a... ,. nmfihuté xt.....xx 1.. :1... ‘ .1... I It... I . . l... v1v3Iy \wx...v3...:.q . .. \, 57.1,: Em $5.91.. 11.: This is to certify that the dissertation entitled VERTICAL AND HORIZONTAL DISTRIBUTION OF DENITRIFIER COMMUNITIES IN PACIFIC NORTHWEST AND ARCTIC MARINE SEDIMENTS CD ‘5 4—: E w a < C 5 presented by .m (u 51>) O) .99 E if) D V ' ' G tz " eronlca riJn i E 9 has been accepted towards fulfillment of the requirements for the PhD. degree in Microbiology and Molecular 1 Genetics \ I ijor Protessor'fi Signature Anzac/#275, QQ?‘ _ Date MSU is an affinnstive-action, equal-opportunity employer PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATEDUE DATEDUE DAIEDUE 6/07 p./ClRC/DaIeDue.indd-p.1 VERTICAL AND HORIZONTAL DISTRIBUTION OF DENITRIFIER COMMUNITIES IN PACIFIC NORTHWEST AND ARCTIC MARINE SEDIMENTS By Veronica Grilntzig A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Microbiology and Molecular Genetics 2007 ABSTRACT VERTICAL AND HORIZONTAL DISTRIBUTION OF DENITRIFIER COMMUNITIES IN PACIFIC NORTHWEST AND ARCTIC MARINE SEDIMENTS By Veronica Grtintzi g Denitrification in marine sediments is a major sink for nitrogen in the global nitrogen budget and the main sink of fixed nitrogen in the oceans. Therefore, understanding the distribution of denitrifier communities, their activity, and types of denitrifiers in sediments may help refine denitrification estimates and better model the marine N cycle. The vertical and horizontal distribution of denitrifiers was studied in sediments from various locations in the Pacific Northwest (Puget Sound, continental slope and abyssal sea floor) and the Arctic with different biogeochemical characteristics, using the heme cd, nitrite reductase gene (nirS) as a molecular marker. Since a method was lacking to rapidly and accurately quantify the organisms involved in this key biogeochemical process, real-time PCR was tested for its applicability for the quantification of a functional gene, the nirS gene of Pseudomonas stutzeri, in the environment. The assay was specific, sensitive down to one copy of the gene, precise and accurate, and indicated that P. stutzeri, though fairly cosmopolitan, might not be as dominant as previously thought. On the other hand, site-specific abundant nirS gene sequences, unrelated to known denitrifiers, were detected at each site. The vertical distribution of denitrifiers was correlated to the bioturbation of the sediments and not to the redox gradients. A conserved denitrifier community structure, as well as denitrification capacity, was detected throughout the mixed sediment layers, reaching as deep as 37 cm in some locations, regardless of the presence of oxygen or nitrate. In the horizontal dimension the different locations seem to harbor distinct denitrifying and bacterial communities, as indicated by statistically significant differences between nirS gene clone libraries as well as 16S rRNA gene clone libraries at each site, in addition to site-specific clustering of highly similar sequences for both genes. Differentiation of denitrifying communities was mostly affected by geographic location, followed by differing environmental factors, often related to water depth, within one area, and least affected by sediment depth. The bacterial community, as studied by the 16S rRNA gene, was less influenced by geographic location, probably due to the lower mutation rate of this gene. Although all sites harbored rich and diverse denitrifier and bacterial communities, a general trend towards higher diversity at intermediate productivity levels was detected, suggesting a possible correlation between these two factors. ACKNOWLEDGMENTS Many, many people have contributed in one or the other way in the completion of this work and although I’m not going to be able to name all of them, be assured that I’m grateful towards, remember and cherish each one of them. First of all I want to thank Dr. James Tiedje for giving me the opportunity to join his lab, for his guidance and teachings, for always being available for me regardless of his busy agenda, taking