterial cell envelop also contributes greatly to metal specificity and binding affinity [139]. Due to their small sizes, bacteria naturally have high surface to volume ratios. Escherichia coli cells achieve 3 a surface area of over 21 m2 /g dry cell weight [108, 158]. With appropriate treatments, including physical crushing, treating and caustic treatment, bacterial biomass can be an efficient adsorbent [177]. Indeed, treated bacterial cells are preferred over live bacterial cells because of convenience (e.g., little maintenance cost and easy to transport). The term bioaccumulation refers specifically to metabolism-dependent metal uptake. While some metals at trace levels are required to maintain life (e.g., Se, Cu, Fe), toxic effects can be observed in bacteria when large amounts of metals are present. Research has shown that 500 µM of Cu2+ (0.125 g/L) is toxic to E. coli. The ability to grow at elevated levels of copper are largely due to resistance mechanisms including bioaccumulation, membrane bound e-flux pumps and metal transporters [147]. The active metal accumulation process is primarily driven by efflux pumps that relocate metals from the interior to the exterior of cells. Active bioaccumulation is likely to be slower than biosorption and subject to growth condition, such as narrow ranges of pH, temperature and salt content. However, bioaccumulation has a great potential in metal remediation because living metal sequestering bacteria have high specificity in metal binding and can regenerate biomass via growth [95]. Biosequestration can be practically applied in metal removal/recovery from industrial waste streams [177]. Wastewater produced from various industries, including mining, leather working, energy production, and electroplating, contain different heavy metals. Discharging this waste into the environment poses serious risk (e.g., toxic metals like Hg and radioactive metals like U) as well as being a waste of resources (e.g., precious metals like Au). Conventional methods like chemical precipitation and coagulation are able to reduce metal concentrations in waste streams only down to the mg/L level. Ion-exchange techniques can remove and recover metals from wastewater to µg/L levels, however, the cost of exchange 4 resins is high and the resins are prone to poisoning by organics and solids in the wastewater. Membrane-based processes (e.g., reverse osmosis) are effective at obtaining pure effluents but are extremely expensive for at large flow rates and they are prone to failure as a results of membrane fouling [81, 177]. Compared to conventional metal removal/recovery methods, biosequestration offers several potential advantages and may offer the potential of a costeffective alternative method. Bacterial biomass can be relatively inexpensive and can serve as a sink to all metals except the alkali cations (e.g., Na+ and K+ ). Biosorbents developed from bacteria biomass have a broad range of metal binding affinities that could be exploited in selective metal removal/recovery. For example, Ogi et al. showed that Shewanella algae can be used in rare metal indium recovery from solutions at low concentrations. This study revealed that at pH 3.8, S. algae cells were able to remove 100% of the soluble indium in solution at concentrations up to 0.2 mol/m3 (229 mg/l). Indium concentrates could be prepared from indium bound to cells by burning at approximately at 800◦ C [115]. Several species of the genus Bacillus have shown to be effective at removing Cu and Pb from aqueous solutions [122, 165, 133]. As reported by Tsuruta, Arthrobacter sp., Bacillus sp. and Lactobacillus sp. isolated from uranium deposits removed 10.9-98.3% of U in solution. Some of the Lactobacillus isolates were able to remove uranium from seawater as well. A selective accumulation study conducted using one of the Arthrobacter isolates revealed that UO2+ is 2 preferentially accumulated over six other metal ions including Cu2+ , Mn2+ , Co2+ , Zn2+ , Cd2+ and Ni2+ [164]. Although biosequestration is still under development and has not yet been applied to real wastewater treatment problems, these initial studies suggest that biosequestration could be a viable process. 5 1.1.1.1 Metal Contamination at Torch Lake Torch Lake, located in Houghton county, MI, has been on the National Priorities List since 1986. The area formerly known as the Copper Country, was well-known for its metallic copper ore that was extensively mined from 1890’s to 1969. During the mining period, the 2,700 acre Torch Lake served as a waste disposal site for milling, smelting and leaching chemicals and by-products. The U.S. Environmental Protection Agency estimated that approximately 20% of the lake volume was made up of copper mill stamp sands (∼200 million tons). It was also estimated that the contaminated sediments are up to 70 ft deep and contain up to 5,500 ppm of copper. The ecosystem around the lake was further damaged by a cupric ammonium carbonate (a common herbicide) spillage (∼27,000 gallons) in 1972 and the disposal of industrial processing water, which contained copper and ammonium that are 2,400 times and 100 times of the local sewage allowable limits, respectively, in the early 1980’s. With the discovery of extensive ecological degradation and three impairments of beneficial use (fish tumors, degraded benthos and fish consumption advisories), Torch Lake was also enlisted as a Great Lakes Areas of Concern in 1987. The area has been under remediation construction since 2003. While tumorous fish were no longer captured, the recovery of the area is minimal [169]. 1.1.1.2 Adaptation of Raltonisa pickettii in High Copper Eleven Ralstonia pickettii strains were previously isolated from the highly contaminated Torch Lake sediments, where copper concentration ranged from 1,000 to 2,500 ppm. Previous studies revealed that all eleven R. pickettii isolates were Cu resistant, while seven of 6 them were also resistant to Zn and Cd and the rest were resistant to Ni and to low levels of Cd. Konstantinidis et al. [80] observed that the colonies of these R. pickettii isolates obtained a greenish blue color when growing on copper supplemented agar plates. It was hypothesized that these isolates could uptake copper from the medium and deposit it in the outer cell envelope [80]. Scanning electron microscopy (SEM) and transmission electron microscopy (TEM) supported this hypothesis and revealed drastic changes to the cells when the isolates were grown in the presence of copper, including a substantial increase of the amount of extracellular material and the deposition of high-density materials in the outer envelope. Electron diffraction scanning of the copper-grown cells revealed the presence of a large amount of copper, which was not observed in the cells grown in the absence of copper. X-ray absorption near edge structure (XANES) and Extended X-ray absorption fine edge structure (EXAFS) analyses suggest that the cell associated copper is likely to be an organic bound tetrachedral cupric complex and the copper sequestration process is not via (hydr)oxide mineral precipitation [191]. A systematic copper sequestration study revealed that the isolate representatives R. pickettii strain 12D (Cu-Ni-Cd resistant) and strain 12J (Cu-Zn-Cd resistant) are well adapted to the high copper environment and the binding mechanisms were different between the viable and heat-killed cells. The live cells can actively sequester copper up to 27.44 (12D) and 38.19 (12J) mg Cu/g dry weight of cells, which is 1.5 to over 2 times more than what heat-killed 12D and 12J cells adsorb. The study also suggests that distinct binding/transporting mechanisms are involved in copper sequestration by live R. pickettii 12D and 12J cells and that live 12J cells bind significantly more copper than live 12D cells [191]. These attributes suggest that R. pickettii strain 12D and 12J may be exploitable in the remediation of copper. 7 From the copper sequestering studies performed on R. pickettii strain 12D and 12J, we concluded that the active uptake of copper by living R. pickettii cells is likely a display of a detoxification process associated with metal resistance systems. In a 2005 review, Silver et al. noted that bacteria developed metal resistance systems to overcome the toxicity of various toxic metal ions, including Zn2+ , Ni2+ , Cu2+ , Cd2+ , Co2+ , and Hg2+ . The majority of these systems function as energy-dependent (e.g., ATPase and chemiosmotic ion/proton exchange) efflux pumps and some also involve metal-binding proteins. For example, a classic Cd2+ , Zn2+ , and Co2+ resistance system (Czc) is a polypeptide chemiosmotic system. It is comprised of membrane polypeptide chemiosmotic efflux pumps, inner and outer membrane proteins, and periplasmic coupling proteins to form channels between the cytoplasm and outside of cells. To reduce metal toxicity, this system utilizes a proton gradient to transfer excess amounts of Cd2+ , Zn2+ , and Co2+ from inside to outside of the cell [147]. Exploiting the mechanisms of metal resistance will increase our understanding of how bacteria adapt and thrive in high metal environments. This information is also important in evaluating the potential and optimizing the ability of live bacterial cells in removing/recovering metals from industrial wastewater. One can envision a number of mechanisms for the acquisition of metal resistance, including the gradual selection of point mutations to develop a new functionality, the duplication/amplification of existing metal resistance genes to increase resistance level, and the acquisition of additional resistance determinants via horizontal gene transfer. Point mutations occur randomly and take a long period of time to evolve a new function from an old one. Starting from small genetic modifications, point mutations alter a single base at a time. Many point mutations are synonymous and therefore will not change the amino acids, which 8 determine genetic functions. Slowly, point mutations may accumulate and eventually change the amino acids and modify the old function. This process may accelerate under selective pressure in that suitable mutants adapt to the new environmental factors better [92]. In conclusion, point mutation is a prolonged process for evolving a completely new function. Unlike point mutations, gene duplication and amplification (GDA) allows bacteria to gain resistance rapidly [135]. GDA, described by its name, allows a bacterium to evolve by duplicating its own existing resistance determinants. This mechanism confers increased levels of resistance by increasing the number of metal resistance gene copies, thereby increasing the levels of metal resistance gene expression [135]. Bacteria can also obtain new functions from other bacteria (sometimes across species, genus, even domain boundaries), referred to as horizontal gene transfer (HGT), through transformation (foreign gene direct uptake), transduction (foreign gene insertion by bacterial phages), and conjugation (plasmid transfer via cell-cell contact). Both GDA and HGT provide a road to rapid acquisition of resistant genes. Because they both depend on existing resistance determinants, GDA and HGT are distinguishable on a sequenced genome. The presence of plasmids or phage indicates that HGT may have occurred via conjugation or transduction. A stretch of genes on a bacterial genome with different guanosine/cytosine (GC) content compared to the rest of the genome also can be used to detect putative foreign genes inserted. Similarly, the possible occurrence of GDA could be identified by finding multiple copies of the same gene, frequently repeated in tandem, in a genome. It is also possible that HGT of the same gene occurred several times, which would result in multiple copies of the same gene on chromosome. To confirm GDA and differentiate from HGT, one has to rely on phylogenetic analyses of gene copies to reveal the evolutionary lineages of each copy [92, 114, 135]. 9 We have demonstrated that R. pickettii strain 12D and 12J have a great potential in copper removal/recovery from wastewater streams. It has also been shown that these two strains shared similar attributes (e.g., same species, ability to thrive in high copper environments, and active sequestration of copper) but have important differences (e.g., different genome sizes, 12J is resistant to Zn but not Ni, 12D is resistant to Ni but not Zn, 12J grows better in the presence of copper and sequesters more copper than 12D). Hence, these two strains are great model organisms for studying how genetic element variations influence the manifestation of biosequestration by living bacteria. Konstantinidis et al. previously observed that R. pickettii 12D and 12J are resistant to multiple metals [80]. Therefore, it is likely that these two bacterial strains obtained multiple metal resistance systems. From both metal resistance and copper sequestration studies, we observed that R. pickettii 12D and 12J are different both genetically and physiologically. It was observed that these two strains belonged to different genomovar groups and strain 12J appeared to adapt to high copper environment better than strain 12D [80, 191]. Hence, we suspect that R. pickettii 12D and 12J either acquired different metal resistant determinants or developed different regulation mechanisms during evolution. In their 2009 review, Sandegren et al. suggested that GDA is a common adaptive mechanism of bacteria to various antibiotics [135]. Chan et al. observed that GDA and HGT contributed greatly to the development of virulence, antibiotic and metal resistance among Staphylococcus isolates [21]. Thus, as an adaptation mechanism, we expect that GDA and HGT played a substantial role in R. picket 12D and 12J. 10 1.1.2 Bioimmobilization of Metals Besides biosequestration, metals also can be removed from aqueous solution via bioimmobilization. Bioimmobilization of metals herein refers to 1) co-precipitation of metals and metabolic by-products and 2) precipitation of metals via dissimilatory reduction. Coprecipitation of metals and bacterial metabolic by-products describes a chemical process where soluble metal cations combine with sulfides [82, 183] or phosphate compounds to form insoluble metal complexes [130]. Large amounts of sulfide in the subsurface result from the reduction of anthropogenically deposited sulfate (e.g., mining). Anaerobic sulfate-reducingbacteria (SRB) obtain metabolic energy (δG = -190 kcal/mol) by reducing sulfate to sulfide. Sulfide is subsequently complexed with soluble heavy metal ions to form precipitates [121]. Besides sulfide, biogenic phosphate has been observed to be effective in bacterial cell associated metal precipitations. Phosphate is liberated from bacterial storage polyphosphate (e.g., Acinetobacter spp., Citrobacter sp.) via phosphatase activities when the growth condition changes, such as changing from aerobic to anaerobic metabolism [9, 91, 97]. Research has revealed that with the presence of soluble heavy metals, the secreted phosphate forms cell-bound metal-phosphate precipitates [9, 91, 97]. Co-precipitation of heavy metals occurs in accord with the chemistry solubility rules and it does not change the oxidation state of metal ions. Compared to co-precipitation of metals and metabolic by-products, metal precipitation via dissimilatory reduction changes the oxidation state of metal ions directly via redox reactions. Similar to the conversion of sulfate to sulfide, anaerobic bacteria conserve energy by enzymatically transferring electrons to metals ions. While the reduced form of iron (Fe2+ ) and manganese (Mn2+ ) are more bioavailable than their oxidized forms (i.e., Fe3+ 11 and Mn4+ ), other heavy metals are more insoluble in the reduced state (e.g., Se6+ VS. Se0 , Cr6+ VS. Cr3+ , U6+ VS. U4+ ). The dissimilatory reduction and precipitation of heavy metals immobilizes metals in situ without additional growth requirements, such as the presence of sulfate and the need of changing growth conditions. However, the disadvantage of this dissimilatory metal reduction is that the reduced metals can readily be reoxidized to their oxidized state by the exposure to oxygen. In contrast, the co-precipitates of heavy metal and sulfide or phosphate are relatively stable in the presence of oxygen. These metabolism dependent characteristics of bioimmobilization are particularly attractive to in situ subsurface metal remediations for it is sustainable and relatively inexpensive for a large area. 1.1.2.1 Metal Contamination at Oak Ridge Reservation Historical weapons development and improper disposal of inorganic, organic and radioactive wastes have caused extensive areas of subsurface contamination at the Oak Ridge Reservation site (Oak Ridge National Laboratory, TN). For 31 years, the unlined S-3 pond was used to store these toxic wastes. It was later discovered that much of the original contaminants are now associated with soil and subsurface sediments and became secondary contaminant sources encompassing an expanding area. The challenge of remediation at Oak Ridge Integrated Field Research Challenge (iFRC) site is the combination of mixed contaminants, large subsurface area and low pH environment. The wide-spread contamination of both soil and groundwater makes traditional remediation methods, including in situ containment and solidification, mechanical separation, and pyrometallurgical processes, unfeasible [106]. While the low pH environment is not favorable for either chemical remediation or bioimmobilization, increasing pH in situ would cause the precipitation of aluminum and calcium 12 and result in clogging of subsurface pores. In addition, denitrification is likely to occur if growth conditions are optimized because of the high concentration of nitrate (133 mM) in the groundwater. in situ denitrification should be avoided because bacteria will preferably use nitrate rather than U6+ and hence inhibit uranium reduction [187]. Wu et al. [187] explored the approach of combining ex situ denitrification and in situ bioimmobilization by installing a nested recirculation system at the Oak Ridge iFRC S-3 site. The system creates a pH neutral zone by pump and treat, where groundwater is extracted for above-ground pH adjustment, removal of clogging agents and bulk denitrification. To maintain required flow rates, treated water was mixed with tap water prior to reinjection. The biostimulation zone is located in the center of the pH neutral zone, where ethanol was injected as a carbon source and electron donor for bioimmobilization of uranium [187]. The system was able to maintain a stable near-neutral pH for over 500 days. In addition to removing bulk Al3+ , Ca2+ , and nitrate, flushing also reduced soluble uranium from over 100 µM to ∼2-7 µM in 69 days. Soluble uranium concentration in the biostimulation zone persisted at around 5 µM until day 177 due to the presence of residue nitrate but decreased to less than 1 µM by day 268 and to levels below EPA drinking water standard (0.126 µM) after 2 year of stimulation. Meanwhile, the uranium content in the biostimulation zone soil increased accordingly and mixed U6+ and U4+ were detected via x-ray absorption near-edge structure spectroscopy (XANES) [188]. Bacterial community studies of the biostimulated soil revealed iron reducing bacteria (FeRB), sulfate reducing bacteria (SRB) and denitrifiers were grown in the treatment area. Bacteria belonging to groups that were previously identified as uranium reducers (e.g., Geobacter, Clostridium) were abundant in the stimulation zone. Cardenas et al. also observed that the abundance of the SRB group (e.g., Desulfovibiro, 13 Desulfosporosinus) followed the U6+ reduction pattern and suggested that SRB played a prominent role in uranium immobilization at Oak Ridge iFRC site [16, 17]. 1.1.2.2 Bacterial Community Profiling Field studies conducted at Oak Ridge iFRC site proved that in situ bioimmobilization of uranium is a suitable strategy. However, complex field conditions complicate the identification of uranium immobilizing bacteria due to the presence of other metals and minerals. It was also extremely difficult to study the interactions between different bacterial strains in response to the presence of uranium. Hence, we took the approach of enriching uranium immobilizing bacterial consortia from the iFRC and anaerobicly isolating potential uranium immobilizers in the laboratory. 1.1.2.3 Reconstruct Uranium Immobilizing Bacterial Communities In general, microbes exist as communities in the environment, where the diversity of species reflects the diversity of nutrients and the abiotic conditions. In such a diverse community, interactions between species is complex and includes mutualism, commensalism, competition and parasitism. One of the most studied relationships is the syntrophism between bacteria and archaea in methanogenesis. The Methanobacillus omelianskii culture Barker et al.