the time to even discuss my thesis in Germany, and for his constant support and encouragement. It has been a great honor and pleasure to form part of the Tiedje lab. I want to thank my committee members Dr. Allan Devol, Dr. Terry Marsh, and Dr. Mike Klug for their guidance and suggestions, for always being available to answer my questions, for turning every meeting into an interesting scientific discussion rather than an examination, and for their patience, constant support and encouragement. Thanks also to Dr. Allan Devol for allowing me to work in his lab and participate in the sampling cruises, which have been wonderful experiences, for teaching me the oceanographic point of view of my work, for his guidance during the sampling and microcosms experiment, and for even getting muddy while helping out with them. Thanks to Dr. Terry Marsh for his guidance and teachings throughout my Ph.D., but especially during my rotation through his lab, a great experience, which has introduced me to many aspects of my later work. Thanks to Dr. Mike Klug, for being absolutely encouraging every time I contacted him, making me smile with each e-mail, and giving me the strength to think that I was able to finish this task. So, again, thank you very much to Dr. Tiedje and my committee for having guided me all the way through my Ph.D. to finally be able to write these acknowledgments! iv Thank you to all the past and present members of the Tiedje lab, a great and stimulating working environment in which people are always open to help and share their knowledge to make everybody’s work better. Special thanks go to my fi'iends He'ctor, Clary and Xiao, who have always been and still are there for me. Thanks for all the scientific discussions and great suggestions, and also for the non-scientific suggestions and baby-tips, for all the encouragement and most of all for your invaluable fiiendship. Thanks to Steve and Gesche for introducing me to the world of denitrification in marine sediments and for great guidance and suggestions. Thanks also go to many other friends in and outside the Tiedje lab who have helped me in one or the other aspect of my work and most of all, who have made my stay in Michigan a wonderful experience. Thanks to Lizy, Blaz, Hanna, Kristian, Nid, Carmen, Claudia, Erick (thanks for the printing, picking me up, etc, etc), Carlos, Vincent, Alban, Ryan, Debora, Jorge, Olga, Lycely, Aviaja, Stephan, Tamara, and many, many others. A very special thank you goes also to Lisa, who has been not only a friend, but like a “big sister” away from home, helping me and my family in many, many ways that I can not be thankful enough for. In addition to Lisa, thank you also to Pat, Nicky, Darlene, Angie, and Suzanne for all their assistance. Thank you to Joyce, for all her help and her infinite kindness. Thanks to Lindsey and Chia Yu who have helped me in different parts of this research. Thanks to the RDP, especially to Benli, for always being available to answer questions and helping me with the sequences. Thanks to Christine at KBS for the help with the GC measurements. And a special thank you also to Dr. Helmut Bertrand and Dr. Robert Hausinger who as Directors of Graduate Studies have always been extremely kind and helpful. The list is long, but thanks to all people at MSU that have contributed in my education and in the research presented here. I am also very grateful towards the people at the University of Washington, especially in Dr. Devol’s laboratory and the Department of Chemical Oceanography, in addition to people from other institutions participating in the sampling cruises, that have helped me with the sample retrieval, microcosms experiments, and made my stay there a great experience, especially Gerard, Wendy, Cindy, Brooke, Ben, Heather, Linda, and Erin. Besides the people that have helped me directly in my research, there is a huge number of people that have helped me to reach this point by giving me their constant support and encouragement. I have to say thank you to my friends in Argentina, Switzerland and Germany, who have always been at my side, cheering me up all along the way. Thank you very much to all my family, for all their love and support. And a very special thank you goes to my mom and my dad, for always being there for me and believing in me, for encouraging me in all my decisions, even if that meant letting me go to do a Ph.D. far away from home, and most of all, for giving me all their love now and always. Danke fiir Alles, Mama und Papa, ich habe Euch sehr lieb! And now, last but definitely not least, I want to thank the two most important persons in my life, without whom none of this would have been possible, my husband Gustavo and my daughter Marina. Thank you for always being there for me, for helping me with everything you could, for encouraging me and supporting me all along the way, and most of all, thank you for your unconditional love. Gracias por todo mis amores, los quiero mucho como el cielo! Thanks to everybody for helping me make this dream a reality. vi PREFACE The work presented in this dissertation was part of a collaborative effort between Michigan State University, the University of Washington, Oak Ridge National Laboratory and the University of Puerto Rico, and funded by the Department of Energy. The Arctic samples analyzed were retrieved by Dr. Allan H. Devol, of the University of Washington. The Pacific Northwest samples were retrieved and processed also under Dr. Devol’s guidance. Dr. Devol’s laboratory was responsible for the biogeochemical analysis of all samples from the Arctic and the Pacific Northwest. The preliminary alignment of nirS gene sequences with the Hidden Markov Model aligner was performed by Dr. Benli Chai of the Ribosomal Database Project. The in silico digestion of nirS sequences with a custom made Perl script was performed by Dr. Héctor L. Ayala-del-Rio of MSU and latter the University of Puerto Rico — Humacao. vii TABLE OF CONTENTS LIST OF TABLES ............................................................................................................ xi LIST OF FIGURES ............................................................................... xiii CHAPTER 1 INTRODUCTION - -- - -- - - - - ..... - ----l OBJECTIVES .................................................................................................................... 11 Objective 1 ................................................................................................................. 11 Objective 2 ................................................................................................................. 12 Objective 3 ................................................................................................................. 13 REFERENCES ................................................................................................................... 14 CHAPTER 2 PSEUDOMONAS S T U TZERI NITRITE REDUCTASE GENE ABUNDANCE IN ENVIRONMENTAL SAMPLES MEASURED BY REAL-TIME PCR ................... 19 ABSTRACT ...................................................................................................................... 19 INTRODUCTION ............................................................................................................... 19 MATERIALS AND METHODS ............................................................................................ 22 Samples ...................................................................................................................... 22 Cultures ..................................................................................................................... 23 DNA extraction and quantitation .............................................................................. 24 Primers and probe ..................................................................................................... 25 Real-time PCR ........................................................................................................... 25 Specificity ................................................................................................................... 