[140] isolated was identified later as a co-culture of S strain and M.o.H strain. The two strains could not be separated because each strain depended on the other under the isolation conditions. Strain S ferments ethanol and strain M.o.H scavenges H2 produced by strain S. The collaborative process generates methane as well as 112 kJ of energy per mole of methane produced [140]. Besides methanogenesis, Kato et al. [74] reported a mixed-culture composed 14 of cellulolytic Clostridium and non-cellulolytic bacteria that can degrade lignocellulose more effectively than isolates. The study showed that it is critical to have non-cellulolytic bacteria scavenge cellulose derivatives and metabolites, maintain the anaerobic condition, and neutralize the pH of the system. It was also revealed that not all non-cellulolytic isolates contributed to cellulose degradation [74]. A further study by the same group showed that the four strains in the assembled cellulolytic community formed a complex network [75]. Besides the collaborative effort on consuming metabolites, a recent study demonstrated that electrons can also be transferred between two different bacterial species. Kato et al. cocultured Geobacter sulfurreducens and Thiobacillus denitrificans. The study showed that G. sulfurreducens transfered electrons from acetate to T. denitrificans, where the electron was further passed onto nitrate to produce ammonia [76]. These studies suggest that microbial interactions and cooperations are important and cannot be ignored in engineered microbial systems. Uranium immobilizing bacterial communities have been studied using two approaches: bottom-up and top-down. A bottom-up approach refers to the strategy of studying the system from individual components, where individual bacterial species are isolated and studied separately. Under strictly controlled conditions, one can determine the uranium immobilization ability of individual bacterial isolates. This approach is effective in understanding cell physiologies of each microorganism. However, the bottom-up approach suffers from a focus on strains that are easily cultivated while neglecting the often more numerous and difficult to cultivate strains. A top-down approach is also used in studying uranium immobilizing communities in order to correlate the total diversity of the community using cultivationindependent techniques with the abiotic conditions (e.g., pH, uranium concentration, etc.). 15 This community information is used to tentatively identify the important bacterial species responsible for uranium immobilization. This approach is extremely powerful because the cultivation-independent approach reveals nearly all populations. However, top-down approaches cannot be used to precisely evaluate population functions due to the large number of variables and in addition, physiology does not always follow phylogeny. Zengler et al. made the point that both top-down and bottom-up approaches should not be treated separately but, rather, should be integrated [193]. It was suggested that bacterial community analyses should be used as a guide for cultivation and in-depth understanding of individual populations. In return, the comprehensive understanding at the cellular level would assist in prediction of complex biological systems. Integrating the two approaches, reconstruction of uranium immobilizing bacterial communities would give insights on population interactions in uranium bioremediation systems. As part of my dissertation, I investigated the uranium immobilizing bacterial communities with a top-down approach. I also cultivated individual bacterial isolates and studied their abilities in uranium immobilization. 1.2 Bioremediation of Agriculture Wastes Agriculture is a significant part of the US economy. The Food and Agriculture Organization of the United Nations (FAO) estimates that approximately $0.2 trillion worth of production (0.67 billion tons) was contributed by agriculture in the US alone in 2010. The top three produced commodities based on value were indigenous cattle meat, dairy products and maize. Maize was also the most abundantly produced based on quantity (at least three times more 16 than other commodities) in the US in 2010. The large level of agricultural production resulted a contaminant amount of waste production. Approximated 6.7 million tons of agriculture wastes were produced in 2010 in the US (estimated based on 1993 facts) [44]. The consequence of improper treatment and disposal of agricultural wastes can be serious. Manures produced by diary and swine farms often contain human pathogens and can be life threatening if the wastes come into contact with water sources. The direct disposal of wastes that are rich in nutrients into the environment can alter the local ecosystems, causing for example, algal blooms. Crop wastes, such as corn stalks are much harder to degrade due to their lignin and cellulose content. Under strict regulations, wastes may be recycled by applying them to fields as fertilizer or they can be used to generate biofuels (e.g., bioethanol). Currently, the most common treatments for agricultural waste are composting and lagoons [166]. Although the typical farm waste management strategy includes waste storage and decomposing, it does lead to concerns regarding air and water quality and greenhouse gas emissions [46]. Methane is a potent greenhouse gas and has a unique role in global climate change and carbon cycling. Petit et al. [125] revealed the high correlation between temperature variation and atmospheric green house gas concentration from the Vostock ice-core record. The past 420 kyr of record also suggested that the present day atmospheric CH4 concentrations are unprecedentedly elevated [125]. As the main component of natural gas, methane is naturally produced by syntrophic microorganisms (i.e., hydrogen producing bacteria and methane generating archaea) during degradation of organic matter via anaerobic methanogenesis [6, 140]. Methanogenesis generates the least amount of energy (∼15% of aerobic degradation of the same organic molecule) and occurs as the last step in microbial anaerobic respiration. 17 Methane production is common in organic-rich sediments [140]. Bastviken et al. [6] estimated that the global freshwater methane emission rate corresponds to at least 0.65 Pg of C (CO2 equivalent, assuming 1 kg of CH4 equals to 25 kg of CO2 over a 100 year period). Besides the naturally emitted methane, anthropogenic methane emission rose substantially along with the development of industry. Stern and Kaufmann [153] estimated that the total amount of global methane emitted from anthropogenic activities increased from 79.3 Tg in 1860 to 371 Tg in 1994 (0.5 Pg and 2.3 Pg CO2 equivalent C, respectively). Over 50% of this was emitted from agricultural activities including livestock production and rice farming [153]. While the terrestrial land surface has been considered an important carbon sink (2.6±1.7 Pg of C), the amount of natural and anthropogenic methane emissions offsets the continental green house gas balance. Thus the impact of global methane emission rate on global climate change should not be overlooked [6]. Besides being a major contributor to global climate change, methane is also an alternative energy source for heat and electricity [46, 78]. With decreasing of fossil fuel discovery rates, the need for renewable energy source increases [15, 66]. Anaerobic digestion of farm wastes not only eliminates some of the environmental concerns of traditional farm waste management methods, such as oder and water contamination, it also can be used to generate and capture methane rich biogas. However, system optimization is critical for anaerobic digestion. Operational parameters, including temperature, HRT, and the carbon to nitrogen ratio of the substrate have been studied extensively [42, 46, 110, 120]. While these studies provide insights on maintaining an efficient anaerobic bioreactor, the microbial communities and species interactions responsible for methane production are not well understood. In this section, I will focus on studying the microbial community aspects of methane producing 18 anaerobic co-bioreactors. 1.2.1 The Role of Class Clostridia in Methane Producing Anaerobic Co-bioreactors Anaerobic methanogenesis is a group effort between bacteria and archaea. Methanogens, a group of archaea, are primary responsible for methane production. However, the process cannot be carried out without the help of bacteria [6, 140]. Schink et al. [140] noted in a recent review that methanogenic degradation of complex organic matter involves a mutual dependency of five groups of microorganisms, primary fermentative bacteria, secondary fermentative bacteria, homoacetogenic bacteria, hydrogen oxidizing methanogens, and acetatecleaving methanogens. The concentration of hydrogen in the system plays a critical role in determining how methane is produced. When the hydrogen partial pressure is less than 10−4 bar, majority of the methane is generated exclusively via the cooporation of primary fermenters and either H2 oxidizing methanogens or acetate cleaving methanogens. Under conditions where hydrogen accumulates in the system (e.g., excess substrate), the secondary fermenters and homo-acetogenic bacteria contribute to methane production by converting hydrogen to acetate. This alternative process shifts the methane producers from hydrogen oxidizing methanogens to acetate-cleaving methanogens [140]. Bacteria class Clostridia categorizes a group of anaerobic bacteria with versatile metabolic pathways. Species belonging to this class are noted for their ability to ferment various substrates, including saccharides, amino acids, and cellulose. Cellulose degradation is particularly relevant as it is a major component of plants. Thus, it is widely abundant, rich in carbon and yet difficult to decompose. Studies have shown that several Clostridium species can decompose and partially hydrolyzed cellulose [36, 74]. Interestingly, many 19 members of Clostridia are also homo-acetogens, such as C. methoxybenzovorans, C. thermoautotrophicum and C. thermoaceticum. These attributes of class Clostridia suggest that this group of bacteria may play a substantial role in anaerobic degradation of agricultural wastes and methane generation. In fact, research studies have repeatedly reported that Clostridia populations were abundant in hydrogen and methane producing anaerobic bioreactors [23, 116, 119, 126, 179]. Bacterial hydrogen production from protons and electrons is catalyzed by hydrogenases. There are two major groups of hydrogenase, [FeFe]-hydrogenase and [NiFe]-hydrogenase, which is composed of two iron atoms and one iron and one nickel atom, respectively. The two groups of hydrogenases are phylogenetically different. While [NiFe]-hydrogenases widely occur in bacteria and archaea, [FeFe]-hydrogenases are found only in anaerobic bacteria and eukaryotes [11, 103]. It is known that [NiFe]-hydrogenases are more abundant in bacteria than [FeFe]-hydrogenase. However, it was suggested that most of [NiFe]-hydrogenases are involved in hydrogen uptake. This is consistent with the observation that [NiFe]-hydrogenases are often found in hydrogen consuming bacteria [10, 176]. In contrast, [FeFe]-hydrogenase is 10-100 times more active than [NiFe]-hydrogenase and energetically prefers to catalyze the reduction of protons to produce hydrogen [47, 189]. Therefore, I hypothesize that class Clostridia is a substantial player in anaerobic co-bioreactors and is the dominant bacterial group in methane producing anaerobic co-bioreactors. 1.3 Managing Complex Microbial Communities for Sustainable Environment Microorganisms reside in various environments, such as soil, water, air, and the intestines of living organisms. Among all microorganisms, bacteria and archaea are highly abundant 20 on earth. Whitman et al. estimated that there are 4-6×1030 bacteria and archaea cells in terrestrial and marine environment [184]. It was estimated that there are at least 4.5×106 different bacterial species globally and 2×104 platonic marine archaea taxa [31]. These bacteria and archaea are the important components of primary global productivity (e.g. photosynthesis) and geochemical cycling. To date, approximately 2.9×105 bacterial and archaeal strains have been isolated and approximately 1.3×104 bacterial and archaeal taxa have been identified [40]. The bacterial and archaeal diversity we have discovered is less than 0.3% of the total diversity based on conservative estimates. Although our knowledge is still limited, studies have shown that microbial communities are valuable resources and effectively managing these resources are the essential to a sustainable environment [175]. Human beings has taken advantage of microbial systems for millennia, dating back to early wine and cheese making. More recently we have used bacterial communities and populations to remediate organic wastes, produce commodity chemicals and contain subsurface contamination. Understanding the interactions among microbial populations is a vital step in managing microbial resources. Microbial populations react to system changes and undesirable shifts in a microbial community can result in system failures. As an example, sludge bulking in wastewater treatment plants is caused by the excessive growth of filamentous microorganisms in activated sludge. While the presence of filamentous microorganisms helps in the degradation of organic matter, when abundant, filamentous microorganisms prevent activated sludge from settling, resulting in failure of the wastewater treatment system [61, 131]. Recent advances in our understanding of the human microbiome also suggest that the changes in composition of bacterial communities, which reside in different parts of human body, can alter human health [4, 127, 150, 173]. Microbial ecology studies the micro21 bial population interactions, community structures and interactions with the environment including changes due to external factors. Knowing more about these interactions will help us to better utilize microbial resources and control engineered systems. Each topic of my dissertation involves a different microbial community with a unique energetic strategy. The lake sediment community of Torch Lake has been under extreme pressure from unnaturally high contamination of metals and the R. pickettii strains isolated from these sediments are uniquely adapted to high copper concentrations. Thus, these strains can provide insight into how microbial populations adapt to high metal concentrations. The sediment communities of Oak Ridge National Laboratory iFRC sites have responded to metal, acid and hydrocarbon contamination and many strains isolated from these communities can immobilize uranium. We have extended the community analyses of these sediments, as well as identified new strains capable of uranium immobilization. This work contributes to our understanding of the interactions between uranium and microbial populations and provide insights to in situ bioremediation of uranium. Finally, The microbial community of an anaerobic bioreactor managing farm waste has a complex community of bacteria and archaea that removes organic matter and generates methane. We have identified a community structure that correlates with production of a large amount of methane. In each instance, knowing the conditions under which the community operates optimally is advantageous to the engineer. By "optimally" we refer to an outcome that is successful for long-term bioremediation and sustainability. In most cases knowing first, who is there (i.e. what species are present and their growth preferences) and second, what is the metabolic strategy of the community, can assist engineers in designing and controlling an environment that imposes sufficient constrains on the community such that it becomes predictable. Successful 22 management requires predictable outcomes. 23 CHAPTER 2 CO-EVOLUTION OF RALSTONIA PICKETTII STRAIN 12D AND 12J: GENOME REARRANGEMENTS AND THEIR INFLUENCES ON THE EVOLUTION OF METAL RESISTANCE 2.1 Abstract The analysis of natural divergence of phylogenetically closely related bacteria residing in the same habitat can reveal fine-scale evolution events. We have previously reported on Ralstonia pickettii strains (12J and 12D) isolated from the same cubic centimeter of coppercontaminated lake sediment that bioaccumulate copper. While these strains have nearly identical 16S rRNA gene sequences (99% similar), they have distinct genomic structures on which we report here. Strain 12J has two chromosomes, a single plasmid, and a filamentous phage, while strain 12D has two chromosomes, three plasmids, and a larger genome by 359,629 bases. Both strains have adapted to the high concentrations of metals using gene duplication and horizontal transfer of metal resistance genes. Plasmids and phage have played a role in the genetic divergence of the strains. Plasmid 2 of strain 12D contains 31 copper and cobalt-zinc-cadmium resistance genes. This plasmid was inserted into the primary chromosome and initiated, we postulate, a cascade of genetic events that have continued the divergence of the two strains. Subsequent events included differences in horizontally transferred genes, acquisition of a filamentous phage, and differences of transposase carriage, with ISXo4 and IS4 unique to strain 12J and the mutator type of transposase unique to strain 12D. Five transposase types are found in both strains. The genomic evolution of strains 12D and 12J provide a view into a collection of dynamic events shaping and reshaping the 24 Ralstonia genome. 2.2 Introduction Microdiversity describes a group of bacterial populations found in the same environmental niche that are phylogenetically closely related (ie. same species) but physiologically distinct [1, 124, 141]. Studying genomic variations within a bacterial population can give us insights into: 1) the impacts of the environmental factors on the bacteria populations; 2) the incidences of horizontal gene transfer (HGT) and gene duplication and amplification (GDA), that led to the divergence; 3) the extent and location of plasticity within the genome; 4) and the functional potential of the microdiversity [21, 98, 124, 135, 146]. We have a collection of bacteria strains from a Ralstonia pickettii microdiversity. Two representative strains, R. pickettii 12D and 12J were isolated from within the same 1 cm3 of heavy metal (especially copper) contaminated lake sediment (16 cm below the sediment surface at a lake depth of 37 m from Torch Lake, MI). Konstantinidis et al. [80] showed that these two strains are 99% identical according to their 16S rRNA gene sequences. Previous research showed that both strains thrived in copper supplemented media with minor differences in growth rates and copper sequestration abilities [191]. Metal resistance tests revealed that R. pickettii 12D is Cu-Ni-Cd resistant and 12J is Cu-Zn-Cd resistant [80]. We also found that R. pickettii 12J contains filamentous phage (p12J) particles and strain 12D is p12J resistant [191]. These physiological differences indicate that R. pickettii 12D and 12J have distinct genome structures. By studying the genomic variations of these two R. pickettii strains we will better un- 25 derstand the evolutionary consequences of a high copper environment, produce a detailed description of the genomic changes that have occurred between strains and improve our understanding of strain divergence and speciation. Herein, we report our findings on the genomic differences, evidence of HGT and GDA, identification of metal resistance determinants and the evolutionary events that led to the differentiation of R. pickettii 12D and 12J. 2.3 2.3.1 Materials and Methods Genome Retrieval and Gene Annotations The genomic DNA of R. pickettii 12D and 12J were isolated in our laboratory and sequenced by Joint Genome Institute (JGI) as reported previously [191]. The genome sequences of R. pickettii 12D and 12J were retrieved from GeneBank database by using the following IDs, NC_012849, NC_012851, NC_012855−57, NC_010678, NC_010682-83, and NC_005131. The basic gene annotations were obtained from Joint Genome Institute Integrated Microbial Genomes (JGI-IMG) database. Basic Local Alignment Search Tool (BLAST) and JGI-IMG genome BLAST was used to identify and confirm gene annotations (E-value ≤ 10−2 , protein identity ≥ 30%, unless specified). 2.3.2 Genome Comparisons Functional groups identified as clusters of orthologous groups (COG) were used to find shared and unique genes present in R. pickettii 12D and 12J. Metabolic pathway information was retrieved from MetaCyc [19] and Kyoto Encyclopedia of Genes and Genomes 26 (KEGG) database [73]. Genome replication origins (ori C) of R. pickettii 12D and 12J were retrieved from the DoriC database [48]. The replication resolution sites (dif ) of 12D and 12J were located by searching for sequences comparable to the R. solanacearum GMI1000 dif sequence by BLAST (BLASTN). The phylogenetic positions of strain 12D and 12J were analyzed, along with a collection of β-Proteobacteria dif sequences [18], by using Molecular Evolutionary Genetics Analysis (MEGA) [161]. Briefly, the dif sequences were aligned by ClustalW and a minimal evolution (ME) phylogenetic tree was constructed with the following parameters: the Jukes-Cantor substitution model assuming uniform rate among sites, gaps/missing data treated as pairwise deletions, and finally clustering with the ME heuristic method by allowing close-neighbor-interchange. Bootstrapping (1000 replications) was used to test the confidence level of branching. Genome statistics (e.g., copies of rRNA coding genes, protein coding genes, horizontally transferred genes, etc.) of R. pickettii 12D and 12J and phylogenetically related bacteria (R. solanacearum GM1000 and Cupriavidus metallidurans CH34) were retrieved from JGI-IMG and compared based on their annotation. Transposases encoding sequences were identified by JGI-IMG. The genomic syntenies of R. pickettii 12D and 12J were assessed by using Mauve [33], a multiple genome alignment tool. 2.3.3 Identification of the Metal Resistant Genes Metal (Co-Zn-Cd, Cu, and Hg) resistant genes in R. pickettii 12D and 12J were extrapolated from JGI-IMG based on the metal resistant genes identified in C. metallidurans CH34. The abundances of these gene copies in R. pickettii strains were compared to those found in C. metallidurans CH34 and R. solanacearum GMI1000. The evolution lineage of the 27 main subunit of these resistance determinant coding genes in R. pickettii 12D and 12J was determined by comparing to similar coding sequences found in 23 other Proteobacteria species retrieved from JGI-IMG (Figure A.1). The phylogenetic analysis of these genes was conducted by using MEGA. Briefly, the DNA sequences were aligned via ClustalW and the bootstrapped (1000 replications) ME trees were constructed with the following parameters: the Jukes-Cantor substitution model assuming uniform rate among sites, gaps/missing data treated as pairwise deletion, and close-neighbor-interchange for ME heuristic method. 2.4 2.4.1 Results Genome Overview JGI-IMG annotations identified 1936 and 1945 COG groups in 12D and 12J, respectively. Between the two strains, 1876 COG groups are shared. KEGG data revealed that the metabolic pathway maps of both strains are highly similar with minor differences (Appendix A). MetaCyc revealed that both strain 12D and 12J contain several detoxification pathways involved in arsenate reduction, cytoplasm Hg2+ reduction, methylglyoxal detoxification, cyanate degradation and superoxide radical degradation (Appendix A). R. pickettii 12D and 12J clustered closely to R. solanacearum GMI1000 and Cupriavidus metallidurans CH34 based on their 16S rRNA gene sequences (Figure A.1). Basic comparisons among these four strains showed that R. pickettii 12D and 12J are smaller in genome size compared to R. solanacearum GMI1000 and C. metallidurans CH34. R. pickettii 12J has the least number of bases and protein coding genes (Table 2.1). Multiple copies of rRNA genes are found in all four bacterial strains. Both R. pickettii strains contain three copies of the 16S 28 rRNA genes. IMG annotations indicate that all four strains contain comparable numbers of orthologs and paralogs and R. pickettii 12D and 12J contain the least number of horizontally transferred genes. All four strains contain many signal peptide and transmembrane protein encoding genes (Table 2.1). The complete genome sequences revealed that the R. pickettii 12D genome is composed of a primary chromosome (12D_C1) of 3.6 Mb, a secondary chromosome (12D_C2) of 1.3 Mb, and three plasmids of 0.39 Mb (12D_P1), 0.27 Mb (12D_P2), and 0.051 Mb (12D_P3) (Figure 2.1). R. pickettii 12J contains a primary chromosome (12J_C1) of 3.9 Mb, a secondary chromosome (12J_C2) of 1.3 Mb, and a plasmid (12J_P1) of 0.08 Mb (Figure 2.2). R. pickettii 12J_C1 is approximately 0.3 Mb larger than 12D_C1. The secondary chromosomes of both strains are approximately the same size and substantially smaller than the primary chromosomes. R. pickettii 12D_P3 and 12J_P1 are comparable in size. Two large plasmids (P1 and P2) in 12D are absent in 12J. We also found that strain 12J contains filamentous phage particles (p12J), which is absent in strain 12D. Horizontally transferred foreign genes (JGI-IMG prediction) are scattered across all genome scaffolds and are twice as abundant in 12D than 12J (Table A.1). A total of 123 genes were predicted as being horizontally transferred genes in R. pickettii 12J and the majority of them are in 12J_C1 (Table A.1). Out of the 287 horizontally transferred genes in strain 12D, 115 are in 12D_C1 and 157 are in 12D_P1 (Table A.1). The JGI-IMG gene profiler revealed that strain 12D and 12J share only one gene among all 410 predicted horizontally transferred genes (JGI-IMG gene profiler with E-value ≤ 1E-5 and protein identity ≥ 80%). This gene is annotated as an IS30 family integrase catalytic region (locus Rpic12D_1028 and Rpic_2522). 29 Table 2.1: Gene summaries of R. solanacearum GMI1000, C. metallidurans CH34, R. pickettii 12D and 12J. Genome Features R. solanacearum GMI1000 C. metallidurans CH34 R.pickettii 12D R. pickettii 12J Bases GC (%) Total Number of Genes Protein Coding Genes rRNA (5S, 16S, 23S) tRNA 5810922 66.98 5204 5120 4,4,4 57 6913352 63.53 6430 6340 2,4,4 62 5685358 63.3 5518 5452 3,3,3 54 5325729 63.65 5092 5027 3,3,3 55 Orthologs Paralogs Signal Peptides Transmembrane Proteins Horizontally transferred 4774 1106 1232 1188 271 (91.74%) (21.25%) (23.67%) (22.83%) (5.21%) 5951 1753 1568 1514 510 30 (92.55%) (27.26%) (24.39%) (23.55%) (7.93%) 4951 1089 1349 1339 287 (89.72%) (19.74%) (24.45%) (24.27%) (5.20%) 4774 1106 1286 1291 123 (93.75%) (21.72%) (25.26%) (25.35%) (2.42%) 0 3.6 0.3 3.3 0.6 3.0 0.9 C1 2.7 1.2 2.4 2.1 1.8 1.5 Figure 2.1: R. pickettii 12D genome scaffolds. The distribution of putative horizontally transferred genes (HGT genes) identified by JGI-IMG are marked as black dashes in rings 4-10. The scaffold sizes are in units of Mb. For interpretation of the references to color in this and all other figures, the reader is referred to the electronic version of this dissertation. 31 Figure 2.1 (cont’d) 1.30 1.2 0.1 0.2 1.1 0.3 1.0 C2 0.4 0.9 0.5 0.8 0.6 0.7 32 Figure 2.1 (cont’d) 0 0.3 0.25 P1 0 0.1 0.05 0.2 0 0 5 .0 0.04 P2 0 .2 0.01 0 .1 0.03 0 5 .1 33 P3 0.02 Figure 2.1 (cont’d) 1 2 3 4 5 6 oriC xe rC xe rD d if d naA re A p 7 8 10 1 1 1 2 RingLe e g nd 9 1. Ge sonforwards ne trand 2. Ge sonre rs trand ne ve es 3. RNA g ne e s(tRNAsg e rRNAsre othe black) re n, d, r 4. HGT g ne e sfromArchae a 5. HGT g ne e sfromEukaryota 6. HGT g ne e sfromVirus s e 7. HGT g ne e sfromphylumothe thanProte r obacte ria 8. HGT g ne e sfromclas sothe thanBe r ta-prote obacte ria 9. HGT g ne e sfromorde othe thanBurkhode r r riale s 10. HGT g ne e swithbe t hitstog ne s e sfromGam a-prote m obacte ria andnohitstog ne e sfromBe ta-prote obacte ria 1 GC conte 1. nt 12. GC s w ke 34 0 0.3 3.6 0.6 3.3 3.0 0.9 C1 1.2 2.7 2.4 1.5 1.8 2.1 Figure 2.2: R. pickettii 12J genome scaffolds. The distribution of putative horizontally transferred genes (HGT genes) identified by JGI-IMG are marked as black dashes in rings 4-9. The scaffold sizes are in units of Mb. 35 Figure 2.2 (cont’d) 0 1.2 0.1 0.2 1.1 1.0 0.3 C2 0.4 0.9 0.5 0.8 0.7 0.6 36 Figure 2.2 (cont’d) 0 0.08 0.01 0.07 P1 0.06 0.02 0.03 0.05 0.04 37 Figure 2.2 (cont’d) 1 2 3 4 5 oriC xe rC xe rD d if d naA re A p 6 10 1 1 7 8 9 RingLe e g nd 1 Ge so . ne nfo ards rw trand 2 Ge so . ne nre rs trand ve es 3 RNA g ne . e s(tRNAsg e rRNAsre o r RNAsblack) re n, d, the 4 HGT g ne . e sfro mArchae a 5 HGT g ne . e sfro mEukaryo ta 6 HGT g ne . e sfro mVirus s e 7 HGT g ne . e sfro mphylumo r than Pro o the te bacte ria 8 HGT g ne . e sfro mclas the thanBe so r ta-pro o te bacte ria 9 HGT g ne . e sfro rde o r than Burkho riale mo r the lde s 1 . GC co nt 0 nte 1 . GC s w 1 ke 38 Both R. pickettii strains are abundant in transposases (Table A.1). Strain 12J has 32 transposase genes which belong to seven transposase families and 12D has 33 transposase genes that belong to eight transposase families. Five of these transpoase types are found in both strains. Among the abundant transposases, type ISXo4 and IS4 are unique to strain 12J and the mutator type of transposase unique to strain 12D. The genome replication initiation positions (ori C) are labeled as black arrows in Figure 2.1 and Figure 2.2. The estimated ori C positions are just upstream of dnaA (C1) or repA (C2) genes. R. pickettii 12D_P1 also has a repA gene (Figure 2.1 P1 black hollow arrow), which was not observed in the other plasmids of 12D and 12J. The identification of dif /xerCD system in both strains suggest that this is used to resolve dimers created during replication [56]. Each strain has a pair of xer CD located on the primary chromosomes (Figure 2.1 and Figure 2.2 black and blue solid triangles). Each strain also contains two dif sites, one is located on C1 and the other one is on C2 (Figure 2.1 and Figure 2.2 red solid triangles). There are four copies of kfr A gene in each strain that presumably contribute to plasmid stability. Two of the four kfr A copies are located on the secondary chromosomes, and the other two were found on 12J_C1 and 12D_P1, respectively. 2.4.2 2.4.2.1 Co-evolution of R. pickettii 12D and 12J 16S rRNA Gene Sequence Variations For the three copies of 16S rRNA genes in each R. pickettii strain, we observed that two copies are located on the primary chromosome and the third copy is found on the secondary chromosome. The three copies in strain 12J are all identical. The 16S rRNA gene copy on 39 the 12D secondary chromosome (12D_C2) is identical to the copies in 12J. The two copies of 16S rRNA on the primary chromosome of 12D (12D_C1) are identical to each other but different to the copy on 12D_C2, with a distance of 0.001. 2.4.2.2 dif Site Sequence Variations Many bacteria employ the dif /xer CD system to resolve dimers formed during chromosome replication [18, 25, 34, 68, 109, 144]. This system is comprised of a set of recombinase encoding genes, xer C and xer D, and a 28 bp binding site (dif site) near the replication terminus [18]. Genome sequences revealed that R. pickettii 12D and 12J also utilizes dif /xer CD system. We found each strain contains one set of xer CD genes and two dif sites (one on each chromosome). The dif site sequences from the primary and secondary chromosomes of R. pickettii vary substantially, although they presumably interact with the same sets of XerCD protein. R. pickettii 12J_C1 dif appears to be the most different. The dif sequences on the secondary chromosomes are highly similar (Figure A.2). The ME phylogenetic tree revealed that the R. pickettii dif sequences clustered together, away from other β-Proteobacteria dif sequences (Figure 2.3). The dif sequence of 12J_C1 appears to be the most ancestral among all four R. pickettii. We observed that 12D_C1 and both secondary chromosomes branched out from 12J_C1 and the secondary chromosome dif sequences appeared to more distant to 12D_C1 than 12D_C1 (Figure 2.3). 40 Burkholderia mallei ATCC 23344 chr. 1 Burkholderia mallei ATCC 23344 chr. 2 46 Dechloromonas aromatica RCB 9 Ralstonia solanacearum GMI1000 7 Thiobacillus denitrificans Bordetella pertussis Tohama I 35 Methylobacillus flagellatus KT 22 Rhodoferax ferrireducens DSM 15236 41 Nitrosomonas europaea ATCC 19718 Neisseria meningitidis MC58 Ralstonia pickettii 12J C1 94 Ralstonia pickettii 12D C1 Ralstonia pickettii 12D C2 94 0.05 Ralstonia pickettii 12J C2 98 71 70 Figure 2.3: Minimal evolution phylogenetic tree of β-Proteobacteria dif sequences. The numbers on branches are bootstrap values. The scale at the left hand lower corner represents the distance. 41 2.4.2.3 Filamentous Phage Elements in R. pickettii 12D and 12J Orf 1 2 34 5 p12J: p12Jp01-10 6 7 8 9 10 12J-C1_1: Rpic_1397-1405 12J-C1_2: Rpic_1391, 1393-95 Replication initiation factor 12J-C1_3: Rpic_2540, 2542-4, 2546-9 Repressor 12J-C2: Rpic_4343-5, 4347, 4349-50 Resolvase invertase Repressor 12D-C1: Rpic12D_2133, 2142 Resolvase invertase 12D-C2: Rpic12D_4471, 4475 Replication initiation factor phiLf Zonular Protein occludens toxin Figure 2.4: Filamentous phage p12J gene compositions and their homologs on R. pickettii 12D and 12J chromosomes. Fragments of the same fragments represent homologous genes. Genes colored yellow with a red border represent genes that share no homology to p12J Orf7 but are orthologs to each other. Grey colored fragments represent other phage related genes (not p12J related). White colored blocks are other non-phage related genes. Filamentous phage p12J (NC_005131) is comprised of ten open reading frames (orf), locus p12Jp01-10 (Figure 2.4 p12J). Three (orf8 - orf10) out ten of the p12J orf’s were annotated as phage proteins but with little information. Orf5, orf7 and orf10 are the main 42 phage assembly proteins and orf7 was identified as a zonular occludens toxin coding gene. We located four incidents of p12J insertion in R. pickettii 12J (Table A.1). It was observed that the complete phage genome is located on 12J_C1, 464 bp upstream from the 12J_C1 dif site (Figure 2.4 12J-C1_1). Upstream adjacent to this complete phage genome is a partial p12J copy with a non-p12J related phage replication initiation factor encoding gene (Figure 2.4 12J-C1_2). The third insertion is also located on on 12J_C1 (Figure 2.4 12JC1_3). We found the fourth partial copy on 12J_C2 (Figure 2.4 12J-C2), which is located 931 bp downstream of 12J_C2 dif site. All three insertions in 12J contain at least four p12J gene homologs, two out of which are primary phage components (orfs 5 & 7). In contrast, only two p12J partial copies were found on 12D_C1 (Figure 2.4 12D-C1) and 12D_C2 (Figure 2.4 12D-C2). Both insertions are distal from R. pickettii 12D dif sites and appear to be highly truncated by either non-phage related genes or other phage (non-p12J) related genes. These 12D partial copies contain only two p12J homologs. Annotation showed that one of the genes encodes zonular occludens toxin. However, it shares no synteny with p12J orf7 (Figure 2.4 12D-C2). 2.4.3 Genome Syntenies of R. pickettii 12D and 12J Genomic DNA sequence dotplot revealed that R. pickettii 12D and 12J share high synteny (Figure A.3). Detailed genomic sequence alignments showed that R. pickettii 12J_C1 could be viewed as homologs of R. pickettii 12D_C1 and 12D_P2. The 12D_P2 homolog was located between 1.56 Mb and 1.95 Mb on 12J_C1, approximately 0.07 Mb downstream of the dif site (Figure 2.5 A). Figure 2.5 panel B indicated where the genomes where rearranged. The replication initiation region of 12D_C2 appears to be drastically different from 43 A. R. pickettii 12D C1 12D P2 R. p t ii 12J C1 icke t B. R. pickettii 12D C2 R. p t ii 12J C2 icke t C. R. pickettii 12D P3 R. p t ii 12J P1 icke t Figure 2.5: Genome syntenies between R. pickettii 12D and 12J scaffolds. Panel A, 12D_C1, 12J_C1 and 12D_P2; Panel B, 12D_C2 and 12J_C2; Panel C, 12D_P3 and 12J_P1. Arrows indicate location of ori C on the chromosomes. Blocks with the same color represent the regions that are similar (connected by lines). The amount of color indicates the percentage of identity. White areas within the blocks indicate absence of a homologous sequence. 44 annotated 0 position. If repositioned, both secondary chromosomes would start with repA (12D_C2 is almost identical with 12J_C2, Figure refmauve B). The 12J_P1 and 12D_P3 alignment revealed a similar genome rearrangement pattern (Figure 2.5 C). We could not locate the origins of replication in this study. The large plasmid of strain 12D (12D_P1) showed little synteny with the other scaffolds and appears unique to itself. The 0.39 Mb 12D_P2 homolog inserted into 12J_C1 is comprised of two distinct portions. The region from 1.56 Mb to 1.72 Mb appears to be abundant with low GC content genes that encode secretion systems and cytoskeleton. Within this region, we observed two copies of plasmid partition protein coding gene par B (locus Rpic_1509 and Rpic_1596) [123]. The section from 1.72 to 1.95 Mb codes for multiple metal resistance genes [191]. This region was extracted and annotated by hand using BLAST targeting the non-redundant protein database of GenBank. We obtained 408 hits with alignment scores over 200 with diverse taxonomy indicating that the region is a phylogenetic mosaic of horizontally transferred genes (Figure 2.6). The query sequence revealed itself as a "mosaic", where multiple genes from different bacteria were assembled to reconstruct the 0.23 Mb region of R. pickettii 12J_C1 (Figure 2.6). As we expected, 12D_P2 shares the most similarities. A fragment of 12J_C1 sequence (between 134000 and 150000) was highly similar (99% identity) to 25 genes on plasmid pMOL30 of C. metallidurans CH34 (0.71 Mb to 0.87 Mb). Finally, the KfrA protein coding gene is also found in the query sequence, which is 100% identical to those kfr A genes observed in R. pickettii 12J_C2, 12D_C2, and 12D_P1 (Figure 2.6). Within this region of R. pickettii 12J_C1, we observed four copies of genes related to cobalt-zinc-cadmium resistance (czcA), one mercury resistance operon and three copper resistance operons. Besides these specific metal resistances, this region also contains multiple copies of genes encoding 45 for heavy metal translocating P-type ATPase, ABC transporter, iron permease, cation diffusion transporter, and heavy metal transport/detoxification protein. Scattered among these resistance genes, were transposes, tyrosine-based site-specific recombinase, and hypothetical protein encoding genes (Figure 2.6). 46 Figure 2.6: BLAST results mapped onto the 1.72 Mb-1.94 Mb region of R. pickettii 12J_C1. The top scale indicates the length of the query region. Similar genes are shown underneath the query sequence. The short thin black bars indicate that the next stretch of sequence is on the opposite strand of DNA. 47 Figure 2.6 (cont’d) 48 2.4.4 2.4.4.1 The Acquisition and Duplication of Resistance Determinants in R. pickettii 12D and 12J Metal Resistant Genes in R. pickettii 12D and 12J Annotations from the R. pickettii 12D_P2 and the inserted region of 12J_C1 revealed a high abundance of ABC transporters, P-type ATPase, and metal (Cu, Co-Zn-Cd, and Hg) resistance determinant encoding genes (Figure 2.6). To determine the number of metal resistant gene copies that reside in R. pickettii 12D and 12J, we employed IMG genome BLAST to extrapolated possible metal resistant genes specifically related to Co-Zn-Cd, Cu and Hg resistance. Co-Zn-Cd (czc operon), Cu (cop operon) and Hg (mer operons) resistant genes found in C. metallidurans CH34 were used as reference genes. Two cop operon genes, copR and copS, were not detected in C. metallidurans CH34. Thus, copRS found in Pseudomonas aeruginosa LESB58 were used as references for identifying copR and copS. The abundance and distribution of these genes are reported in Table 2.2. We observed multiple occurrence of czc operon in all four bacterial strains. The major Co-Zn-Cd pump encoding gene, czcA is the most abundant, which was found 13 times in CH34, 9 times in 12J, 8 times in 12D, and 4 times in GMI1000. Two different sets of czcBCD were located on all four strains. R. pickettii 12J contains the most copies of czcD (13) and the least amount of czcC (2). R. pickettii 12J contains the most amount of czcE, a copper binding protein coding gene [195], and czcP, which is a transition metal efflux coding gene [138]. Of the four strains, only C. metallidurans CH34 and R. pickettll 12J carry czcN while czcI was found only in CH34. Copper resistant determinants were found in all four strains but were most abundant in R. pickettii 12D and 12J. C. metallidurans contains two copies of each cop operon gene and 49 only one copy of each cop gene was found in R. solanacearum GMI1000. The major cop operon component, copA, a Cu(I)-translocating P-type ATPase was observed 5 and 3 times in 12J and 12D, respectively. The regulatory and signaling protein encoding genes of czc and cop operon appear to be highly similar and were reported together. A mercury resistant determinant is found in R. pickettii 12J and C. metallidurans CH34 only. Although the regulatory gene, mer R, was observed in both R. pickettii 12D and R. solanacearum GMI1000, the gene encoding for mercury reductase (mer A) was absent from both. We identified two copies of mer P, a mercuric transporter protein encoding gene, and one copy of mer A in R. pickettii 12J. In contrast, three copies of mer A and four copies of mer P were found in C. metallidurans CH34. The distribution of metal resistant genes on R. pickettii 12D and 12J genome revealed that majority of the metal resistant genes are located on 12D_P2 and 12J_C1. The secondary chromosomes of both strains carry the same number of the metal resistant genes. No Co-Zn-Cd, Cu, or Hg resistant genes were found on R. pickettii 12D_P1 and 12D_P3. The total copy number of each metal resistant gene revealed that strain 12J contains an equal or great number of copies than 12D, with the exception of czcC (Table 2.2). 50 Table 2.2: The abundance of Co-Zn-Cd, Cu, and Hg resistant genes in R. pickettii 12D and 12J, R. solanacearum GMI1000, and C. metallidurans CH34. czcA P2 12D C2 12D 12J P3 12D P1 12J P1 12D czcE czcP czcI czcN 1 0 0 0 0 1 2 0 0 0 0 5 1 4 0 0 7 5a 4 3 0 1 4 3 2 0 1 1 0 1 2 0 0 3 0 1 1 1 0 0 0 0 0 0 3 0 1 1 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 cop Operon C1 Metal Resistance Genes 12D 12J copA copB copC copD 0 0 0 0 5 4 4 3 3 4 4 2 0 0 1 1 0 0 1 1 0 0 0 0 0 0 0 0 0 0 0 0 czc/cop czcR/copR czcS/copS 0 1 5 6 2 2 0 1 0 1 0 0 0 0 0 0 mer Operon czc Operon czcB czcC czcD mer A 0 1 0 0 0 0 0 mer E 0 0 0 0 0 0 0 mer P 0 2 0 0 0 0 0 mer R 3 5 0 2 2 0 0 mer T 0 1 0 0 0 0 0 a. Two out of five matches are shared by Rmet_5979 matches. b. IMG genome BLAST with e-value of 1E-2 and at least 40% identity. 51 0 0 0 0 0 R. pickettii 12D R. pickettii 12J R. solanacearum GMI1000 C. metallidurans CH34 Reference Locus czcE czcP czcI czcN 8 3 3 1 2 2 2 1 2 0 0 9 1 5 1 1 7 6 4 3 0 1 4 1 4 1 2 1 1 0 0 0 0 13 4 4 2 1 1 1 3 2 1 1 Rmet_5980 Rmet_5981 Rmet_4121 Rmet_5982 Rmet_4120 Rmet_5979 Rmet_3406 Rmet_5976 Rmet_5970b Rmet_4595 Rmet_5984 cop Operon Table 2.2 (cont’d) copA copB copC copD 3 4 5 3 5 4 5 4 1 1 1 1 2 2 2 2 Rmet_6112c Rmet_6113 Rmet_6114 Rmet_6115 czc/cop czcR/copR czcS/copS 2 4 5 7 2 2 5 12 Rmet_5978/PLES_22631d Rmet_5977/PLES_22621e Metal Resistance Genes czcA mer Operon czc Operon czcB czcC czcD mer A 0 1 mer E 0 0 mer P 0 2 mer R 5 7 mer T 0 1 c. IMG genome BLAST with e-value of 1E-5 and at d. IMG genome BLAST with e-value of 1E-5 and at e. IMG genome BLAST with e-value of 1E-2 and at 0 3 Rmet_6174 0 1 Rmet_6176 0 4 Rmet_6173b 5 12 Rmet_6171 0 3 Rmet_6172 least 50% identity. Matches contain all 3 Pfam domains. least 60% identity. Both reference loci return the same results. least 40% identity for PLES_22621. 52 2.4.4.2 The Phylogenetic Lineage of czcA Copies in R. pickettii 12D and 12J Table 2.2 shows that copies of czcA are distributed on 12D_C1 (1), 12D_C2 (3), and 12D_P2 (4) in R. pickettii 12D and 12J_C1 (5), 12J_C2 (3), and 12J_P1 (1) in R. pickettii 12J. To determine their phylogenetic lineage, a collection of 85 czcA sequences from 25 Proteobacteria (Figure A.1) was used to construct a minimal evolution phylogenetic tree (Figure 2.7). We observed that both R. pickettii 12D and 12J contain czcA genes that share evolutionary lineages to copies found in C. metallidurans CH34, R. eutropha JMP134, R. solanacearum Po82 and GMI1000. We observed that of eight copies of czcA genes in 12D, six copies clustered with those from 12J (Figure 2.7 Group I-VI). The phylogenetic tree showed that some copies of czcA from the same strain clustered together, such as 12D P2 copy 1 and copy2, and 12J C1 copy 4 and copy 5. One of the R. pickettii 12J czcA copy (12J C1 copy 1) appeared to be unique and did not cluster with the rest of the R. pickettii copies (Figure 2.7). 