26 Sensitivity and detection limit .................................................................................... 27 Quantitation of nirS in environmental samples ......................................................... 28 Accuracy .................................................................................................................... 28 RESULTS ......................................................................................................................... 29 Specificity ................................................................................................................... 29 Sensitivity and detection limit .................................................................................... 30 Accuracy .................................................................................................................... 3] Analyses of environmental samples ........................................................................... 32 DISCUSSION .................................................................................................................... 44 Acknowledgements ..................................................................................................... 46 REFERENCES ................................................................................................................... 48 CHAPTER 3 DENITRIFIER COMMUNITY STRUCTURE AND ACTIVITY WITH RESPECT TO REDOX GRADIENTS AND BIOTURBATION-- - - -52 ABSTRACT ...................................................................................................................... 52 viii INTRODUCTION ............................................................................................................... 53 MATERIALS AND METHODS ............................................................................................ 55 Stuay area and sediment sampling ............................................................................ 55 Chemical analyses of sediment samples .................................................................... 56 DNA extraction and purification ............................................................................... 56 T —RF LP analysis of nirS gene .................................................................................... 5 7 Analysis of nirS T -RFLP profiles ............................................................................... 59 Cloning of nirS sequences ......................................................................................... 60 Phylogenetic analysis of cloned nirS sequences ........................................................ 61 Quantification of nirS from specific clusters and strains by real-time PCR ............. 62 Denitrification capacity in sediments ........................................................................ 64 RESULTS ......................................................................................................................... 65 Biogeochemical characteristics of the sediments ...................................................... 65 Denitrifier community structure by T -RFLP of nirS genes ....................................... 66 Phylogenetic analysis of nirS clone sequences .......................................................... 69 Quantification of clone clusters by real-time PCR .................................................... 71 Denitrification capacity of sediments studied in microcosms ................................... 73 DISCUSSION .................................................................................................................. 102 REFERENCES ................................................................................................................. 11 1 CHAPTER 4 DENITRIFIER AND BACTERIAL DIVERSITY AND DISTRIBUTION PATTERNS IN SEDIMENTS FROM THE ARCTIC AND THE PACIFIC NORTHWEST _ - -- - 116 ABSTRACT .................................................................................................................... 