53 Ralstonia eutropha H16 C2 (-) Cupriavidus necator N-1 C2 (-) 82 Cupriavidus taiwanensis str. LMG19424 C2 (-) Ralstonia solanacearum GMI1000 plasmid pGMI1000MP (+) 100 Ralstonia solanacearum Po82 megaplasmid (+) 41 Ralstonia eutropha JMP134 C2 (-) 100 Cupriavidus metallidurans CH34 C2 (-) 90 67 Cupriavidus metallidurans CH34 plasmid 1 (-) 100 Ralstonia pickettii 12J C1 (-) copy 1 79 Cupriavidus taiwanensis plasmid pRALTA (-) Pseudomonas aeruginosa PA7 (+) 100 Pseudomonas fluorescens Pf-5 (-) Pseudomonas syringae pv. tomato str. DC3000 (-) 99 100 Pseudomonas putida F1 (-) 93 Pseudomonas putida F1 (-) 38 Burkholderia xenovorans LB400 C2 (-) Burkholderia sp. 383 C2 (-) 100 100 100 Burkholderia multivorans ATCC 17616 plasmid pTGL1 (-) Burkholderia thailandensis E264 C2 (+) 99 Burkholderia mallei ATCC 23344 C2 (-) 100 100 100 Burkholderia pseudomallei 1710b C2 (-) Cupriavidus necator N-1 plasmid BB1p (+) Ralstonia pickettii 12D P2 (+) copy 1 Ralstonia pickettii 12D P2 (-) copy 2 100 Cupriavidus metallidurans CH34 plasmid 1 (-) 100 Cupriavidus metallidurans CH34 plasmid 2 (+) 100 100 100 Figure 2.7: Phylogenetic position of R. pickettii 12D and 12J czcA genes. The numbers on the branches represent bootstrap values. 54 Figure 2.7 (cont’d) 100 Ralstonia pickettii 12J C2 (-) copy 2 Group I Ralstonia pickettii 12D C2 (-) copy 3 Ralstonia eutropha JMP134 C2 (+) Cupriavidus metallidurans CH34 C2 (-) Curpiavidus metallidurans CH34 C1 (-) Cupriavidus metallidurans CH34 C2 (-) Burkholderia thailandensis E264 C2 (-) 100 Geobacter uraniumreducens Rf4 (+) 100 Ralstonia pickettii 12J C2 (-) copy 3 Group II 100 Ralstonia pickettii 12D C2 (-) coy 4 Ralstonia solanacearum Po82 megaplasmid (+) Ralstonia solanacearum GMI1000 plasmid pGMI1000MP (-) 100 Cupriavidus metallidurans CH34 C2 (-) Ralstonia eutropha H16 C2 (+) 100 100 Cupriavidus necator N-1 C2 (+) 100 Cupriavidus taiwanensis str. LMG19424 C2 (+) Ralstonia eutropha JMP134 C2 (+) 56 Cupriavidus metallidurans CH34 C2 (+) Cupriavidus metallidurans CH34 C2 (-) Ralstonia solanacearum Po82 megaplasmid (-) 100 100 Ralstonia pickettii 12J C1 (+) copy 4 100 Ralstonia pickettii 12J C1 (-) copy 5 Ralstonia pickettii 12J P1 (-) copy 6 100 Group III 100 Ralstonia pickettii 12D C1 (-) copy 5 66 76 67 75 39 100 99 62 100 55 Figure 2.7 (cont’d) Geobacter uraniumreducens Rf4 (-) Geobacter lovleyi SZ (-) Geobacter metallireducens GS-15 (+) Ralstonia solanacearum GMI1000 plasmid pGMI1000MP (+) 100 Ralstonia solanacearum Po82 megaplasmid (+) 100 44 Ralstonia pickettii 12J C2 (-) copy 7 Group IV 100 Ralstonia pickettii 12D C2 (-) copy 6 100 Cupriavidus metallidurans CH34 C2 (+) Ralstonia eutropha JMP134 C2 (+) 100 Cupriavidus taiwanensis str. LMG19424 C2 (-) 100 Ralstonia eutropha H16 C2 (-) 100 100 Cupriavidus necator N-1 C2 (-) Geobacter lovleyi SZ (+) Geobacter metallireducens GS-15 (-) 100 100 Geobacter uraniumreducens Rf4 (+) 100 Burkholderia mallei ATCC 23344 C1 (+) Burkholderia pseudomallei 1710b CI (+) Burkholderia multivorans ATCC 17616 C3 (+) 100 Burkholderia multivorans ATCC 17616 plasmid pTGL1 (+) 60 100 78 94 100 56 Figure 2.7 (cont’d) Cupriavidus metallidurans CH34 plasmid 1 (+) Cupriavidus metallidurans CH34 C2 (+) 100 100 Ralstonia eutropha H16 C1 (+) Cupriavidus necator N-1 C1 (+) Ralstonia solanacearum GMI1000 plasmid pGMI1000MP (+) 100 95 Ralstonia solanacearum Po82 megaplasmid (+) 100 Ralstonia pickettii 12J C1 (-) copy 8 Group V 98 Ralstonia pickettii 12D P2 (-) copy 7 Ralstonia pickettii 12J C1 (+) copy 9 82 Group VI 100 Ralstonia pickettii 12D P2 (-) copy 8 Pseudomonas putida F1 (+) Pseudomonas aeruginosa PA7 (+) 100 Escherichia coli APEC O1:K1:H7 plasmid pAPEC-O1-R (+) Escherichia coli APEC O1:K1:H7 (+) 100 Escherichia coli O157:H7 Sakai EHEC (+) 100 53 Escherichia coli O44:H18:042 EAEC (+) 0.05 57 2.4.4.3 The Phylogenetic Lineage of copA Copies in R. pickettii 12D and 12J We obtained 24 copA sequences from 25 genomes reported in Figure A.1 and constructed a minimal evolution tree (Figure 2.8). Fewer copA copies were found and they appeared to be highly similar to those found in R. solanacearum Po82, C. metallidurans CH34, and γproteobacteria (Pseudomonas and Escherichia groups). We observed three incidences where copA copies from 12J and 12D clustered together (Figure 2.8 Group I-III). R. pickettii 12J C1 copy 2 clustered close to Group I and 12J C1 copy 5 appeared to be distant from all of the other copies (Figure 2.8). R. pickettii 12D_P2 carries three copA copies which were clustered into two distinct groups (Figure 2.8 12D P2 copy 1/2 and 12D P2 copy 3), suggesting the occurence of multiple horizontal transfer events. 58 Ralstonia pickettii 12J C1 (+) copy 1 Group I 90 Ralstonia pickettii 12D P2 (+) copy 1 96 Ralstonia pickettii 12J C1 (-) copy 2 Ralstonia solanacearum Po82 megaplasmid (-) 76 Ralstonia solanacearum GMI1000 plasmid pGMI1000MP (+) Ralstonia pickettii 12J C1 (-) copy 3 Group II 68 100 Ralstonia pickettii 12D P2 (+) copy 2 Cupriavidus metallidurans CH34 plasmid 1 (+) Cupriavidus metallidurans CH34 C2 (-) 69 Ralstonia eutropha JMP134 C1 (-) 58 Ralstonia eutropha H16 C2 (-) 76 98 Cupriavidus taiwanensis str. LMG19424 C2 (-) Burkholderia multivorans ATCC 17616 C1 (-) Geobacter lovleyi SZ (+) Ralstonia pickettii 12J C1 (+) copy 4 100 Group III 56 Ralstonia pickettii 12D P2 (-) copy 3 Pseudomonas fluorescens Pf-5 (-) 62 Pseudomonas aeruginosa PA7 (+) 99 Pseudomonas syringae pv. tomato str. DC3000 (+) 45 Escherichia coli APEC O1:K1:H7 plasmid pAPEC-O1-R (+) 97 Pseudomonas putida F1 (-) 69 100 Pseudomonas putida F1 (-) Ralstonia pickettii 12J C1 (+) copy 5 100 0.05 Figure 2.8: Phylogenetic position of R. pickettii 12D and 12J copA genes. The numbers on the branches represent bootstrap values. 59 2.5 Discussion Studying the development of microdiversities gives insights on how environmental factors influence the evolutionary events and how the specific adaptation strategies evolved. In this study, we investigated the genomic differences of R. pickettii 12D and 12J and identified possible evolutionary events that occurred as a result of the selective pressure of metal contamination. Under the same amount of heavy metal stress, R. pickettii diverged and evolved into two distinct genomovars. Herein, we report the genomic differences between the two strains and the possible evolutionary events behind the divergence. Interestingly, in what would appear to be a relatively constant environment, the genome structures of R. pickettii 12D and 12J are drastically different (5 scaffolds in 12D and 3 scaffolds in 12J). However, both strains share large amounts of genetic information (Figure 2.5 and Figure A.3). Figure 2.5 revealed that 12J_C1 is composed of 12D_C1 and 12D_P2, the secondary chromosomes of both strains are nearly identical, and 12J_P1 could be viewed as 12D_P3 with the addition (or insertion) of 0.028 Mb of 12D_C1. The largest plasmid 12D_P1 is unique to strain 12D. Thus R. pickettii 12D and 12J in fact share genetic information across four genome scaffolds. We observed substantial variations in the R. pickettii dif site sequences. Both R. pickettii 12D and 12J utilize the same set of xer CD genes, located on the primary chromosomes, to resolve dimers created during replications on both the primary and secondary chromosomes. Carnoy et al. [18] have reported that this is a common phenomenon observed in bacteria with multiple chromosomes. It has also been shown that a single set of xer CD can bind to distinct dif sequences [18, 170]. The ME tree of β-Proteobacteria dif sequences suggested 60 that the 12D_C1 dif copy is the closest to the ancestral R. pickettii copy and the other three copies from 12D and 12J were derived from the dif sequence that is similar to 12J_C1 dif (Figure 2.3). We have previously reported that live filamentous phage particles were isolated from strain 12J and the intracellular replicative form of the virus [191]. This is consistent with our genome analysis, where an intact copy of the filamentous phage genome was found on R. pickettii 12J_C1. Besides the integrated p12J, there are three copies of p12J like sequences in R. pickettii 12J (two on 12J_C1, and one on 12J_C2). These three copies contain at least four genes that share syntenies with p12J genes (Figure 2.4). Two out of three vestigial copies were integrated near 12J dif sites. Huber et al. [63] has shown that filamentous phage can be integrated into bacterial chromosomes near dif sites only with the assistance of dif /xer CD systems. This is consistent with our observation of the integrated filamentous phage. Copy 12J-C1_3, which locates distant to the dif site, might be originally integrated into ancestral R. pickettii and degraded through evolutionary events. Similarly, we located two vestigial p12J copies in 12D_C1 and 12J_C2. These two copies resemble p12J in structure but contain only two genes that share homology with p12J and their insertion positions are distal to 12D dif sites. This observation suggests that copies 12D-C1 and 12DC2 were integrated into ancestral R. pickettii and later mutated (Figure 2.4). Unpublished work in our lab demonstrated that R. pickettii 12D could not be infected by p12J. One could speculate that the two vestigial p12J copies in 12D_C1 and 12J_C2 that resemble p12J in structure might be contributing to the phage resistance. The analyses of 16S rRNA gene sequences, dif sites variations, and p12J insertions suggest that the secondary chromosome was present in the common ancestor to 12J and 12D. 61 It has been frequently proposed that bacterial secondary chromosomes evolved from plasmids with insertions of house keeping genes [123, 149]. Supporting this hypothesis, we have found that both secondary chromosomes of R. pickettii 12D and 12J contain the plasmid stabilization protein encoding gene, kfr A. Cooper et al. [28] also demonstrated that secondary chromosomes evolve faster than primary chromosomes. Surprisingly, the secondary chromosomes of R. pickettii 12D and 12J appeared to undergo many fewer changes than the primary chromosomes (Figure 2.5). We do not know the reasons for this apparent discrepancy in Ralstonia. One obviously major evolutionary event in R. pickettii 12D and 12J was the integration of 12D_P2 into the ancestral R. pickettii primary chromosome. This inserted region is 0.07 Mb downstream of 12J_C1 dif site. The "mosaic" pattern of the inserted region in 12J_C1 indicates that the ancestral primary chromosome acquired more genes from other bacteria after the insertion (Figure 2.6). With this in mind, it is logical to assume that plasmid 12D_P2 transferred into the common ancestor and eventually became inserted into chromosome 1 of what was to become 12J. This common ancestor also contained 12J_P1 (12D_P3) that has been preserved in the two lineages along with C1 and C2. A 0.03 Mb portion of 12D_C1, which was missing in 12J_C1, was also found in 12J_P1 (Figure 2.5 C). This suggests that the insertion in 12J_P1 was due to the integration of 12D_P2 into the ancestral primary chromosome. Hence, the ancestral R. pickettii likely contained two chromosomes and two plasmids and lacked a virulent copy of the filamentous phage p12J. Both R. pickettii 12D and 12J have multiple copies of Cu and Co-Zn-Cd resistant genes (Table 2.2). The majority of these copies are located either on a plasmid (12D_P2) or the insertion region that share high synteny to the plasmid (12J_C1). The phylogenetic trees of 62 czcA and copA genes showed that ancestral R. pickettii contained a large number of czcA and copA genes. Both Figure 2.7 and Figure 2.8 revealed that several groups of czcA (Group I, II, and VI) and copA (Group I-III) genes grouped together on the same branch of lineage. These co-existing groups suggest that these gene copies existed in the ancestral R. pickettii genome. Thus, the divergence of R. pickettii 12D and 12J appeared to be a recent event. Horizontal gene transfer (HGT) and gene duplication/amplification (GDA) played a substantial role in the R. pickettii 12D and 12J adaptation to the high copper environment. We observed a large number of horizontally transferred foreign genes in both strains (Figure 2.1 and Figure 2.2). R. pickettii 12D P1 is especially abundant with foreign genes from Eukaryotes (Figure 2.1 and Table A.1). Previous studies suggested that HGT and GDA have been frequently used by other bacterial strains (e.g. C. metallidurans and Staphylococcus sp.) to acquire or amplify metal and antibiotic resistance genes [21, 98, 104, 135, 146]. Figure 2.5 showed that strain 12J obtained metal resistant genes, such as mer operon and czc operon, from other β-Proteobacteria. The phylogenetic trees of czcA and copA genes revealed that many copies of czcA and copA genes share lineages with R. metallidurans CH34, R. eutropha JMP134, R. solanocerum Po82 and GMI1000 and γ-Proteobacteria (copA only). These close lineages to other bacteria further confirmed that many of these genes were obtained via HGT. We also observed copies of czcA genes (12D P2 copy 2, 12J C1 copy 4 and 5, and Group V) and copA genes (12J C1 copy 2) grouped together with other R. pickettii 12D and 12J czcA and copA genes (Figure 2.7 and Figure 2.8). This suggests that these copies of genes were duplicated to confer the metal stress. We observed that R. pickettii 12D and 12J contain fewer copies of Co-Zn-Cd resistant genes and more copies of Cu resistant genes compared to C. mtallidurans CH34, the model 63 organism to study the metal resistant determinants (Table 2.2) [67, 71, 105, 111, 160]. This is interesting because C. metallidurans CH34 was isolated from a Zn rich environment and R. pickettii 12D and 12J were isolated from a Cu rich environment [80, 171]. It is clearly shown here that C. metallidurans contains more Co-Zn-Cd resistant genes because it was isolated from a high Zn environment and R. pickettii 12D and 12J contains more Cu resistant genes because it was isolated from a high copper environment. This is consistent with the environmental conditions from which the strains were isolated and presumably reflecting the selection pressures they were subjected to. We observed that R. pickettii 12D and 12J contain a large number of cop gene copies and they are more abundant in 12J than 12D (Table 2.2). We found four copies of czcE, a copper binding protein encoding gene described by Zoropogui et al. in 12J and only three in 12D [195]. Together, these high copy numbered copper resistant and binding genes could elucidate the larger amount of copper sequestered by R. pickettii 12D and 12J [191]. The larger copy numbers of czcE genes in 12J also correlates with the larger amount of copper sequestered by 12J cells which was previously reported [191]. Passot et al. previously suggested that we are witnessing the emergence of the third chromosome in 12J [123]. However, our results indicate that 12J actively integrates genes rather than splitting chromosomes. Interestingly, we noticed that R. pickettii 12D_P1 is large in size and contains a large amount of horizontally transferred genes and IS elements. This plasmid is substantially different from the other plasmids in 12D and 12J. It contains kfr A and replication repA gene like the secondary chromosomes. Thus, it resembles a possible early stage tertiary chromosome. All of this adaptation and evolution evidence suggests that we are witnessing the co-evolution of two strains and the rising of diversity within the species. 64 The different evolutionary strategies R. pickettii 12D and 12J took for adaptation in a heavy metal environment might be the beginning of speciation. 2.6 Conclusion In this study, we compare the genomic variations between two R. pickettii strains to reveal the development of R. pickettii microdiversity in Torch Lake, Houghton, MI. Our results indicated that R. pickettii 12D and 12J acquired metal (eg. copper, cobalt-zinc-cadmium, mercury) resistant genes from other bacteria, such as R. solanocerum and C. metallodurans, via horizontal gene transfer. Under the pressure of high copper in the environment, both strains also increased the number of metal resistant gene copies via gene duplication/amplification. We found that a larger number of copper resistant gene copies (12J > 12D) corresponds to a greater amount of actively sequestered copper (12J > 12D). The comparative genomic analysis suggests that both strains co-evolved under the pressure of high copper. Gaining the resistance to filamentous phage in R. pickettii 12D might be the beginning of the divergence. We predict that the ancestral R. pickettii had one primary chromosome (similar to 12J_C1 minus the insertion from 12D_P2), a secondary chromosome (similar to 12J_C2 or 12D_C2), two plasmids (similar to 12D_P2 and 12D_P3) and a filamentous phage particle. After the divergence, strain 12J actively integrated mobile elements containing metal resistant genes and strain 12D maintained original plasmids and acquired another large plasmid (12D_P1), which was largely composed of horizontally transferred foreign genes. We suspect that 12D_P1 is undergoing the process of being transformed to a third chromosome. The differences in metal-resistant gene copies and genome scaffolds suggested the development 65 of R. pickettii microdiversity might be the first step in speciation. Our results also showed that secondary chromosomes of R. pickettii 12D and 12J appeared to be much more stable than the primary chromosomes during the evolution. This finding was to our surprise because it is the opposite of the observations of Cooper et al., which suggested that secondary chromosomes mutate and evolve faster than primary chromosomes [28]. We suspect that secondary chromosomes are preferably preserved in microdiversity but preferably modified across species. 66 CHAPTER 3 RESOURCE-INDUCED BACTERIAL STRUCTURE SHIFTS AND THEIR INFLUENCE ON COMMUNITY FUNCTIONS 3.1 Abstract Subsurface uranium contamination poses serious health risks and is costly to remove by traditional physical-chemical methods. A more cost-effective method has proven to be in situ immobilization of subsurface uranium via the microbially mediation. The full potential of this approach would be facilitated by the identification of all microbial species, including previously uncultured strains, that play a substantial role in uranium immobilization. However, in situ studies are influenced by uncontrolled environmental factors and soil bacterial communities are complex. Thus, we used standard enrichment protocols on bacterial communities from uranium contaminated sediments to select for strains capable of immobilizing uranium. Previous studies have indicated that many uranium reducing bacteria are also able to reduce nitrate and iron. Therefore, minimal media supplemented with lactate in combination with nitrate, ferric citrate, or uranyl acetate were used as the conditions of enrichment. Culture independent bacterial community analyses showed that bacterial communities shifted drastically upon enrichment and Firmicutes were observed in all enrichments. Genera Pyschrosinus and Clostridium were particularly abundant in uranyl acetate enrichments and were commonly detected in both uranium and iron enriched communities. Despite the drastic structural shifts, we observed a stable functional gene profile in all enrichment communities. Our findings indicated that phylum Firmicutes might play a substantial 67 role in uranium immobilization. The stable functional gene profiles targeting over 44,000 genes of the enriched bacterial communities suggested that the net function of the bacterial communities that resided in the uranium contaminated environments remained unchanged. 