116 INTRODUCTION ............................................................................................................. 1 17 MATERIALS AND METHODS .......................................................................................... 119 Study area and sediment sampling .......................................................................... 119 Chemical analyses of sediment samples .................................................................. 120 DNA extraction and purification ............................................................................. 120 T-RFLP analysis of nirS gene .................................................................................. 120 Cloning of nirS and 16S rRNA gene sequences ....................................................... 122 Phylogenetic analysis of cloned nirS and 16S rRNA gene sequences ..................... 123 Community description and comparison based on 16S rRNA and nirS clone libraries ................................................................................................................................. 1 24 Quantification of nirS from specific clusters and strains by real-time PCR ........... 125 RESULTS ....................................................................................................................... 126 Biogeochemical characteristics of the sediments .................................................... 126 Denitrifier community structure by T -RF LP of nirS genes ..................................... 127 Phylogenetic analysis of nirS gene clone sequences ............................................... I30 Quantification of clone clusters by real-time PCR .................................................. 132 Phylogenetic analysis of 16S rRNA clone sequences .............................................. 134 Community description and comparison based on nirS and 16S rRNA clone libraries ................................................................................................................................. 13 7 DISCUSSION .................................................................................................................. 192 REFERENCES ................................................................................................................. 200 ix CHAPTER 5 CONCLUSIONS AND FUTURE PERSPECTIVES--. -- - _ ‘ 206 FUTURE PERSPECTIVES ................................................................................................. 210 REFERENCES ................................................................................................................. 213 APPENDIX RECOVERY OF NOVEL nirS GENES FROM NATURE AND TESTING OF THEIR FUNCTIONALITY - ...... - - - 215 REFERENCES ................................................................................................................. 217 LIST OF TABLES TABLE 2.1. PRIMER AND PROBE SEQUENCES COMPARED To EXAMPLE POSITIVE AND NEGATIVE CONTROL MRS GENE SEQUENCES ............................................................... 33 TABLE 2.2. SPECIFICITY OF REAL-TIME PCR REACTIONS FOR PSEUDOMONAS STUTZERI DNA ................................................................................................................................... 34 TABLE 2.3. P. STUTZERI MRS GENE ABUNDANCE IN VARIOUS HABITATS. ............................ 42 TABLE 3.1. LOCATIONS AND CHARACTERISTICS OF SAMPLING STATIONS. .......................... 78 TABLE 3.2. SEDIMENT SAMPLES ANALYZED FROM DIFFERENT SITES. ................................. 79 TABLE 3.3. DESCRIPTION OF CLUSTERS IDENTIFIED IN MRS PHYLOGENETIC TREE. ............. 91 TABLE 3.4. PRIMERS AND PROBES USED To QUANTIFY THE MRS GENE OF SPECIFIC CLUSTERS AND STRAINS BY REAL-TIME PCR ............................................................................... 92 TABLE 3.5. NITROUS OXIDE PRODUCTION RATES IN MICROCOSMS AMENDED WITH 400 MM NANO3. ...................................................................................................................... 99 TABLE 4.1. LOCATIONS AND CHARACTERISTICS OF SAMPLING STATIONS. ........................ 