3.2 Introduction Uranium is a naturally occurring radioactive metal that can be detected in rocks, soil and streams. Due to weathering, oxidized uranium (U6+ ) is commonly found in the environment. Human beings are exposed to this background uranium at low levels via food and water (∼0.7-1.1 µg/day) without potential health threats. However, human activities such as mining and weapon development released elevated amounts of uranium in the environment. Among 1,517 former and current United States Environmental Protection Agency (USEPA) National Priority List (NPL) sites, at least 54 sites were identified as uranium contaminated [167, 168]. The uranium contaminants can permeate into ground water and become a health threat to human beings. The Oak Ridge National Laboratory (ORNL) integrated Field Research Center (iFRC) is an example of a site with extensive underground uranium contamination. The uranium containing wastes penetrated an unlined waste pond (S-3 area), contaminated the groundwater, and formed secondary contaminant sources with the soil sediments in subsurface. This large area subsurface contamination site poses a great challenge the development of cost-effective remediation methods. Lovley et al. [89, 88] discovered that some microorganisms were able to utilize oxidized metals (e.g., Fe3+ , Mn4+ ) as electron acceptors to generate energy under anaerobic conditions. It was also shown and proven by Gorby et al. that U6+ could also be used as the 68 electron acceptor during bacterial anaerobic respiration [52]. The reduced uranium, U4+ is less stable (can be readily oxidized to U6+ ) but insoluble in water [14, 43]. Thus, by decreasing the solubility of uranium (e.g., forming uraninite), uranium could be removed from ground water. Several studies have also shown that iron and sulfate reducing bacteria have the potential to immobilize uranium via enzymatic activities [97, 102, 148, 186]. Encouraged by these laboratory results, researches on in situ bioremediation have been performed at several Department of Energy (DOE) uranium contaminated research sites [16, 17, 64, 77, 187, 188, 190]. Wu et al. conducted in situ uranium immobilization studies at Oak Ridge National Laboratory (ORNL) integrated Field Research Challenge (iFRC) site S-3 area [187, 188]. Uranium reduction in this area is particularly challenging because of the inhibition caused by pore water with low pH and high concentrations of nitrate and aluminum [187]. To overcome the inhibition of uranium reduction, Wu et al. [187] designed a nested circulation system, where a pH neutral zone (the outer loop) was created and a biostimulation zone (the inner loop) was established within the pH neutral zone. Within the biostimulation zone, Wu et al. successfully demonstrated that U6+ was immobilized in its reduced form (U4+ ) [188]. Cardenas et al. investigated the bacterial community shifts accompanied the reduction of uranium in the biostimulation zone and suggested that sulfate reducing bacteria might play a substantial role in uranium reduction [16, 17]. In situ bacterial community shifts reflecting uranium reduction are particularly informative in the identification of novel uranium immobilizers. However, these populations are usually difficult to isolate and their activities cannot be directly confirmed. In addition, the in situ community dynamics are strongly influenced by the field environmental factors 69 [35]. For these reasons, a controlled system that leads to cultivation of potential uranium immobilizers is desirable. Herein, we report our studies on identifying the potential uranium immobilizing populations by enriching uranium exposed soil sediments in the laboratory. Soil sediments were collected from ORNL iFRC site, specifically within the biostimulation zone (FW102-2 and FW102-3) and pH neutral zone (FW107) of the nested circulation system. Sample FW102-2 and FW102-3 were collected from the same monitoring well but at two depths (13.7 m bgs and 12.2 bgs, respectively). Cardenas et al. [16] previously reported that soluble uranium concentrations decreased in monitoring well FW102. No bacterial community studies have been conducted in the pH neutral zone. All three samples contained a large amount of uranium (700 ppm in FW107, 1600 ppm in FW102-2, 300 ppm in FW102-3). Thus, we hypothesized that all three samples contain strains capable of immobilizing uranium. Besides uranium, the ORNL iFRC S-3 area where these samples were collected are also abundant in nitrate and iron [187]. It is known thermodynamically that anaerobic energy yield ranks as NO− >Fe3+ >U6+ . Research studies have also shown that uranium reducers are capable of 3 nitrate and iron reduction [2, 163]. Therefore, we enriched soil sediments in minimal media supplemented with NO− , or Fe3+ , or U6+ . By doing this we expected to identify specific 3 bacterial community shifts associated with uranium enriching processes. 70 3.3 3.3.1 Materials and Methods Anaerobic Media and Stocks Minimal growth media (FWM and FWM-Fe) were modified based on previously reported freshwater-medium and ferric-citrate media [87]. The detailed composition of FWM is listed in Table B.1, Table B.2 and Table B.3. FWM-Fe medium is the same as FWM but supplemented with 56 mM of ferric citrate. Supplement stocks, sodium lactate (2 M) and sodium nitrate (1 M) solutions were prepared by dissolving chemicals in deionized water to desired concentrations. Stocks solutions were filter (0.22 µm) sterilized. Uranyl acetate stock (100 mM) was prepared by dissolving uranyl acetate powder in 30 mM pH7 sodium bicarbonate solution. Filter sterilization was also applied. All media and stock solutions were bubbled under N2 :CO2 (80:20) to eliminate oxygen. 3.3.2 Bacterial Community Enrichments Three uranium contaminated soil samples (FW102-2, FW102-3, and FW107) were collected anaerobically from the nested recirculation experimental site at Oak Ridge iFRC S-3 area (denoted as "O" in sample labels). Bacterial consortia from these soil sediments (10%, v/v) were enriched in FWM [86] supplemented with sodium lactate (20 mM) combined with either sodium nitrate (10 mM, denoted as "NO3" in sample labels), or ferric citrate (56 mM, denoted as "Fe" in sample labels) or uranyl acetate (0.25 mM, denoted as "U" in sample labels) in duplicates at 15 ◦ C. One control culture without the addition of inocula was also set up for each enrichment condition. The soil sediments were enriched for a total of 35 days and another 20mM lactate was injected on the 17th day after sampling to provide additional 71 nutrients (after sampling). Two-ml samples were taken from ferric citrate enrichments and passed through 0.22 µm HPLC certified syringe filters (Sun SRI Titan2 17mm Syringe Filters, Thermo Fisher Scientific) on the 1st, 17th and 35th day of incubation. Samples were mixed with HCl to obtain a final concentration of 0.5 M HCl and stored at -20◦ C until being processed. Half of the sample was used to monitor the depletion of lactate and half for the reduction of Fe3+ . The enrichments were then used as inocula (1% v/v) for secondary enrichments under the same conditions. Additional sodium lactate was added on the 13th day. The ferric citrate enrichments were again monitored for 35 days. Bacterial consortia samples were collected at the end of second enriching period (70th day) for 16S rRNA gene and community functional gene profiling. From each enrichment, two 1 ml samples were withdrawn and centrifuged at 10,000 RPM for 20 minutes in a Sorvall SS-34 rotor and the pellet was stored at -20◦ C until extraction of community DNA. The remaining material was processed using GeoChip [57]. 3.3.3 Quantitative Analyses The concentration of sodium lactate was determined using high-performance liquid chromatography (HPLC) with an Aminex HPX-87H column (polystyrene-divinylbenzene ion exchange resins, Bio-Rad, Hercules, CA). Sodium lactate and acetate mixtures were used as standards (0.04 µM - 0.4 µM) and freshly prepared in triplicate prior to sample analyses. A mixture of acetylnitrile and sulfuric acid (5%) was used as the eluent. Iron reduction was determined by the previously published ferrozine essay [154]. Briefly, 0.1 ml of filtered acidified ferric citrate sample was added to 0.9 ml 1 M HCl. This mixture (0.1 ml) was then added to 1 ml ferrozine buffer (a mixture of 50 mM HEPES and 0.1% 72 ferrozine with a final pH of 7) and incubated for 10 minutes at room temperature. After incubation, 200 µl was added to a clear flat-bottom Corning 96-well microtiter plate (Corning Scientific, Tewksbury, MA) and absorbance was measured at 520 nm. 3.3.4 DNA Extraction and Community Profiling Community DNA was extracted by using Powersoil DNA Extraction Kit (Mo Bio Laboratory, Carlsbad, CA). Two pyrosequencing chemistries (454 FLX and 454 Titanium) were used to evaluate the bacterial 16S community structure. For samples in replicate 1, bacterial 16S rRNA gene V4-V5 regions (Forward 5’AYTGGGYDTAAAGNG3’ and Reverse 5’TACNVGGGTATCTAATCC3’) were amplified as previously described [157]. Samples from replicate 2 were amplified with primer sets developed by Human Microbiome Project (HMP) targeting the V3-V5 region (357F 5’CCTACGGGAGGCAGCAG3’ and 926R: 5’CCGTCAATTCMTTTRAGT3’) of 16S rRNA genes. Both polymerase Chain Reactions (PCR) were carried out by mixing 1X Taq reaction buffer, 2 mM MgSO4 , 0.1 mg/ml BSA, 0.2 mM of each dNTP, 0.33 µM of each primer, 0.125 U/µl of Platinum High Fidelity Taq Polymerase (Invitrogen, Carlsbad, CA), 10 ng DNA template and DNase and RNase free water. The amplification conditions include an initial denaturing step at 95◦ C for 5 minutes, followed by 30 cycles of 3 temperature step consisted of denaturing at 95◦ C for 45 seconds, annealing at 50◦ C for 45 seconds, and elongation at 72◦ C for 90 seconds, and a final extension at 72◦ C for 5 minutes. All amplified 16S rRNA gene PCR products were visualized on 2% Tris-Borate-EDTA (TBE) agarose gel. DNA fragments at ∼450 bp (V4-V5) and ∼569 bp (V3-V5) were excised and extracted from the gels by using Qiagen Gel Extraction Kit and further purified with QiaQuick PCR Product Purification Kit (Qiagen, Valencia, CA). All 73 purified DNA products were quantitated with Quibit HS Double-stranded DNA Kit (Invitrogen, Carlsbad, CA) then composited at equal mass for a final DNA concentration of 0.5 ng/µl. Pyrosequencing was performed at Michigan State University RTSF by using a Roche 454 GSFLX and 454 GSFLX Titanium Sequencers. Functional gene profiling of the enriched consortia was performed at Oklahoma University. Community DNA were extracted as previously described [172]. GeoChip 3.0 was used to access the functional gene profiles of U6+ and NO− enriched samples in replicate 2. GeoChip 3 5.0 was used to access the functional gene profiles of the original soil sediment communities and the Fe3+ , U6+ and NO− enriched communities in replicate 1 and Fe3+ and NO− enriched 3 3 communities in replicate 2. GeoChip 3.0 and GeoChip 5.0 data were combined and GeoChip 3.0 replicate 2 NO− enrichment data were omitted for analysis purpose. 3 3.3.5 Community Analysis Bacterial 16S rRNA gene sequences obtained from pyrosequencing were processed and analyzed using the Ribosomal Database Project (RDP) Pyrosequencing pipeline [180]. Sequences that were shorter than 250 bp, containing any N’s in the sequencing region, or with an average quality score less than 20 were discarded. Chimeras were removed from RDP processed sequences by using USEARCH (UCHIME reference mode with Silva gold alignment database as reference). RDP complete linkage clustering was used to determine bacterial operational taxonomic units (OTUs) based on 97% sequence similarity. To eliminate the clustering artifacts, sequences were trimmed to the same length (144 bp model position) based on RDP alignment profiles prior to clustering. RDP Multi-classifier was used to identify each sequence to bacterial genus level (threshold 80). Any identified genus that has a 74 total count of 1 across all samples (21 samples) was omitted from analyses. RDP SeqMatch was used for further identification of bacterial clusters of interest. OTU- and phylotype-based community analyses were performed by using statistical software R [128]. Ecological indices were estimated for each bioreactor sample based on clustered sequences by using R package Vegan [117]. Specifically, community richness (ChaoI), diversity indices (Shannon H’) and community evenness (E) were calculated based on equations previously described [94]. Diversity index Shannon H’ was expressed as magnified H’ (eH ) to emphasize the differences among samples [94]. The sampling coverage (C) of each sample was calculated based on Good’s method, C=1-n1 /N, where n1 represents number of OTUs that was observed once and N is the total number of sequences in the sample [39]. R package Vegan was also used to perform non-metric multidimensional scaling analysis (NMDS) to correlate the dissimilarities among bioreactor samples and the shifts in phylotype abundance [117]. To identify the shared genera among different enrichments, genera found in both duplicated enrichments were extrapolated first. Venn diagrams of these dereplicated genera were then constructed by using the R package Vennerable [159]. 3.4 3.4.1 Results Monitoring the Enrichment Cultures Three soil sediments were collected from locations containing 700 ppm (FW107), 1600 ppm (FW102-2) and 300 ppm (FW102-3) uranium. Among three electron acceptor supplemented enrichments, we observed growth in nitrate supplemented enrichments as early as the 5th day. All enrichments grew equally well. Between the 15th and 30th day, we observed Fe3+ 75 reduction in ferric citrate enrichments, where the culture turned dark brown. The change of colors were first observed in FW102-2 and FW102-3 enrichments. FW107 enrichments changed color by day 30. A large amount of precipitates were formed upon the addition of uranyl acetate. Hence, it was difficult to differentiate the growth of the uranium enrichment cultures from the turbidity due to precipitation. However, we did observe abundant live cells under the microscope throughout the enrichment process. The monitored ferric citrate enrichments revealed that iron was actively reduced to Fe2+ while sodium lactate was used as the electron donor and carbon source (Figure 3.1). No growth was observed in the control microcosms and we did not detect Fe2+ (above background level) in these control cultures. Under the same enrichment condition, we observed that the FW-107 culture reduced the least amount of Fe3+ (Figure 3.1 A and B) and accumulated the most amount of acetate (Figure 3.1 C and D). Some acetate accumulated in FW102-2 and FW102-3 ferric citrate enrichments by the 13th day of incubation. However, this acetate was consumed by the end of the incubation period (Figure 3.1 C and D). Second generation enrichments were more efficient at iron reduction and lactate utilization than the first generation. 3.4.2 Overview of the Community Dynamics Bacterial communities of the original iFRC soil sediments and the enrichment cultures were investigated based on bacterial 16S rRNA sequences. We obtained 703 to 8213 sequences per sample. Although two different sequencing chemistry were used, the ecological indices indicated that the data obtain between two replicates are comparable (Table 3.1). Pyrosequencing was able to cover at least 97% of the sample community, except for FW1022_O, 76 the original soil sediment sample from well FW102-2, which also had the least number of sequences (Table 3.1 column C). All of the original soil sediments appeared to be more even and diverse than the enriched samples, as we expected. We also found that the original community in FW107 contained only 75 OTUs, considerably fewer than were found in FW102-2 and FW102-3. ChaoI, Magnified Simpson (eH ) and evenness (E) indices also showed that the original community of FW107 was much less diverse and more skewed than the other two. Uranyl acetate and ferric citrate enrichments were the most skewed and least diverse samples in both FW102 groups and FW107 group, respectively (Table 3.1). 3.4.3 Shifts in the Enrichment Bacterial Community Structures A detailed comparison of the bacterial community structures showed substantial shifts among enrichments (Figure 3.2). The ten most abundant genera summed from all 21 samples and their distributions in each community were compared side-by-side. Consistent with the ecological indices, we observed that community structures between replicates were similar, however substantial differences were observed among enrichments. FW107 group appeared to be different from the others. FW107 original sample and enrichment samples were highly dominated by one or two genera. The original soil sediment sample collected from FW107 was abundant with Sulfuricurvum, which was not found in other samples. Over 80% of sequences from the FW107 ferric citrate enrichments were unclassified Veillonellaceae. Psychrosinus were found dominant in FW107 uranyl acetate enrichment cultures. FW102-2 and FW102-3 bacterial consortia enriched under the same condition appeared to be similar. While all ten genera constituted less than 10% of the populations in FW102-2 and FW102-3 soil sediment samples, Sulfuricurvum comprised of over 70% of the FW107 original community 77 Table 3.1: Ecological indices of iFRC enrichment bacterial communities. O represents original soil sediment samples. Fe, NO3, and U supplemented enrichment cultures. Replicate Total Sequences OTU ChaoI eH E C (%) FW1022_O 0 703 146 286.05 53.33 0.80 88.76% FW1022_Fe 1 2 6623 5891 57 72 91.20 107.00 13.48 0.64 14.86 0.63 FW1022_NO3 1 2 2293 8213 67 65 92.00 89.00 6.49 9.14 0.44 0.53 98.91 99.81 FW1022_U 1 2 1125 5152 16 44 19.75 49.14 1.95 3.81 0.24 0.35 99.47 99.83 FW1023_O 0 6078 203 309.24 26.87 0.62 98.70 FW1023_Fe 1 2 3700 4315 60 50 77.10 71.38 7.68 7.08 0.50 0.50 99.49 99.56 FW1023_NO3 1 2 878 5488 31 70 37.43 101.63 5.76 0.51 9.31 0.53 98.86 99.58 FW1023_U 1 2 6008 6055 53 49 78.67 62.00 4.09 5.38 0.35 0.43 99.63 99.79 FW107_O 0 1410 75 117.00 4.04 0.32 97.45 FW107_Fe 1 2 1664 2956 17 15 22.60 29.00 2.17 1.21 0.27 0.07 99.52 99.73 FW107_NO3 2 1 3596 4102 56 62 123.67 76.25 5.05 0.40 7.64 0.49 99.19 99.54 FW107_U 1 2 771 4679 32 46 36.67 50.20 3.87 0.39 3.40 0.32 98.96 99.85 99.71 99.64 sequences. A large amount of Geobacter were detected in the ferric citrate and uranyl acetate enrichments. The proportion of these ten genera increased as the enrichment stress went from nitrate to iron to uranium (Figure 3.2). We also observed that Psychrosinus was found in all samples (except FW107 ferric citrate enrichments) and its abundance increased as the enrichment stress increased (Figure 3.2). 78 A. FW107 FW102-2 FW102-3 15 10 5 0 B. Day 17 Lactate Day 35 Acetate 15 FW107 10 FW102-2 FW102-3 Day 1 Day 1 25 D. 20 Lactate (mM) Fe(II) (µM) 45 40 35 30 25 20 15 10 5 0 20 Day 35 1st Enrichment C. 5 0 Day 17 Day 35 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 Acetate (mM) Day 17 Day 1 Acetate Acetate (mM) 25 Lactate (mM) Fe(II) (µM) 45 40 35 30 25 20 15 10 5 0 Day 1 Day 17 Lactate Day 35 FW107 FW102-2 FW102-3 FW107_lac_control FW102-3_lac_control FW102-2_lac_control 2nd Enrichment Figure 3.1: The concentration of Fe2+ (panels A and B), lactate and acetate (panels C and D) in each ferric citrate enrichment cultures. The first generation enrichments are presented in panel A and C and the second generation enrichments are presented in panel B and D. Error bars represent the differences between two replicates. 79 FW FW 10 102 FW 22 2_ 10 _NO O   2 3 FW 2_N _2   10 O3 FW 22 _1   10 _Fe _ FW 22_ 2   10 Fe_ FW 22 1   10 _U_ 22 2   FW _U FW _ 10 102 1   FW 23 3_ 10 _NO O   2 3 FW 3_N _2   10 O3 FW 23 _1   10 _Fe _ FW 23_ 2   10 Fe_ FW 23 1   10 _U_ 23 2   _U FW FW _1 10 107   FW 7_ _O 10 NO   7_ 3_2 FW NO   10 3_ FW 7_ 1   10 Fe_ 2 FW 7_F   10 e_1 FW 7_   10 U_2 7_   U_ 1   Abundance  (%)   100   90   80   70   60   50   40   30   20   10   0   Psychrosinus   Simplicispira   unclassified_Clostridiales   Geobacter   Pseudomonas   Clostridium  XlVb   unclassified_Veillonellaceae   Sulfuricurvum   unclassified_Oxalobacteraceae   Figure 3.2: The top ten (summed across all samples) most abundant genera identified via RDP multiclassifier. Sample replicates were labeled as "_1" and "_2". Original soil sediment samples were denoted with "_O." 80 FW107_Fe FW107_U 5 1 4 6 27 0 1 0 FW107_NO3 0 0 0 1 4 2 0 FW107_O Figure 3.3: The shared genera among FW107 samples. The color of sample font is the same as the cognate ellipse. 3.4.4 Shared Genera We identify the bacterial genera that were shared among enrichments. Venn diagrams were constructed on essential genera that were shared between the replicated cultures of FW107, FW102-2 and FW102-3 samples. Among 51 genera identified in the original FW107 sediment and enriched consortia, over 50% were unique to FW107 and only one (Sulfurospirillum) was shared by all four samples (Figure 3.3). The unique genus in FW107 iron enrichment was Trichococcus. FW107 uranium and nitrate enrichment cultures shared 5 genera, out of which, one was also detected in iron enrichment consortia and the other four were Psychrosinus, Pseudomonas and two groups of Clostridium. Five genera, unclassified Ruminococcaceae, Oscillibacter, Clostridium sensu stricto, Clostridium IV, and Acidovorax were found only in uranium enriched FW107 cultures. Unclassified Veillonellaceae was found in uranium, iron 81 FW1022_Fe FW1022_U 0 10 0 1 47 3 4 3 FW1022_NO3 2 1 1 0 4 4 0 FW1022_O Figure 3.4: The shared genera among FW102-2 samples. The color of sample font is the same as the cognate ellipse. and nitrate enriched consortia. A large number of unique genera (47) were found in FW102-2 original soil sediment sample (Figure 3.4). Uranyl acetate enrichment of FW102-2 appeared to be a subset of FW102-2 ferric citrate enrichment. The four genera shared by ferric citrate, nitrate and uranium enrichments were Sulfurospirillum, Sporomusa, Psychrosinus, and Geobacter. FW102-3 samples contained a total number of 107 bacterial genera, 70 of which were unique in the FW102-3 original soil sediment (Figure 3.5). Psychrosinus was found common in iron, nitrate and uranium enrichment cultures. Geobacter and unclassified Veillonellaceae were shared by all four samples (Figure 3.5). Unclassified Ruminococcaceae, unclassified Bacteroidetes, Sporomusa, Desulfosporomusa, Clostridium III were only found in uranyl acetate enriched FW102-3 consortia. Among FW107, FW102-2 and FW102-3 uranium enrichments, 82 FW1023_Fe FW1023_U 5 13 2 0 70 2 1 0 FW1023_NO3 5 2 2 0 0 5 0 FW1023_O Figure 3.5: The shared genera among FW102-3 samples. The color of sample font is the same as the cognate ellipse. Sporomusa and Psychrosinus were the two numerically dominant taxa based on sequence frequency. 3.4.5 Functional Gene Profiles GeoChip profiles revealed that there were 1532 functional genes (between different GeoChip platforms) found in different microorganisms. Four uranium enriched samples, FW107_U(2), FW107_U, FW1022_U(2), and FW1023_U(2) appeared to be substantially different from the others (Figure 3.6). These four samples had many fewer abundant functional genes suggesting that uranium selects for a smaller collection of taxa. Three out of four samples were analyzed by GeoChip3. Although both FW107 uranium enrichment duplicates clustered close to each other, uranium enrichment duplicate 2 of FW102-2 and FW102-3 appeared 83 to be different from duplicate 1 (Figure 3.6). Analysis on GeoChip5.0 enrichment culture replicate 1 alone showed that the overall variation among all samples was small (Figure B.1). Thus, we suspect the different GeoChip platforms contributed substantially to the differences among samples we observed. 84 Color Key 024 FW1022_NO3(2) FW1023_NO3(2) FW1023_NO3 FW1023_Fe(2) FW1022_Fe(2) FW107_NO3 FW1022_NO3 FW107_NO3(2) FW107_O FW1023_U FW1022_U FW1023_Fe FW107_Fe FW1023_O FW1022_O FW107_Fe(2) FW1022_Fe FW1023_U(2) FW1022_U(2) FW107_U FW107_U(2) Row Z-Score Figure 3.6: The functional gene profiles of the original soil sediment and enrichment samples. The functional gene signals of all samples (columns) were standardized to the same amount. Each functional gene (row) was then scaled among all 21 samples to show the differences in abundance. Green indicates high abundance and red indicates low abundance. 