142 TABLE 4.2. SEDIMENT SAMPLES ANALYZED FROM DIFFERENT SITES. ............................... 143 TABLE 4.3. CLUSTER SPECIFIC AND P. STUYZERI MRS GENE ABUNDANCE IN PACIFIC NORTHWEST AND ARCTIC SEDIMENTS As MEASURED BY REAL-TIME PCR. .............. 168 TABLE 4.4. CLUSTER SPECIFIC AND P. STUTZER! MRS GENE ABUNDANCE IN PACIFIC NORTHWEST SEDIMENTS AS MEASURED BY REAL-TIME PCR AND DETERMINATION OF SIGNIFICANT DIFFERENCES BY ANOVA ................................................................... 169 TABLE 4.5. ASSIGNMENT OF 168 RRNA CLONE SEQUENCES FROM PACIFIC NORTHWEST AND ARCTIC SEDIMENTS INTO TAXONOMIC GROUPS ................................................. 184 TABLE 4.6.SIGNIFICANT DIFFERENCES IN THE TAXONOMIC COMPOSITION BETWEEN 16S RRNA LIBRARIES FROM PACIFIC NORTHWEST AND ARCTIC SEDIMENTS .................. 185 TABLE 4.7. DIVERSITY AND PREDICTED RICHNESS OF 168 RRNA AND MRS GENE FRAGMENTS IN SEDIMENTS FROM DIFFERENT SITES BASED ON OTU ASSIGNMENT OF CLONES BY DOTUR. ................................................................................................ 189 xi TABLE 4.8. DETERMINATION OF SIGNIFICANT DIFFERENCES BETWEEN LIBRARIES BY THE APPLICATION OF THE INTEGRAL FORM OF THE CRAMER-VON MISES STATISTIC WITH I- LIBSHUFF. ............................................................................................................. 190 TABLE 4.9. SIMILARITY BETWEEN LIBRARIES DESCRIBED BY THE ABUNDANCE-BASED SORENSON SIMILARITY INDEX (LAM/ND) AND WITH A NONPARAMETRIC RICHNESS ESTIMATOR OF SHARED OTUS ANALOGOUS TO CHAol (S AB CHAO) As DETERMINED WITH SONS. ............................................................................................................. 191 xii LIST OF FIGURES Some of the figures in this dissertation are presented in color. FIGURE 1.1. DIAGRAM OF THE NITROGEN CYCLE INDICATING PROCESSES AND NITROGEN COMPOUNDS INVOLVED. ............................................................................................... 2 FIGURE 1.2. DENITRIFICATION PATHWAY SHOWING THE NITROGEN SPECIES AND ENZYMES INVOLVED. .................................................................................................................... 6 FIGURE 2.1. GENERATION OF STANDARD CURVE. ............................................................... 35 FIGURE 2.2. LIMIT OF DETECTION OF THE P. STUTZERI MRS GENE USING REAL-TIME PCR. .37 FIGURE 2.3. RELATIONSHIP BETWEEN THRESHOLD CYCLE (CT) AND ERROR. ...................... 39 FIGURE 2.4. QUANTIFICATION OF MRS IN ARTIFICIAL MIXTURES OF P. STUTZERI KC, P. AERUGINOSA, AND E. cou CELLS. ............................................................................... 40 FIGURE 2.5. QUANTIFICATION OF MRS IN KBS SOIL (0) AND WINTERGREEN LAKE SEDIMENT (A) SAMPLES SPIKED WITH P. STUTZERI CELLS. ......................................... 41 FIGURE 2.6. P. STUTZERI MRS COPY NUMBER MEASURED BY REAL-TIME PCR WITHOUT (BLACK BARs) AND WITH (GRAY BARS) THE ADDITION OF 106 P. STUTZERI MRS COPY NUMBERS T0 DNA EXTRACTED FROM DIFFERENT ENVIRONMENTAL SAMPLES. .......... 42 FIGURE 3.1. LOCATIONS OF SAMPLING STATIONS AT PUGET SOUND AND THE WASHINGTON MARGIN SLOPE. ........................................................................................................... 77 FIGURE 3.2. PORE WATER PROFILES FOR 02 AND NO3' FOR PUGET SOUND (SHALLOW BUDD INLET, TURNING BASIN, AND CARR INLET) AND WASHINGTON MARGIN SEDIMENTS. 80 FIGURE 3.3. PORE WATER PROFILES FOR FE(II), SO42} MN(II), AND NH4+ FOR PUGET SOUND (SHALLOW BUDD INLET, TURNING BASIN, AND CARR INLET) AND WASHINGTON MARGIN SEDIMENTS. ........................................................................... 82 FIGURE 3.4. T-RFLP PROFILES OBTAINED BY AMPLIFICATION OF MRS GENES FROM PUGET SOUND AND WASHINGTON MARGIN MARINE SEDIMENTS FROM VARIOUS DEPTHS ...... 84 FIGURE 3.5. DENDROGRAM OBTAINED BY CLUSTER ANALYSIS OF T-RFLP PROFILES ......... 85 FIGURE 3.6. PCA ORDINATION OF T-RFLPS OF MRS GENES FROM PUGET SOUND AND WASHINGTON MARGIN SEDIMENTS. ........................................................................... 86 xiii FIGURE 3.7. PHYLOGENETIC TREE SHOWING THE AFFILIATION OF MRS CLONE SEQUENCES RETRIEVED FROM SHALLOW BUDD INLET, PUGET SOUND, SEDIMENTS TO SELECTED REFERENCE SEQUENCES. ............................................................................................. 87 FIGURE 3.8. DETAIL OF SUBTREE A AS DESCRIBED IN FIGURE 3.7. ..................................... 88 FIGURE 3.9. DETAIL OF SUBTREE B As DESCRIBED IN FIGURE 3.7. ..................................... 89 FIGURE 3.10. COMPARISON OF T-RFLP PROFILES OBTAINED BY AMPLIFICATION OF MRS GENES FROM SHALLOW BUDD INLET, PUGET SOUND, SEDIMENTS FROM VARIOUS DEPTHS To DISTRIBUTION OF T-RF SIZES OBTAINED BY IN sruco DIGESTION OF MRS CLONES RETRIEVED FROM THAT SAME SITE (SEDIMENT DEPTH 2-2.5 CM). .................. 90 FIGURE 3.1 1. CLUSTER SPECIFIC AND P. STUTZERI MRS GENE ABUNDANCE IN SHALLOW BUDD INLET AND TURNING BASIN SEDIMENTS FROM DIFFERENT DEPTHS As MEASURED BY REAL-TIME PCR. ................................................................................................... 93 FIGURE 3.12. CLUSTER SPECIFIC AND P. STUTZERI MRS GENE ABUNDANCE IN CARR INLET AND WASHINGTON MARGIN SEDIMENTS FROM DIFFERENT DEPTHS AS MEASURED BY REAL-TIME PCR .......................................................................................................... 94 FIGURE 3.13. NITROUS OXIDE PRODUCTION IN MICROCOSMS OF SHALLOW BUDD INLET (PUGET SOUND) SEDIMENTS FROM VARIOUS DEPTHS AFTER ADDITION OF DIFFERENT ELECTRON ACCEPTORS ................................................................................................ 95 FIGURE 3.14. NITROUS OXIDE PRODUCTION IN MICROCOSMS OF CARR INLET (PUGET SOUND) SEDIMENTS FROM VARIOUS DEPTHS AFTER ADDITION OF DIFFERENT ELECTRON ACCEPTORS. ................................................................................................................ 96 FIGURE 3.15. NITROUS OXIDE PRODUCTION IN MICROCOSMS OF TURNING BASIN (PUGET SOUND) SEDIMENTS FROM VARIOUS DEPTHS AFTER ADDITION OF DIFFERENT ELECTRON ACCEPTORS. ................................................................................................................ 97 FIGURE 3.16. NITROUS OXIDE PRODUCTION IN MICROCOSMS OF WASHINGTON MARGIN SEDIMENTS FROM VARIOUS DEPTHS AFTER ADDITION OF DIFFERENT ELECTRON ACCEPTORS. ................................................................................................................ 98 FIGURE 3.17. T-RFLP PROFILES OBTAINED BY AMPLIFICATION OF MRS GENES FROM SHALLOW BUDD INLET, PUGET SOUND, SEDIMENTS COMPARED To PROFILES OBTAINED AFTER INCUBATION OF SEDIMENTS FROM THE SAME SITE WITH 400 MM NANO3 IN MICROCOSMS. .......................................................................................... 100 FIGURE 3.18. PCA ORDINATION OF T-RFLPS OF MRS GENES FROM SHALLOW BUDD INLET, PUGET SOUND, SEDIMENTS NOT SUBJECTED AND SUBJECTED To INCUBATION WITH 400 MM NANO3 IN MICROCOSMS. ................................................................................... 101 xiv FIGURE 4.1. LOCATIONS OF SAMPLING STATIONS. ............................................................ 141 FIGURE 4.2. PORE WATER PROFILES FOR 02 AND NO;' IN ARCTIC (BARROW CANYON SHALLOW, BARROW CANYON DEEP, EAST HANNA SHOAL SHALLOW, AND EAST HANNA SHOAL DEEP) SEDIMENTS. ........................................................................... 144 FIGURE 4.3. PORE WATER PROFILES FOR FE(II), MN(II), AND NH4+ IN ARCTIC (BARROW CANYON SHALLOW, BARROW CANYON DEEP, EAST HANNA SHOAL SHALLOW, AND EAST HANNA SHOAL DEEP) SEDIMENTS. .................................................................. 146 FIGURE 4.4. PORE WATER PROFILES FOR 02, N03] AND NH4+ IN SEDIMENTS FROM THE ABYSSAL SEA FLOOR WEST OF THE JUAN DE FUCA RIDGE. ........................................ 148 FIGURE 4.5. T-RFLP PROFILES OBTAINED BY AMPLIFICATION OF MRS GENES FROM ARCTIC (EHS, EAST HANNA SHOAL SHALLOW; EHSD, EAST HANNA SHOAL DEEP; BC, BARROW CANYON SHALLOW; BCD BARROW CANYON DEEP) AND PACIFIC NORTHWEST (WJF, WEST OF JUAN DE FUCA; WM, WASHINGTON MARGIN; CI, CARR INLET; SB, SHALLOW BUDD INLET; TB, TURNING BASIN) MARINE SEDIMENTS FROM VARIOUS DEPTHS. ..................................................................................................... 