85 3.4.6 Non-metric multidimensional scaling (NMDS) analysis Non-metric multidimensional scaling (NMDS) analysis was performed to correlate the βdiversity of enrichment samples and bacterial community or functional gene shifts. The original soil sediment sample community compositions appeared to be substantially different from the enriched communities (Figure 3.7). This difference also influenced the bacterial genera and functional gene correlations substantially (Figure 3.7 and Figure 3.8). Both Figure 3.7 and Figure 3.8 showed that the most significant shifts were observed between original soil sediment samples and the enrichments. Significant decreases in bacterial genera, including many unclassified bacteria were observed in the enriched cultures (Figure 3.7). The functional gene profile revealed that genes involved in the sulfur cycle (e.g., sox and dsr A) and lincosamide resistance (linB) decreased significantly in enriched cultures. Within the enrichments we also noticed some significant phylogenetic and functional shifts among bacterial communities due to the enriching conditions (Figure 3.9 and Figure 3.10). In the nitrate enriched communities, unclassified bacteria (β-Proteobacteria, Chitinophagaceae, and Comamonadaceae), Dechloromonas, and Simplicispira were found significantly more abundant than the other enrichments. We observed that unclassified Enterobacteriaceae, Clostridium XIVa, and unclassified Clostridiales were significantly more abundant in FW102-2 and FW102-3 Fe and U enrichment cultures. Unclassified Veillonellaceae, Sulfurospirillum and Trichococcus were significantly more abundant in FW107 Fe and U enrichments. The abundance of Geobacter shifted significantly towards FW102-2 and FW102-3 U enrichments (Figure 3.9). The abundance of a large number of functional genes shifted substantially among all enrichments (P ≤ 0.01, data not shown). We observed 86 the abundance of three groups changed most significantly (P ≤ 0.001). Group III includes seven different bacterial functional genes, which appeared to be significantly more abundant in FW107 Fe enrichments. The abundance of chromate transporter encoding gene (chr A, specifically found in Burkholderia xenovorans LB400) decreased significantly from FW107 enrichments to FW102-2 and FW102-3 enrichments (Figure 3.10). 87 0.6 I unclassified Deltaproteobacteria 0.4 FW107_NO3(2) 0.2 FW107_U 0.0 -0.2 FW1022_NO3 FW1023_O FW107_NO3 FW1023_NO3 FW107_Fe FW1023_Fe(2) I FW1023_U II FW1023_Fe FW1023_U(2) FW1022_Fe(2) FW107_Fe(2) FW1022_O III FW1022_Fe FW1022_U FW107_O FW1022_U(2) unclassified Cytophagaceae Subdivision3 genera incertae sedis unclassified Burkholderiaceae unclassified Syntrophobacteraceae unclassified Desulfuromonadales unclassified Cystobacterineae unclassified Cystobacteraceae unclassified Opitutaceae Gp6 Campylobacter Arcobacter Prevotella Desulfobulbus Roseomonas III -0.6 -0.4 II FW1023_NO3(2) FW107_U(2) NMDS2 FW1022_NO3(2) unclassified Desulfobulbaceae -0.5 0.0 0.5 NMDS1 Figure 3.7: NMDS analysis of NO− enrichments (grey), Fe3+ enrichments (pink), U6+ enrichments (yellow), 3 and original samples (green). Each ellipse was drawn around each group centroid based on a group distance standard error at 95% confidence level. The changes in genera abundances that drive comparative community differences are indicated with red arrows. The arrows point to the areas of high abundance. The red arrows indicate the abundance of genera that changed significantly (P ≤ 0.001). The specific genera are shown in red according to the group labeling. 88 0.6 0.0 -0.2 II 2) FW1023_Fe( FW1023_U FW1023_Fe FW1023_U(2) FW1022_U II III FW1023_O FW1022_O ] IV FW1022_Fe FW1022_Fe(2) -0.4 sox Silicibacter pomeroyi DSS-3 _N O3 FW107_U FW107_NO3 FW1023_NO3 FW107_Fe I ohbAB Paracoccus denitrificans PD1222 III adpB Kineococcus radiotolerans SRS30216 IV gyrB Legionella pneumophila subsp. fraseri linB Methylobacterium populi BJ001 FW107_O FW107_Fe(2) dsrA uncultured sulfate-reducing bacterium nbaC Sclerotinia sclerotiorum 1980 linB Bradyrhizobium sp. BTAi1 FW1022_U(2) -0.6 NMDS2 0.2 FW1023_NO3(2) FW107_U(2) I FW 10 22 0.4 FW107_NO3(2) FW1022_NO3(2) -0.5 0.0 0.5 NMDS1 − enrichments NO3 Figure 3.8: NMDS analysis of (grey), Fe3+ enrichments (pink), U6+ enrichments (yellow), and original samples (green). Each ellipse was drawn around each group centroid based on group distance standard error at 95% confidence level. The changes in functional gene (with the specific microorganism from which the probes were designed) abundance were correlated to the shifts observed in community structures. The red arrows indicate the abundance of functional genes that changed significantly (P ≤ 0.001). The specific genera are shown in red according to the group labeling. The arrows point to the areas of high abundance. 89 FW107_NO3(2) IV Simplicispira 0.5 unclassified Comamonadaceae ] FW1023_NO3(2) 0.0 FW107_NO3 FW107_U FW1 023_ Fe FW 10 23 I _F e( 2) V VI e _F 22 10 FW _U 23 10 FW FW1022_U -0.5 NMDS2 FW1023_NO3 FW107_Fe FW1023_U(2) FW107_Fe(2) III FW1022_U(2) -1.0 -0.5 Geobacter II FW1022_Fe(2) 0.0 unclassified Enterobacteriaceae II Clostridium XlVa FW1022_NO3 FW107_U(2) VII I FW1022_NO3(2) III unclassified Clostridiales IV unclassified Betaproteobacteria unclassified Chitinophagaceae Dechloromonas V unclassified Veillonellaceae VI Sulfurospirillum VII Trichococcus 0.5 NMDS1 Figure 3.9: NMDS analysis of NO− enrichments (grey), Fe3+ enrichments (pink), and U6+ enrichments (yellow). 3 Each ellipse was drawn around each group centroid based on group distance standard error at 95% confidence level. The changes in genera abundance were correlated to the shifts observed in community structures. The colored arrows indicate the abundance of genera that changed significantly (red: P ≤ 0.001, blue: P ≤ 0.05, and green: P ≤ 0.01). The genera names are shown in the same color according to the group labeling. The arrows point to the areas of high abundance. 90 FW107_NO3(2) FW1022_NO3(2) I FW1023_NO3(2) 0.5 II I phtA Comamonas testosteroni FW1022_NO3 FW107_U(2) FW107_NO3 FW107_U 0.0 FW107_Fe III FW1023_U FW1023_Fe(2) FW1023_Fe FW1023_U(2) FW107_Fe(2) FW1022_U -0.5 NMDS2 FW1023_NO3 FW1022_U(2) -1.0 -0.5 II ChrA Burkholderia xenovorans LB400 III gyrB Legionella jamestowniensis adpB Geobacillus thermodenitrificans NG80-2 tfdA Aspergillus fumigatus Af293 catB Planctomyces maris DSM 8797 GCoADH Sclerotinia sclerotiorum 1980 merP Idiomarina loihiensis L2TR CadA Sinorhizobium meliloti 1021 FW1022_Fe FW1022_Fe(2) 0.0 0.5 NMDS1 − enrichments NO3 Figure 3.10: NMDS analysis of (grey), Fe3+ enrichments (pink), and U6+ enrichments (yellow). Each ellipse was drawn around each group centroid based on group distance standard error at 95% confidence level. The changes in functional gene abundance were correlated to the shifts observed in community structures. The red arrows indicate the abundance of functional genes that changed significantly (P ≤ 0.001). The functional genes and the specific microorganisms from which the probes were designed are shown in the red according to the group labeling on the side. The arrows point to the areas of high abundance. 91 3.5 Discussion The identification of uncultured uranium immobilizers in sub-surface environments is instrumental in the optimization of in situ bioremediation. Our assessment of bacterial communities indicated, not surprisingly, that enrichment was an effective method for identifying bacterial populations that might play important roles in sub-surface uranium immobilization. We were able to reduce the bacterial diversity and identify common populations under different stressors (ie. iron, nitrate, and uranium) to evaluate their potentials. From the ferric citrate enrichment cultures, we observed that FW107 enrichments were not as effective at iron reduction as the FW102-2 and FW102-3 enrichments. Bacterial community analysis revealed that FW107 iron enrichments were predominantly populated by unclassified Veillonellaceae (Figure 3.2). Gihring et al. observed that although this group was abundant in their enrichment cultures, Veillonellaceae did not contribute to uranium reduction but to initial nutrient degradation in their study [50]. This might explain why FW107 ferric citrate enrichments appeared to be less effective at iron reduction and the accumulation of acetate than FW102-2 and FW102-3 ferric citrate enrichments (Figure 3.1). Thus we suspect that Veillonellaceae did not contribute to uranium immobilization directly but was responsible for early nutrient consumption. The other two genera, Trichococcus and Sulfurospirillum, observed in FW107 iron enrichments were reported to be associated with sulfate and metal reducers [41, 152]. We suspect the iron reduction observed in FW017 ferric citrate enrichments were mainly contributed by members of Trichococcus and Sulfurospirillum. In contrast to the FW107 iron enrichment culture, FW102-2 and FW102-3 ferric citrate enrichments appeared to be much more diverse. Genus Geobacter, well-known for 92 its metal-reducing and acetate utilizing members [132], was found abundant in both sets of consortia. The genus is likely a major contributor to iron reduction and acetate utilization in FW102-2 and FW102-3 iron enrichments. The shifts in the bacterial communities revealed that FW102-2 and FW102-3 enrichments became abundant with Proteobacteriia while Firmacutes became numerically dominant in FW107 enrichments (Figure 3.2). Venn diagrams revealed that uranium enrichments shared genera with both iron and nitrate enrichments (Figure 3.3, 3.4, and 3.5). Among all of the shared genera, Psychrosinus and Sporomusa, genera of phylum Firmicutes, were found in all uranium enrichment cultures. A member of genus Psychrosinus was previously isolated from near freezing temperature and identified as a lactate fermenting bacterium [137]. No evidence has been shown that members of Psychrosinus are sulfate or metal reducers. However, the Psychrosinus group was previously identified as Pelosinus according to RDP (spring 2012 update). We found the 16S rRNA sequences from our Psychrosinus genus were highly similar to Pelosinus sp. UFO1, which was reported as a uranium reducer by Ray et al. [129]. Genus Sporomusa has been reported to be found in various U6+ -reducing and sulfate-reducing environments [7, 12, 113]. However, no specific members of this group has been identified as a metal reducer. The top ten most abundant genera in all samples revealed that the FW107 original soil sediment sample was drastically different from FW102-2 and FW102-3 (Figure 3.2). This was expected as well FW107 was in the neutralized zone and was never stimulated [187, 188]. The most significant influence was the enrichment process. The NMDS analyses revealed that the most significant shifts were decreases in genera abundance (Figure 3.7 and functional gene abundance (Figure 3.8) from original soil sediment samples to enrichment cultures. Analysis 93 of bacterial community shifts induced by the addition of NO− , Fe3+ and U6+ revealed that 3 bacteria belonging to phylum Firmicutes were significantly more abundant in Fe3+ and U6+ consortia (Figure 3.9). The shifts in bacterial functional genes are less conclusive. Figure 3.10 suggests that the original soil sediment samples had a substantial impact on the functional profile variations. We suspect this result was also due to different GeoChip platforms. When eliminating GeoChip3.0 samples, we were surprised to observe that despite the drastic shifts revealed by 16S rRNA profiles, little variation weas found in their functional gene profile (Figure B.1). This observation suggested that the underlying functional profile of these communities were resilient and highly similar. However, it is also possible that important genes were not included on the GeoChip microarray. Thus, a different and more effective method might be needed to investigate enrichment community function shifts. 3.6 Conclusion We studied the shifts in bacterial communities, obtained from uranium contaminated soil sediments (i.e., FW107, FW102-2, FW102-3) and bacterial community shifts induced by supplemented resources (i.e., NO− , Fe3+ , and U6+ ). The enriched community structures 3 changed drastically according to bacterial 16S rRNA sequences. We observed that Geobacter was abundant in the FW102-2 and FW102-3 U6+ enrichment cultures, while Psychrosinus was abundant in the FW107 U6+ enrichment cultures. Psychrosinus was also found common in all enrichment cultures at various levels of abundance. The functional gene profiles of the enrichments and original soil samples appeared to be highly similar despite the large changes we observed in 16S rRNA based community structure analysis. Our observations suggest that 94 the function potential of these bacterial community requires did not change as much as its corresponding phylogeny structure. According to this study, we hypothesize that Firmicutes play an important role in subsurface uranium immobilization. 95 CHAPTER 4 THE ROLE OF FIRMICUTES IN URANIUM IMMOBILIZATION 4.1 Abstract The observation of uranium immobilization by anaerobic bacterial communities is a composite result of bacterial interactions. To understand the bacterial interactions and their impact on uranium immobilization, it is not only important to study the bacterial community structural changes but it is also critical to study individual isolated members of the community. In uranyl acetate supplemented enrichment cultures seeded with microbial communities from uranium contaminated sediments, we observed that the concentration of soluble uranium decreased and the abundance of Firmicutes bacteria increased as a function of time. We hypothesized that members of Firmicutes play an important role in subsurface uranium immobilization. We employed standard cultivation techniques under strictly anaerobic conditions and isolated 51 bacterial strains from the uranium enrichment cultures. Bacterial 16S rRNA gene sequences revealed that these isolates belong to four distinct phylogenetic clades of Firmicutes. However, comparative analysis of growth rates and genome structure using REP-BOX-GTG5 -PCR profiles indicated that these isolates are distinct strains and might react differently in the presence of soluble uranium. We screened ten isolates (based on their 16S rRNA identifications and REP-BOX-GTG5 -PCR profiles) for their abilities to immobilize uranium and six isolates were more proficient at removing soluble uranium than the original enrichment community. 96 4.2 Introduction Bacterial phylum Firmicutes are comprised of a group of ancient and metabolically versatile anaerobic bacteria [38, 96, 99]. They were frequently found in various environments and reported in bacterial enrichments. Many members of this group have also been identified as sulfate and metal reducing bacteria using carbon sources as electron donors [3, 32, 45, 58, 60, 118]. Members of Firmicutes, such as Clostridia are also well-known for their ability to produce hydrogen via [FeFe]-hydrogenase (HydA) [10, 103]. In resource scarce subsurface environments, metabolic hydrogen could be readily used as an electron donor to support bacterial growth and dissimulatory reduction (eg. Fe, Mn, and etc.) [79, 151, 194]. We have found large numbers of Firmicutes 16S rRNA gene sequences in iFRC (integrated Field Research Challenge site at Oak Ridge National Laboratory, TN) original soil sediment samples as well as our sequentially enriched microcosms (see Chapter 3). In the previous chapter, we were able to identify potential genera that might play important roles in subsurface uranium immobilization. In this chapter we report on the isolation and characterization of strains capable of immobilizing U6+ . We monitored the decrease of soluble U6+ , the corresponding changes in the U6+ enriched bacterial communities, and isolate potential U6+ immobilizers. We hypothesize that members of Firmicutes play a substantial role in U6+ removal in iFRC. 97 4.3 4.3.1 Materials and Methods Uranium Immobilization by FW107 and FW102-2 Enrichment Cultures Enrichment cultures with uranyl acetate as the terminal electron acceptor (Chapter 3) were used as a starting culture. To obtain an initial high cell density, they were first enriched anaerobically in 1/2X R2B medium at 15◦ C for 10 days. The 1/2X R2B enrichments were then centrifuged at 5000 RPM to collect all of the cells. Cell pellets were washed twice in 30 mM anaerobic sodium bicarbonate buffer (pH 7. cells were collected via centrifuging in between) inside of an aerobic chamber (2.5% H2 ). Final washed cell pellets were resuspended in anaerobic PIPES buffered artificial ground water media (PBAGW) supplemented with 10 mM sodium pyruvate [69] to a final OD600 of 0.01. Microcosms were set up in duplicate and incubated at 15◦ C. Over a period of 64 days, the concentration of soluble U6+ was determined by a spectrophotometric U6+ -azide assay [145]. All samples were taken from microcosms anaerobically. Samples for HPLC and U6+ analysis were passed through 0.22 µm HPLC certified syringe filters (Sun Titan2, Fisher Scientific) under anaerobic condition. Community analysis samples were also taken from one set of the microcosms (replicate 1) on days 0 (1/2X R2B cultures), 10, 20 and 64 of incubation with uranyl acetate. 4.3.2 Quantitative Analyses A uranyl-azide assay was modified from a previously described method [145]. Briefly, HNO3 was added to filtered samples to a final concentration of 4.9%. Acidified samples (40 µl) were then added to sodium azide solution (3M, 160 µl) in a clear bottom 96-well microtiter plate (Corning Scientific, Tewksbury, MA). Absorbance measurements were performed at 340 nm. 98 To correlate the absorbance and U6+ concentration, a standard curve was constructed (in duplicate) for each measurement (Figure C.1). This assay was found to be valid for soluble U6+ ranging from 5 µM to 1000 µM. 4.3.3 DNA Extraction and Sequecing Preparation Community DNA was extracted by using Powersoil DNA Extraction Kit (Mo Bio Laboratory, Carlsbad, CA). Pyrosequencing chemistries (454 FLX) were used to evaluate the bacterial 16S community structures. Bacterial 16S rRNA gene V4-V5 regions (Forward 5’AYTGGGYDTAAAGNG3’ and Reverse 5’TACNVGGGTATCTAATCC3’) were amplified as previously described with 454 FLX chemistry [157]. Polymerase Chain Reactions (PCR) were carried out by mixing 1X Taq reaction buffer, 2 mM MgSO4 , 0.1 mg/ml BSA, 0.2 mM of each dNTP, 0.33 µM of each primer, 0.125 U/µl of Platinum High Fidelity Taq Polymerase (Invitrogen, Carlsbad, CA), 10 ng DNA template and DNase and RNase free water. The amplification conditions include an initial denaturing step at 95◦ C for 5 minutes, followed by 30 cycles of 3 temperature steps consisting of denaturing at 95◦ C for 45 seconds, annealing at 50◦ C for 45 seconds, and an elongation at 72◦ C for 90 seconds, and a final extension at 72◦ C for 5 minutes following the completion of 30 cycles. Amplified 16S rRNA gene PCR products were visualized on 2% Tris-Borate-EDTA (TBE) agarose gel. DNA fragments at ∼450 bp were excised and extracted from the gels by using Qiagen Gel Extraction Kit. The extracted products were further purified by using a QiaQuick PCR Product Purification Kit (Qiagen, Valencia, CA). All final purified DNA products were quantitated by using Qubit HS Double-stranded DNA Kit (Invitrogen, Carlsbad, CA) prior to combining at equal mass for a final DNA concentration of 0.5 ng/µl. Pyrosequencing was performed at Michigan 99 State University RTSF by using Roche 454 GSFLX Sequencer. FeFe-hydrogenase (HydA) genes were amplified as described previously [11] and cloned with Topo Cloning (Invitrogen, Carlsbad, CA) and sequenced at the MSU-RTSF. Briefly, the primers used were FeFe-272F (5’GCHGAYMTBACHATWATGGARGA3’) and FeFe-427R (5’GCNGCYTCCATDACDCCDCCNGT3’) [11]. Each HydA PCR consists of a mixture of 1X Taq reaction buffer, DNase and RNase free water, 2 mM MgCl2 , 0.1 mg/ml BSA, 0.2 mM of each dNTP, 2 µM of each primer, 0.05 U/µl of Platinum Taq Polymerase (Invitrogen, Carlsbad, CA), and 20 ng DNA template. The amplification conditions include an initial denaturing step at 94◦ C for 4 minutes, followed by 35 cycles of 3 temperature steps consisting of denaturing at 94◦ C for 60 seconds, annealing at 56.5◦ C for 60 seconds, and an elongation at 72◦ C for 90 seconds, and a final extension at 72◦ C for 20 minutes following the completion of 35 cycles. All PCR products were stored at -20◦ C until needed. HydA PCR products (∼500 bp) were confirmed via electrophoresis (1% Tris-Acetate-EDTA agarose gel). Clone libraries of FeFe-hydrogenase were constructed by using Invitrogen Topo PCR2.1 Cloing Kit (Invitrogen, Carlsbad, CA) following the vendor’s protocol. The cloned fragments were purified and sequenced at RTSF via an ABI 3730 Genetic Analyzer. 4.3.4 Community Analysis Bacterial 16S rRNA gene sequences obtained from pyrosequencing were processed and analyzed by using Ribosomal Database Project (RDP) Pyrosequencing pipeline [180]. Sequences that were shorter than 250 bp, containing any N’s in sequencing region, or with an average quality score less than 20 were discarded. Chimeras were removed from RDP processed sequences by using USEARCH (UCHIME reference mode with Silva gold alignment database 100 as reference). RDP Multi-classifier was used to identify each sequence to bacterial genus level (threshold 80). RDP SeqMatch was used for further identification of bacterial clusters of interest. To eliminate the clustering artifacts, sequences were also trimmed to the same length (250 bp) based on alignment profiles prior to RDP complete linkage clustering. [FeFe]-hydrogenase gene sequences were identified by BLAST against NCBI non-redundant amino acid database. The BLAST E-value (expect value) was set at E < 10−5 . Top 50 best matches were generated for each sequence. The annotated match with the best E-value was then selected to be used to determine bacterial species information. 4.3.5 Isolation of Potential Uranium Immobilizing Bacteria Potential uranium immobilizers were isolated from FW107 and FW102-3 uranyl acetate enrichments on R2A agar plates in a Coy Lab anaerobic chamber ((H2 :CO2 :N2 = 3:10:87). A volume of 100 µl enrichment culture from each microcosm was initially spread on R2A plates anaerobically. Individual colonies were then picked and restreaked for isolation. Final isolates were stored in 7.5% DMSO at -80◦ C. 4.3.6 Growth Curves of Isolates The growth of each isolate under anaerobic conditions (H2 :CO2 :N2 = 3:10:87) was determined by measuring turbidities at 620 nm wavelength (OD620). Isolates were grown in 1/2X R2B broth in 96 well microtiter plates for 63 hours at 25◦ C. To determine if any isolates can utilize U6+ as an electron acceptor at 25◦ C, isolates were grown for 200 hours in minimal media (30 mM NaHCO3 , pH 6.8) supplemented with 20 mM sodium pyruvate as a carbon source and electron donor and 500 µM uranyl acetate. Isolates were also grown in minimal 101 media (30 mM NaHCO3 , pH 6.8) supplemented with 20 mM sodium pyruvate but without addition of an electron acceptor to determine if these strains were capable of fermentation. The turbidity measurements were performed under the same anaerobic conditions by using a Bio-Rad PR 3100 TSC Microplate Reader (Bio-Rad, Hercules, CA) 4.3.7 Isolate Identifications Final isolates were identified via 16S rRNA sequencing, which was carried out by the MSU RTSF. Isolate cultures were supplied in 96 well microtiter plate and MSU RTSF carried out 16S rRNA gene amplification and sequencing (primer 27F, 5’AGAGTTTGATCCTGGCTCAG3’ was employed). REP-BOX-(GTG)5 -PCR was also used to differentiate isolates that are substantially different at genomic level [49]. We used four primers, REP1RI (5’IIIICGICGICATCIGGC3’), REP2-I (5’IIICGNCGNCATCNGGC3’), GTG5 (5’GTGGTGGTGGTGGTG3’), and BOXA1R (5’CTACGGCAAGGCGACGCTGACG3’) for each reaction and set up PCR reactions as previously described [174]. Briefly, each 15 µl PCR reaction consisted of 3 µl 5X Gitschier buffer (Table C.1), 1.5 µl 10% DMSO, 30 pmol of each primers except BOXA1R (50 pmol), 0.75 µl of dNTP (25 mM of each), 0.85 mg/ml BSA, 0.24 µl of 5U/mul DNA Taq polymerase, 100 ng DNA template and water. To standardize the REP-BOX-(GTG)5 -PCR profile, Desulfitobacterium hafniense DCB-2 was also included in each batch of reactions. Gel electrophoresis was conducted with 2% TAE (tris-acetateEDTA) agarose gels and a 72 Volt electric potential. Final REP-BOX-(GTG)5 -PCR files were analyzed using Kodak Molecular Imaging Software. 102 4.3.8 Cell Suspension Assay To evaluate the uranium immobilization abilities of the isolates, we selected nine strains based on their REP-BOX-(GTG)5 -PCR profiles (presence and absence of DNA fragments). All of these strains were isolated from FW107 uranium enriched culture. Cell suspension assays [27, 88] were performed on these nine isolates along with the FW107 enrichment culture and a false positive control (sterile media without the inoculation of cells). Briefly, individual isolates (-80◦ C freezer stocks) were inoculated into anaerobic 1/2X R2B broth and grown to late exponential phase (OD600 ∼ 0.2). Cells were then washed with 30 mM bicarbonate buffer twice (centrifugations at 13,000 g for 15 minutes were carried out at the end of each wash). Final clean cell pellets were suspended in the reaction buffer (30 mM sodium bicarbonate at pH 6.8, 20 mM sodium pyruvate, 1 mM uranyl acetate) to OD600 0.1. Suspended cells were incubated at 25◦ C for 70 hours in anaerobic chamber (H2 :CO2 :N2 = 3:10:87) 4.4 4.4.1 Results Soluble U6+ Immobilization by FW107 and FW102-2 Enrichments Over a period of 64 days, we observed that both FW107 and FW102-2 enrichments were able to reduce the soluble U6+ concentrations from over 250 µM to under 150 µM (Figure 4.1). The U6+ concentration variations at day 0 were likely due to the inoculation method. An excess amount of uranyl acetate in the needle was injected, which was supposed to be retained. A slight increase in U6+ concentration was also observed during the early stage of incubation (0-10 days) in both enrichment microcosms and the control culture. This could be 103 due to interactions between media and uranyl acetate. All four enrichment cultures yielded a similar U6+ removal trend (Figure 4.1). 4.4.2 Bacterial Community Compositions of FW107 and FW102-2 Enrichments Pyrosequencing results showed that genera Clostridium and Pelosinus were abundant in the FW107 uranium enrichment culture. Together, they constituted over 80% of the community. As time increased, we observed an increasing amount of unclassified Clostridiales and a decreasing amount of Clostridium (Figure 4.2). Unlike FW107, the FW102-2 uranium enrichment culture was dominated by Tolumonas, which was initially found over 90% abundant. The relative abundance of this genus decreased as a function of time. Meanwhile, we observed that Pelosinus (currently identified as Psychrosinus by RDP) became more abundant (from approximately 0 to almost 15%) as the incubation time increased (Figure 4.3). The uranium enriched FW102-2 consortia initially (at time 0) contained few copies of [FeFe]-hydrogenase genes initially. With the same amount of DNA template, the hyd A amplicon fragment was barely visible (data not shown). Fewer hyd A clones were derived from FW102-2 enrichments than from FW107 enrichments at time 0 (8 clones VS. 66 clones) (Figure 4.4). The difficulty in [FeFe]-hydrogenase gene amplification and cloning could be due to either the lack of a hyd A containing population or experimental procedure bias. The abundance of hyd A genes detected increased significantly (to 50) after enriching in uranyl acetate (Figure 4.4 FW102-2 U64days). While the changes in hydA clones obtained were not quantitatively proven, it is likely due to the increase of [FeFe]-hydrogenase containing bacteria after incubation. A large number of hyd A sequences in the FW102-2 uranium consortia were found in 104 uncultured bacteria. In contrast, hyd A genes identified in both FW107 1/2X R2B enrichment and the uranium enriched consortia belonged to two large genera groups, Clostridium and Desulfotomaculum. We also observed that hyd A sequences from these two groups became more abundant after the uranium enrichment process (Figure 4.4 FW107). 105 Figure 4.1: The removal of soluble U6+ in FW107 and FW102-2 enrichments. The high soluble U(VI) concentration in the control culture (No Cells) was due to the excess amount of uranyl acetate carried over from the syringe needle during inoculation. 106 3.5 3 2.5 2 1.5 1 0.5 0 Pelosinus Clostridium Minor Groups 10 20 30 40 Time (days) Sulfurospirillum unclassified_ "Ruminococcaceae" Trichococcus Ralstonia Desulfitobacterium 50 60 0 Minor Group Percentage Pelonsinus/Clostridium Percentage 80 70 60 50 40 30 20 10 0 unclassified_Clostridiales unclassified_Veillonellaceae Herbaspirillum Desulfosporosinus Succinispira Figure 4.2: Bacterial community shifts with the increasing of incubation time in FW107 U-PBAGW microcosm. 107 Pelosinus Tolumonas Minor Groups 16 14 12 10 8 6 4 2 0 0 10 20 30 40 Time (days) Herbaspirillum Clostridium unclassified Gammaproteobacteria Sulfurospirillum 50 8 7 6 5 4 3 2 1 0 60 Minor Group Percentage Pelosinus Percentage Tolumonas Percentage 100 90 80 70 60 50 40 30 20 10 0 unclassified_Clostridiales Desulfosporosinus unclassified_"Lachnospiraceae" Microvirgula unclassified_Enterobacteriaceae Figure 4.3: Bacterial community shifts with the increase of incubation time in FW102-2 U-PBAGW microcosm. 108 Hyd Clone OTU 8 6 50 11 66 8 66 12 FW102-2 Hyd Clone OTU FW107 U 0 days U 64 days Figure 4.4: Distribution of [FeFe]-hydrogenase containing bacteria in FW107 and FW102-2 grown in 1/2X R2B media (U 0 days) and in uranyl acetate enrichments (U 64 days). The bacterial populations represented by the small fractions of each pie chart are listed on the right hand side with color blocks corresponding to the colors found in each pie chart. The distribution of genera Clostridium (cyan color) and Desulfotomaculum (purple color) are shown as the outer ring of each pie chart. 109 Figure 4.4 (cont’d) 110 4.4.3 Potential U6+ Immobilizers We have isolated 51 anaerobic bacteria from FW107 and FW102-2 uranium enrichments. We were able to identify 48 of them according to their 16S rRNA sequences. All of these isolates belonged to phylum Firmicutes. RDP SeqMatch identified these isolates as Clostridium XI, Clostridium XIVa, Clostridium sensu stricto, and Pelosinus. Phylogenetic analysis of the isolates 16S rRNA sequences by Arb agreed with RDP SeqMatch and revealed that 29 isolates clustered closest to Clostridium sardiniense (Figure 4.5 Cluster I), 13 isolates clustered closest to uncultured bacteria that are similar to Clostridium methylopentosum (Figure 4.5 Cluster II), and 4 isolates clustered closest to Pelosinus fermentans (Figure 4.5 Cluster III). Two isolates, A04 and B03 did not belong to any of the clusters but appeared more closely related to Cluster I than the other groups (Figure 4.5). Although the 16S rRNA sequences indicated that majority of the isolates were highly similar, we observed substantial differences in their growth rates at 25◦ C in anaerobic 1/2X R2B (Figure 4.6). Thus the 16S rRNA sequence alone is not sufficient to differentiate these isolates. We employed REPBOX-(GTG)5 -PCR and the profiles revealed that isolates identified as the same bacterial species according to their 16S rRNA sequences were substantially different in their genome organization based on REP-BOX-(GTG)5 profiles. These profiles were clustered and are presented in Figure 4.7. 111 Figure 4.5: The phylogenetic distribution of the anaerobic isolates based on their 16S rRNA gene sequences. The numbers on the branches represent the bootstrap values (1000 permutations). 112 Figure 4.5 (cont’d) 113 Figure 4.6: The growth curves of FRC isolates in 1/2X R2B under anaerobic condition (H2 :CO2 :N2 3:10:87). See next page for legends. 114 Figure 4.6 (cont’d) A1_big E1 C12_small B9_small A12_big A3_small E3_big D3_big C10_small B11 A5 E7 D7_big C1_2 B3_(small) A8 neg D11 D5 C5 A11_small A1_small E2 C7_small B9_big A12_2_small A3_big E5 D11_2 C10_big B1_2_small A6 E9 C12_big C2 B5_small A9 A4_small E6_small D6 B8_small A11_big Blank D8_small C7_big B9_2_small A12_2_big A10_small neg2 D12 115 B12 B1_2_big A2_small E3_small D2 C3 B5_big B1_small A4_big E6_big C6_small B8_big A7 E10_small D8_big C8 B9_2_big B6 A10_big nge3 C11 C1 A12_small A2_big E3_median D3_small C4_small B10 B1_big D7_small C6_big B2 A8_small E10_big D9 C9 C4_big B7 D2 D7 A10 B12 E10 A4 A9 B11 B8 E6 D3 E7 A1 A3 B5_2 C4 C10 C11 Dhaf_32.50 Dhaf_1.8 Dhaf_17.32 Dhaf_51.68 Dhaf_1.16 Dhaf_redo B9 E3 C12_A D6 A2 B10 D8 B9_2 A12 C1 C2 E1 B6 E9 D12 A11 B3 A5 C8 D9 D11_2 B7 C5 D5 D11 C9 C3 C8_2 B2 B1 A8 C6 A7 4 3 2 Distance 1 0 Figure 4.7: REP-BOX-(GTG)5 -PCR profiles of potential uranium immobilizing bacteria. Branches labeled Dhaf represent control samples. The rest represents enrichment isolates. 116 4.4.4 U6+ Immobilization By Isolates We observed that some isolates were able to grow anaerobically in minimal media supplemented with uranyl acetate and sodium pyruvate, while little or no growth was observed in minimal media supplemented with sodium pyruvate alone (Figure C.2 B9 and D5, and Figure C.3 A1, C9 and E2). We also observed isolates grew better in sodium pyruvate media instead of uranyl acetate and sodium pyruvate media (Figure C.2 B10 and C1, and Figure C.3 A10 and C1). However, due to the long incubation time in 96 well plates, we could not obtain reliable results (Figure C.2 and Figure C.3). Therefore we employed REPBOX-(GTG)5 profiles and 16S rRNA identifications to dereplicate our isolate collection. Within each bacteria group that was identified by 16S rRNA gene sequences, the isolates were subgrouped based on their REP-BOX-(GTG)5 -PCR profiles. From FW107 consortia, nine strains were selected based on their unique REP-BOX-(GTG)5 -PCR profiles and 16S rRNA sequences combinations to examine their ability of anaerobic uranium immobilization. These nine isolates are shown in Figure 4.8 as A4 (Figure 4.5 A04), A5 (Figure 4.5 Cluster III), A8 (Figure 4.5 Cluster I), A9 (Figure 4.5 Cluster I), A11 (Figure 4.5 Cluster I), A12 (Figure 4.5 Cluster I), B1 (Figure 4.5 Cluster I), B8 (Figure 4.5 Cluster II), and B10 (Figure 4.5 Cluster III). Sample FW107 represent the FW107 uranium enrichment culture. The cell suspension assays were performed on these ten cultures to determine the amount of soluble uranium removed. After 70 hours of incubation, we observed that isolates A11, A12, B1 and B8 removed the largest amount of U6+ , some uranium removal was observed in A8 and A9, A4 and A5 showed a comparable amount of removal to FW107, and little to no uranium removal was observed in B10. (Figure 4.8). 117 1200 Soluble U(VI) (µM) 1000 800 T0_avg 600 T70_avg 400 200 0 neg A4 A5 A8 A9 A11 A12 B1 B8 B10 FW107 Figure 4.8: Soluble U6+ removal by Firmicutes isolates (from FW107) and FW107 uranium enriched community after 70 hours (T70) of incubation at room temperature in an anaerobic chamber (H2 :CO2 :N2 3:10:87). Sample "neg" represents a control culture without the addition of any bacterial cells. The error bars represent measurements of two replicates. 118 4.5 Discussion In this part of our study, we were able to monitor the reduction of soluble U6+ in FW107 and FW102-2 uranium enrichments. Consistent with our findings in studies performed in Chapter 3, we observed that the abundance of Pelosinus (currently identified as Psychrosynus by RDP) increased along with the enrichment time (Figure 4.2 and 4.3). In addition, unclassified Clostridiales become more abundant in both enrichments during the same period, while soluble U6+ decreased (Figure 4.1). Thus, it is likely that members of Pelosinus and unclassified Clostridiales contributed to the U6+ removal processes. [FeFe]-hydrogenase distributions in the enrichment cultures suggested that Clostridium and Desulfotomaculum might be contributing to metabolic hydrogen production. However, some members of these two groups are known uranium reducers [45, 72], their predominant presence in the [FeFe]-hydrogenase gene communities might also be a simple statement of their uranium reduction dependent growth from U6+ removal. The combined sequence data from the 16S rRNA genes and [FeFe]-hydrogenases indicates that the Clostridia are numerically important members of the community with possible uranium reduction responsibilities, as well as participants in the hydrogen economy of the bioreactors. Although RDP SeqMatch and Arb phylogenetic analysis were consistent with each other and showed that the isolates grouped into four major Firmicute categories, RDP SeqMatch showed different species identifications. Among 48 Firmicute bacterial isolates, RDP SeqMatch revealed that the majority of the sequences in Clostridium XIVa group were identical to the 16S rRNA gene sequence of Desulfotomaculum guttoideum, which is a rarely studied sulfate-reducing bacterium [51]. At 95% to 97% similarity, sequences in Clostridium sensu 119 stricto group were closely related to Clostridium acetobutylicum, Clostridium butyricum, and Clostridium puniceum. While Clostridium acetobutylicum was recently reported as a uranium reducing bacterium [32], the other two were mostly studied for their chemical producing ability, such as butanol and 1,3-propandiol [22, 62]. The Pelosinus isolates we have identified were closely related to Pelosinus sp. strain UFO1, which was recently reported for its role multi-mode uranium immobilization [129]. Out of 48 isolates, only one was identified as Clostridium venationis in group Clostridium XI and no studies on anaerobic respiration or metal immobilization have been reported on this species. The above taxonomy information and literature reports suggest that studying Firmicutes has a great potential in understanding bacterial community activities in U6+ removal. The growth patterns and REP-BOX-(GTG)5 -PCR profiles suggested that these 48 isolates should not be treated simply as four groups. Instead, it is important to test isolates based on their REP-BOX-(GTG)5 -PCR profiles for their abilities to immobilize U6+ . Out of the ten isolates we selected, four removed substantially more soluble uranium than the others, including the FW107 uranium enriched community. It was to our surprise that the Pelosinus isolate A5 was not proficient at uranium removal despite its high abundance in the enrichment cultures. It was even more surprising to observe the drastic differences in uranium removal abilities between strains (e.g., Cluster I A9 VS. Cluster I A11 in Figure 4.8). Considering the high variabilities among isolates of the same species, it is very likely that we did not obtain a uranium removal efficient Pelosinus strain during the isolation processes. 120 4.6 Conclusion We previously described our observations on the prominence of Firmicute bacteria in uranium supplemented enrichment cultures (Chapter 3). In this work, we studied the shifts of bacterial communities during uranium immobilization. We have anaerobically isolated and identified 48 bacterial strains, which belong to phylum Firmicutes. These isolates were categorized into four different species according to their 16S rRNA gene sequences. However, their growth patterns and REP-BOX-(GTG)5 -PCR profiles indicated that they are physiologically and genomely distinct within each species. We evaluated the ability of uranium immobilization of isolate representatives. Our results indicated that many isolates with close phylogenetic affiliation to Clostridium sardiniense could immobilize uranium. 121 CHAPTER 5 PHYLOGENETIC COMPOSITION OF BACTERIA IN ANAEROBIC CO-DIGESTION OF DAIRY MANURE AND CORN STOVER 5.1 Abstract More than 160 million dry tons of animal waste are produced annually from animal farms in the United States. These wastes pose a great environmental and health risk if not disposed of properly. Anaerobic digestion is an effective method in breaking down animal wastes to reduce the risks and it produces methane as an alternative energy source. Previous studies suggested that optimization of feed composition, HRT, and other operational conditions can greatly improve total solids removal and increase methane productivity. Despite these improvements, the relationships between system performance and the underlying bacterial community shifts are not well understood. Therefore, we constructed anaerobic bioreactors fed with dairy manure (DS) and dairy manure supplemented with corn stover (DM+CS) to compare the bacterial community structures and identify important bacterial populations correlating to total solids removal and methane production. Our results suggested that unique uncultivated strains of Bacteroidetes played a substantial role in cellulose/hemicellulose degradation in the DM+CS bioreactors. Aminomonas paucivorans and Clostria populations utilize secondary metabolites produced during cellulose/hemicellulose degradation to generate hydrogen and acetate. Thus, promoting the growth of unclassified Bacteroidetes, Clostridia, and A. paucivorans might optimize the efficiency of co-digestion system and control the variability. 122 5.2 Introduction Agriculture is a significant part of the US economy. The Food and Agriculture Organization of the United Nations (FAO) estimated that approximately $0.2 trillion of production was contributed from agriculture in the US in 2010. The top three commodities based on production values were indigenous cattle meat, dairy products and maize. Maize was the most produced commodity based on quantity (at least three times more than other commodities). This large agricultural output resulted in considerable amounts of waste being produced. Approximately 6.7 million tons of agriculture wastes were produced in 2010 in the US (estimated based on 1993 data) [44]. The wastes were mainly used as fertilizer for crop growth. However, applying agricultural wastes, especially manures, to fields raises significant environmental concerns of greenhouse gas emissions, spread of animal pathogens, and release of excess nutrients to the environment, all of which may pose a great risk to human health and the ecosystem [20, 24, 100, 112]. Hence, a controlled animal waste management strategy that simultaneously treats the wastes and utilizes them for fuel/chemical production would have both environmental and economic benefits. Anaerobic digestion can effectively degrade farm wastes and concomitantly produce methane, thereby reducing the health and environmental risks while producing energy [59]. Methane is produced via methanogenesis in anaerobic digestion. Methanogenesis is a syntrophic process that takes place under anaerobic conditions when organic matter is in high concentration [140]. It is a highly cooperative process carried out by different groups of bacteria and archaea with distinct responsibilities [13]. Schink [140] schematically showed that methanogenic degradation of organic matter was executed by three groups of microor- 123 ganisms. During anaerobic methanogenesis, organic substrates are degraded by fermentative bacteria. Fermentation by-product hydrogen is used by hydrogen-oxidizing methanogens to produce methane, while acetate-cleaving methanogens produce methane from acetate. Previous research has shown that the substrate C:N ratio influences the production of methane in anaerobic digestion [46]. Several studies have demonstrated that a mixture of nutrient-rich manures and carbohydrate-rich crop and food wastes are suitable substrates for anaerobic digestion and support methanogenesis [30, 42, 46, 85, 110, 179]. While a large number of studies were focused on optimizing system operational parameters, the microbial communities within the anaerobic digestion have not been not well explored. As pointed out by Schink et al. [140], these microbial communities are the primary force behind methane production and community shifts could effect overall performance. Herein we report on a pilot study describing microbial communities of anaerobic bioreactors fed with dairy manure (DM) and corn stover (CS) using gene targeted metagenomics. 5.3 5.3.1 Materials and Methods Anaerobic Bioreactors Feedstock composition and the laboratory-scale anaerobic bioreactors (5 L working volume anaerobic fermenters) were set up and described by Yue et al. [192]. Briefly, six continuously stirred tank reactors (CSTR) were operated at 35◦ C with hydraulic retention times (HRTs) of 30 days, 40 days, and 50 days. The feed for diary manure bioreactors was 5% dry matter and 95% water. The feed for manure and corn stover (DM+CS) bioreactors was a mixture of dairy manure and corn stover in a ratio of 4:1 (wt:wt) and diluted to 5% DM using water. 124 Both feeds were homogenized using a blender prior to feeding. Analytical analyses of biogas, total solids (TS), and fiber composition were reported by Yue et al. [192]. 5.3.2 DNA Extraction and Targeted Gene Amplifications Anaerobic bioreactors were fed with DM or DM+CS at three different HRTs. Samples were taken from each bioreactor once digestion reached the stabilized phase (after two HRTs). The samples were centrifuged at 10,000 RPM for 20 minutes in a Sorvall SS-34 rotor and the pellet was stored at -20◦ C until extraction of community DNA. DNA was extracted by using Powersoil DNA Extraction Kit (Mo Bio Laboratory, Carlsbad, CA). Polymerase Chain Reactions (PCR) were carried out to amplify bacterial 16S rRNA gene (16S) and FeFe-hydrogenase genes (HydA) with universal primer sets: 357F (5’CCTACGGGAGGCAGCAG3’) and 926R (5’CCGTCAATTCMTTTRAGT3’) for 16S, and FeFe-272F (5’GCHGAYMTBACHATWAT GGARGA3’) and FeFe-427R (5’GCNGCYTCCATDACDCCDCCNGT3’) for HydA [11]. Bacterial 16S rRNA genes were amplified with sequence-tagged primers developed by the Human Microbiome Project (HMP) targeting the V3-V5 region of 16S rRNA genes. Briefly, 1X Taq reaction buffer, 2 mM MgSO4 , 0.1 mg/ml BSA, 0.2 mM of each dNTP, 0.33 µM of each primer, 0.125 U/µl of Platinum High Fidelity Taq Polymerase (Invitrogen, Carlsbad, CA), and 10 ng DNA template were mixed with DNase and RNase free water for each PCR reaction. The amplification conditions include an initial denaturing step at 95◦ C for 5 minutes, followed by 30 cycles of 3 temperature steps consisting of denaturing at 95C for 45 seconds, annealing at 50◦ C for 45 seconds, and elongation at 72◦ C for 90 seconds, and a final extension at 72C for 5 minutes following the completion of 30 cycles. All PCR products were stored at -20◦ C until sequencing at the MSU Research Technologies Support Facilities 125 (MSU-RTSF). FeFe-hydrogenase genes were amplified as described previously [11] and cloned with Topo Cloning (Invitrogen, Carlsbad, CA) and sequenced at the MSU-RTSF. Each HydA PCR consists of a mixture of 1X Taq reaction buffer, DNase and RNase free water, 2 mM MgCl, 0.1 mg/ml BSA, 0.2 mM of each dNTP, 2 µM of each primer, 0.05 U/µl of Platinum Taq Polymerase (Invitrogen, Carlsbad, CA), and 20 ng DNA template. The amplification conditions include an initial denaturing step at 94◦ C for 4 minutes, followed by 35 cycles of 3 temperature steps consisting of denaturing at 94◦ C for 60 seconds, annealing at 56.5◦ C for 60 seconds, and elongation at 72◦ C for 90 seconds, and a final extension at 72◦ C for 20 minutes following the completion of 35 cycles. All PCR products were stored at -20◦ C until needed. 5.3.3 Preparation for Pyrosequencing and Construction of Clone Libraries The 16S rRNA gene PCR products were visualized on 2% Tris-Borate-EDTA (TBE) agarose gel. DNA fragments at ∼569 bp were excised and extracted from the gels by using Qiagen Gel Extraction Kit and further purified with QiaQuick PCR Product Purification Kit (Qiagen, Valencia, CA). All purified DNA products were quantitated with Quibit HS Double-stranded DNA Kit (Invitrogen, Carlsbad, CA) then composited at equal mass for a final DNA concentration of 0.5 ng/µl. Pyrosequencing was performed at Michigan State University RTSF by using a Roche 454 GSFLX Titanium Sequencer. HydA encoding gene sequence PCR products (∼500 bp) were confirmed via electrophoresis (1% Tris-Acetate-EDTA agarose gel). Clone libraries of FeFe-hydrogenase were constructed by using Invitrogen Topo PCR2.1 Cloning Kit (Invitrogen, Carlsbad, CA) following the vendor’s protocol. The cloned fragments were purified and sequenced at RTSF via an 126 ABI 3730 Genetic Analyzer. 5.3.4 Sequence and Community Analyses Bacterial 16S rRNA gene sequences obtained from pyrosequencing were processed and analyzed by using the Ribosomal Database Project (RDP) Pyrosequencing pipeline [26]. Sequences that were shorter than 250 bp, containing any N’s in sequencing region, or with an average quality score less than 20 were discarded. Chimeras were removed from RDP processed sequences by using USEARCH (UCHIME reference mode with Silva gold alignment database as reference). RDP Multi-classifier was used to identify each sequence to bacterial genus level (threshold 80). RDP SeqMatch was used for further identification of bacterial clusters of interest. To eliminate the clustering artifacts, sequences were also trimmed to the same length (250 bp) based on alignment profiles prior to complete linkage clustering at the RDP. [FeFe]-hydrogenase gene sequences were identified by BLAST against the NCBI nonredundant amino acid database. The top 50 best matches were generated for each sequence. The annotated match with the best expected value (E < 10−5 ) was selected in determining bacterial species information. OTU- and phylotype-based community analyses were performed by using statistical software R [128]. Ecological indices were estimated for each bioreactor sample based on clustered sequences by using the R package Vegan [117]. Specifically, community richness (chaoI), diversity indices (Shannon H’ and Simpson’s D) and community evenness (E) were calculated based on equations previously described [94]. The sampling coverage (C) of each sample was calculated based on Good’s method, C=1-n1 /N, where n1 represents the number of OTUs 127 that is a singleton and N is the total number of sequences in the sample [39]. The R package Vegan was also used to perform non-metric multidimensional scaling analysis (NMDS) to correlate the dissimilarities among bioreactor samples and the shifts in phylotype abundance [117]. Shared OTUs among samples were calculated and plotted by using the R package Vennerable [159]. 5.4 5.4.1 Results The Influence of HRT and Corn Stover on System Performance For bioreactors fed with DM, a substantial increase on biogas productivity was observed once the HRT increased from 30 to 40 days (Figure 5.1). The biogas productivity of the DM bioreactors increased from 209 to 399 ml/g TS loading/day. A similar increase was also observed in the DM+CS bioreactors (from 270 to 497 ml/g TS loading/day). Further increases in HRT had no significant (P > 0.05) influence on the biogas productivity for either feeds. The TS reduction for both feeds also increased with the increases in HRT (Figure 5.1). The HS reduction increased from 27% (HRT of 30 days) to 50% (HRT of 50 days) in the DM bioreactors. A similar trend was observed in the DM+CS bioreactors (from 35% at a HRT of 30 days to 53% at a HRT of 50 days). In addition, digestion performance demonstrated that under the same HRT, both biogas productivity and TS reduction were significantly (P < 0.05) higher in the DM+CS bioreactors than in the DM bioreactors, presumably due to the introduction of fast-hydrolyzing components from corn stover in DM+CS feed [192]. 128 Figure 5.1: The total solids (TS) reduction efficiency (diagonal and grid lined columns) and the production rate of biogas (grey and white columns) in the DM and DM+CS bioreactors at four HRTs: a HRT of 30 days (HRT30d), a HRT of 40 days (HRT40d) and a HRT of 50 days (HRT50d). Solid diamonds and solid triangles represent the amount of methane produced in the DM and DM+CS bioreactors, respectively. 129 The trend of biogas productivity and TS reduction indicated that bioreactor performance underwent significant (P < 0.05) changes from 30 days of HRT to 40 days or longer of HRT. The biogas productivities for both DM and DM+CS bioreactors were maximized at 40 days of HRT. Longer HRTs did not increase the biogas productivity but significantly more TS was removed (P < 0.05). In spite of different biogas production rates and TS reduction levels from different bioreactors, the methane content of biogas from all six bioreactors fell within the narrow range from 65.7% to 67.4%, and this was relatively stable throughout the experiment. The pH measurements indicated that all six bioreactors were maintained at near neutral condition (pH 6.58-6.94). 5.4.2 Phylogenetic Comparisons of Anaerobic Bioreactor Bacterial Communities Bacterial 16S rRNA gene targeted sequencing revealed that we sampled over 98% of each anaerobic bioreactor community (Table D.1 and Figure D.2). The evenness of the bacterial communities were relatively stable (0.68 to 0.78) but bacterial diversity shifted drastically from sample to sample. All diversity indices indicated that the DM+CS bioreactor with a HRT of 40 days had the most diverse bacterial community (Table D.1). 130 40 50 45 40 35 30 25 20 15 10 5 0 A1. 35 30 25 20 15 10 5 0 HRT30d HRT40d Number of OTU's Bacteroidetes Abundance (%) 45 HRT50d Figure 5.2: The abundance and diversity (OTU at 97% sequence similarity) of phyla Bacteroidetes (panel A1), class Clostridia (panel B1), and Escherichia / Shigella (panel C1) in each bioreactor bacterial community. The distribution of OTUs among samples are presented as Venn diagrams in panel A2 (Bacteroidetes), B2 (Clostridia) and C2 (Escherichia/Shigella). The colors of the bioreactor/HRT font corresponds to an area bounded by the same color. Labels HRT30d, HRT40d and HRT50d represent bioreactors with a HRT of 30, 40, and 50 days, respectively. 131 Figure 5.2 (cont’d) A2. 7 0 0 1 0 0 2 0 0 0 0 0 1 1 DM HRT40d 0 0 0 DM+CS HRT50d 0 DM+CS HRT30d 0 0 0 0 0 9 0 0 2 0 1 0 0 2 0 3 1 0 0 2 0 0 1 1 0 0 0 0 DM HRT30d 11 0 1 0 0 0 0 0 10 0 0 1 6 1 12 2 5 DM HRT50d Bacteroidetes Venn Diagram 132 DM+CS HRT40d 30 25 B1. DM DMCS 120 100 DM_OTU 20 DMCS_OTU 80 15 60 10 40 5 20 0 0 HRT30d HRT40d 133 HRT50d Number of OTU's Clostridia Abundance (%) Figure 5.2 (cont’d) Figure 5.2 (cont’d) B2. 15 0 0 3 1 1 1 0 2 2 2 9 0 1 DM+CS HRT30d DM HRT40d 1 0 0 DM+CS HRT50d 0 0 4 2 1 1 0 1 3 3 1 0 1 2 0 1 1 3 2 1 6 5 0 2 1 0 1 0 4 4 1 DM HRT30d 7 1 4 0 1 0 0 11 1 18 3 34 5 8 DM HRT50d 46 Clostridia Venn Diagram 134 DM+CS HRT40d 30 Abundance (%) Escherichia/Shigella 35 9 8 C1. 7 25 6 20 5 15 4 3 10 2 5 1 0 0 HRT30d HRT40d 135 HRT50d Number of OTU's Figure 5.2 (cont’d) Figure 5.2 (cont’d) C2. 1 0 0 0 0 0 0 0 0 0 1 0 0 DM HRT40d 1 1 0 DM+CS HRT50d 0 0 0 0 0 1 0 DM+CS HRT30d 0 1 2 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 DM HRT30d 0 0 0 0 0 0 0 0 0 0 2 0 0 0 1 0 DM HRT50d Escherichia/Shigella Venn Diagram 0 0 0 DM+CS HRT40d Bacterial taxa were determined based on 16S rRNA genes via RDP Multi-Classifier with a minimal bootstrap value of 80 (assigned as unclassified if score is smaller than 80). A total of 16 bacterial phyla were assigned and approximately 5% of total sequence was categorized 136 as unclassified bacteria (3-7% unclassified bacteria in each sample). At the genus level, a total of 198 bacterial groups (137 classified and 61 unclassified) were identified. Firmicutes, Bacteroidetes, and Proteobacteria were the most abundant phyla in all six anaerobic bioreactor samples (Figure D.2 wide columns). Within these three phyla, Clostridia, unclassified Bacteroidetes, and Escherichia/Shigella were found to be highly abundant (Fgiure D.2 thin columns). We observed that Clostridia comprised 74-84% of the Firmicutes in the bioreactors. Phylum Bacteroidetes was largely composed of unclassified Bacteroidetes, except in the DM a HRT of 30 days bioreactor that had abundant Alkaliflexus, Petrimonas, unclassified Porphyromoadaceae and Flavobacterium. Although the abundance of phylum Bacteroidetes was similar across the six bioreactors, the abundance of unclassified Bacteroidetes increased as the HRT increased. In the DM+CS bioreactor with a HRT of 40 days, 35.45% of the community populations were identified as unclassified Bacteroidetes, representing 87% of the Bacteroidetes populations. Within phylum Proteobacteria, we observed that genera Escherichia/Shigella were consistently abundant and its abundance variation reflected the abundance of Proteobacteria except in the DM+CS bioreactor with a HRT of 40 days. Thus, the population of Escherichia/Shigella increased as a function of HRT and on average constituted 17% of the bacterial community except in the DM+CS bioreactor with a HRT of 40 days, where it decreased to 1.53%. Over 50% of the phylum Proteobacteria in the DM+CS bioreactors with a HRT of 30 days, the DM bioreactors with a HRT of 40 days and the DM+CS bioreactors with a HRT of 50 days were identified as Escherichia/Shigella, while 37.71% and 21.21% of Proteobacteria were classified as Escherichia/Shigella in the DM bioreactor with a HRT of 30 days and DM+CS bioreactor with a HRT of 40 days, respectively (Figure D.2). Besides the changes in abundance, we also observed the diversity shifts within 137 phylum Bacteroidetes, class Clostridia, and genera Escherichia/Shigella (Figure 5.2). When operated at the HRTs of 30 days and 50 days, phylum Bacteroidetes was more abundant in DM bioreactors than in DM+CS bioreactors. Phylum Bacteroidetes also appeared to be less abundant in both DM and DM+CS bioreactors under the HRT of 50 days than when they were operated with a HRT of 30 days. At a HRT of 40 days, the abundance of phylum Bacteroidetes was substantially larger in DM+CS bioreactors than DM bioreactors (40.38% in DM+CS a HRT of 40 days VS. 21.47% in DM a HRT of 40 days). The Bacteroidetes diversity (OTU) pattern followed its abundance trend (Figure 5.2 A1). A larger numbers of OTUs were observed in DM bioreactors than in DM+CS bioreactors at HRTs of 30 and 50 days. The number of OTUs decreased in DM and DM+CS bioreactors slightly as HRT increased from 30 to 50 days. The lowest Bacteroidetes diversity was observed in the DM bioreactor at a HRT of 40 days (25 OTUs). In contrast, the DM+CS bioreactor at a HRT of 40 days revealed the largest Bacteroidetes diversity (43 OTUs). A total nine OTUs were shared by all six samples. The proportion of unique OTUs in each sample ranged from 19% in the DM+CS bioreactor at a HRT of 50 days bioreactor to 32% in DM a HRT of 30 days and DM a HRT of 50 days bioreactors (Figure 5.2 A2). Similar to phylum Bacteroidetes, the abundance of class Clostridia decreased when HRT increased from 30 days to 50 days (Figure 5.2 B1). When operated at HRTs of 30 and 50 days, class Clostridia was more abundant in DM bioreactors than DM+CS bioreactors. The abundances of class Clostridia were similar in the DM (25.5%) and DM+CS (26.6%) bioreactors operated at a HRT of 40 days. The diversity within class Clostridia followed the same trend as observed in the abundance profiles at HRTs of 30 and 50 days. At a HRT of 40 days, class Clostridia was substantially more diverse in the DM+CS bioreactor than in the DM bioreactor. With an 138 abundance difference of 1.1% at a HRT of 40 days, the DM+CS bioreactor was observed to have 105 OTUs of Clostridia, while only 75 Clostridia OTUs were detected in the DM bioreactor (Figure 5.2 B1). There were four Clostrdia OTUs shared by all six samples. A large portion of the OTUs were unique to each sample, ranging from 20% in DM a HRT of 50 days and DM+CS a HRT of 50 days bioreactors to 42% in DM a HRT of 30 days bioreactor (Figure 5.2 B2). Unlike phylum Bacteroidetes and class Clostridia, the abundance of genera Escherichia/Shigella increased in the bioreactors as HRT increased from 30 to 50 days (Figure 5.2 C1). It was also observed that the genera Escherichia/Shigella was more abundant in the DM+CS bioreactors than in the DM bioreactors when operational HRTs were 30 and 50 days. Escherichia/Shigella were the least abundant (1.53%) in the DM+CS bioreactor operated at HRT 40 days. In contrast, the DM bioreactor operated at a HRT of 40 days was substantially more abundant with Escherichia/Shigella (19.55%). The diversity profile revealed that when the abundance of Escherichia/Shigella increased from 5.82% (the DM bioreactor with a HRT of 30 days) to 19.55% (the DM bioreactor a HRT of 40 days), the number of OTUs stayed the same (6). However, the Escherichia/Shigella OTUs decreased from 8 to 3 in DM+CS bioreactors when the Escherichia/Shigella abundance decreased substantially at a HRT of 40 days. Finally, the number of Escherichia/Shigella OTUs remained the same (from 3 to 4) when their abundance increased from 1.53% (the DM+CS bioreactor with a HRT of 40 days) to 29.56% (the DM+CS bioreactor with a HRT of 50 days) (Figure 5.2 C1). Among these OTUs, two were found in all six samples. Two of six OTUs were unique to the DM bioreactor with a HRT of 30 days, and one of eight was unique to the DM+CS bioreactor with a HRT of 30 days and the DM bioreactor with a HRT of 50 days. No unique Escherichia/Shigella OTUs were found in the DM bioreactor with a HRT of 40 139 days, the DM+CS bioreactor with a HRT of 40 days and the DM+CS bioreactor with a HRT of 50 days (Figure 5.2 C2). 5.4.3 Bacteroidetes Bacteroidetes was one of the most abundant phyla in the bioreactor samples described above. The genera abundance revealed that unclassified Bacteroidetes was also predominant and highly variable (Appendix C Figure C.2). Sequences identified as members of phylum Bacteroidetes by RDP Classifier were extracted and separated into groups of unclassified Bacteroidetes and classified Bacteroidetes (all taxon groups that could be identified further than phylum Bacteroidetes). The abundance of phylum Bacteroidetes in bacterial communities varied from 21% to approximately 40% (Figure 5.3 vertical striped area). We observed that the composition of phylum Bacteroidetes shifted from classified Bacteroidetes to unclassified Bacteroidetes. Across six bioreactor samples, unclassified Bacteroidetes comprised as much as 88% of the phylum Bacteroidetes (the DM+CS bioreactor a HRT of 40 days). The least amount of unclassified Bacteroidetes was observed in the DM bioreactor with a HRT of 30 days, where only 15% of Bacteroidetes were unable to be identified further than the phylum level. The composition of phylum Bacteroidetes revealed that unclassified Bacteroidetes were more abundant than classified Bacteroidetes in the DM+CS bioreactors regardless of HRT (Figure 5.3). However, in the DM bioreactor samples, classified Bacteroidetes were substantially more abundant than unclassified Bacteroidetes at HRT 30 days (the DM bioreactor at a HRT of 30 days) and became much less abundant in the DM bioreactor with a HRT of 50 days. At the same retention time, unclassified Bacteroidetes were more abundant in the DM+CS bioreactors than in the DM bioreactors and vice versa for classified 140 Bacteroidetes. It was also observed that unclassified Bacteroidetes increased in abundance as the HRT increased in both the DM and DM+CS bioreactors, while the abundance of classified Bacteroidetes decreased (Figure 5.3). At 97% similarity, sequences in phylum Bacteroidetes were clustered into different groups. It was observed that phylum Bacteroidetes was abundant (40%) and diverse (over 40 OTU) in the DM+CS bioreactor operated at a HRT of 40 days (Figure 5.2A). In contrast, Bacteroidetes was the least abundant and diverse (approximately 25 OTU) in DM bioreactor operated at a HRT of 40 days. The Bacteroidetes clusters (at 97% similiarity) in the DM+CS bioreactor with a HRT of 40 days were compared to the Bacteroidetes clusters in other samples and four distinct clusters (#11, #14, #43, and #69) were identified (Figure 5.4). These four clusters were abundant in the DM+CS bioreactor with a HRT of 40 days where together they comprised of 62% of the Bacteroidetes populations in the DM+CS bioreactor with a HRT of 40 days. These clusters were among the unclassified Bacteroidetes. Cluster #69 appeared unique to the DM+CS bioreactor with a HRT of 40 days as it was not detected in other samples (the DM bioreactor with a HRT of 30 days, the DM bioreactor with a HRT of 40 days, the DM bioreactor with a HRT of 50 days and the DM+CS bioreactor with a HRT of 50 days) or detected with a 0.3% abundance (the DM+CS bioreactor with a HRT of 30 days). Cluster #11 and #14 were observed in the DM+CS bioreactor samples (the DM+CS bioreactor with a HRT of 30 days, the DM+CS bioreactor with a HRT of 40 days and the DM+CS bioreactor with a HRT of 50 days) at high abundance (6-27% and 15-29% of the phylum Bacteroidetes in each sample, respectively). These two clusters were also observed in the DM bioreactor samples (the DM bioreactor with a HRT of 30 days, the DM bioreactor with a HRT of 40 days and the DM bioreactor with a HRT of 50 days) but at a substantially 141 Abundance of Bacteroidetes (%) Classified Bacteroidetes 100 80 60 40 20 Unclassified Bacteroidetes 0 20 40 60 80 100 DM HRT 30d DM+CS DM HRT 40d DM+CS DM HRT 50d DM+CS 0 20 40 60 80 100 Bacteroidetes Abundance in Communities (%) Total DM DM+CS Figure 5.3: The abundance of phyla Bacteroidetes (vertical stripes) in each bioreactor bacterial communities. The distribution of classified and unclassified Bacteroidetes within phylum Bacteroidetes were calculated and displayed in left panel (white and diagonal striped bars) and right panel (black and gridded bars) respectively. Labels HRT30d, HRT40d and HRT50d represent bioreactors with a HRT of 30, 40, and 50 days, respectively. 142 lower abundance (1-7% and 0.4-2% of the phylum Bacteroidetes in each sample, respectively) but became more abundant as the HRT increased. Thus, the abundance of #11 and #14 could be ranked as the DM bioreactor with a HRT of 30 days