149 FIGURE 4.6A. RICHNESS OF THE MRS-CONTAINING COMMUNITY IN PACIFIC NORTHWEST (TB, TURNING BASIN; SB, SHALLOW BUDD INLET; CI, CARR INLET, WM, WASHINGTON MARGIN; WJ F, WEST OF JUAN DE FUCA) AND ARCTIC (BC, BARROW CANYON SHALLOW; BCD, BARROW CANYON DEEP; EHS, EAST HANNA SHOAL SHALLOW; EHSD, EAST HANNA SHOAL DEEP) SEDIMENTS FROM DIFFERENT DEPTHS, BASED ON NUMBER OF INDIVIDUAL T-RFS DETECTED AFTER RESTRICTION WITH HHAI. ................................................................................................................................. 150 FIGURE 4.6B. DIVERSITY OF THE MRS-CONTAINING COMMUNITY IN PACIFIC NORTHWEST (TB, TURNING BASIN; SB, SHALLOW BUDD INLET; CI, CARR INLET, WM, WASHINGTON MARGIN; WJ F, WEST OF JUAN DE FUCA) AND ARCTIC (BC, BARROW CANYON SHALLOW; BCD, BARROW CANYON DEEP; EHS, EAST HANNA SHOAL SHALLOW; EHSD, EAST HANNA SHOAL DEEP) SEDIMENTS FROM DIFFERENT DEPTHS, AS DETERMINED BY THE SHANNON DIVERSITY INDEX BASED ON THE T-RFLP RESULTS. THE T-RFLP ANALYSIS WAS PERFORMED WITH RESTRICTION ENZYME HHAI .............. 151 FIGURE 4.7. DENDROGRAM OBTAINED BY CLUSTER ANALYSIS OF T-RFLP PROFILES ....... 152 FIGURE 4.8. PCA ORDINATION OF T-RFLPS OF MRS GENES FROM PACIFIC NORTHWEST (WM, WASHINGTON MARGIN; TB, TURNING BASIN; SB, SHALLOW BUDD INLET; CI, CARR INLET; WJF, WEST OF JUAN DE FUCA) AND ARCTIC (BCD, BARROW CANYON DEEP; BC, BARROW CANYON SHALLOW; EHSD, EAST HANNA SHOAL DEEP; EHS, EAST HANNA SHOAL SHALLow) SEDIMENTS ............................................................ 153 FIGURE 4.9. CANONICAL CORRESPONDENCE ANALYSIS (CCA) BASED ON T-RFLPS OF MRS GENES AND ENVIRONMENTAL VARIABLES FROM PACIFIC NORTHWEST (TB, TURNING XV BASIN; SB, SHALLOW BUDD INLET; CI, CARR INLET, WM, WASHINGTON MARGIN; WJF, WEST OF JUAN DE FUCA) AND ARCTIC (BC, BARROW CANYON SHALLOW; BCD, BARROW CANYON DEEP; EHS, EAST HANNA SHOAL SHALLOW; EHSD, EAST HANNA SHOAL DEEP) SEDIMENTS. ........................................................................................ 154 FIGURE 4.10. SCHEMATIC REPRESENTATION OF THE MRS PHYLOGENETIC TREE SHOWN AND DESCRIBED IN FURTHER DETAIL IN FIGURE 4.11. ...................................................... 155 FIGURE 4.1 1. PHYLOGENETIC TREE SHOWING THE AFFILIATION OF MRS CLONE SEQUENCES RETRIEVED FROM SHALLOW BUDD INLET (SB), WASHINGTON MARGIN (WM), WEST OF JUAN DE FUCA (WJF), BARROW CANYON SHALLOW (BC), AND EAST HANNA SHOAL SHALLOW (EHS) SEDIMENTS To SELECTED REFERENCE SEQUENCES. ........... 156 FIGURE 4.12. DETAIL OF SUBTREE A AS DESCRIBED IN FIGURES 4.10. AND 4.11. ............. 158 FIGURE 4.13. DETAIL OF SUBTREE B As DESCRIBED IN FIGURES 4.10. AND 4.1 1. ............. 160 FIGURE 4.14. DETAIL OF SUBTREE C AS DESCRIBED IN FIGURES 4.10. AND 4.11. ............. 162 FIGURE 4.15. DETAIL OF SUBTREE D As DESCRIBED IN FIGURES 4.10. AND 4.11. ............. 163 FIGURE 4.16. DETAIL OF SUBTREE E AS DESCRIBED IN FIGURES 4.10. AND 4.11. ............. 165 FIGURE 4.17. COMPARISON OF T-RFLP PROFILES OBTAINED BY AMPLIFICATION OF MRS GENES FROM ARCTIC (EHS, EAST HANNA SHOAL SHALLOW; BC, BARROW CANYON SHALLOW) AND PACIFIC NORTHWEST (WJF, WEST OF JUAN DE FUCA; WM, WASHINGTON MARGIN; SB, SHALLOW BUDD INLET) SEDIMENTS FROM VARIOUS DEPTHS To DISTRIBUTION OF T-RF SIZES OBTAINED BY IN SILICO DIGESTION OF MRS CLONES RETRIEVED FROM THOSE SAME SITES. .......................................................... 167 FIGURE 4.18. PHYLOGENETIC TREE SHOWING THE AFFILIATION OF 16S RRNA CLONE SEQUENCES FROM THE GAMMA PROTEOBACTERIA RETRIEVED FROM SHALLOW BUDD INLET (SB), WASHINGTON MARGIN (WM), WEST OF JUAN DE FUCA (WJF), BARROW CANYON SHALLOW (BC), AND EAST HANNA SHOAL SHALLOW (EHS) SEDIMENTS To SELECTED REFERENCE SEQUENCES. .......................................................................... 170 FIGURE 4.19. PHYLOGENETIC TREE SHOWING THE AFFILIATION OF 16S RRNA CLONE SEQUENCES FROM THE DELTA PROTEOBACTERIA RETRIEVED FROM SHALLOW BUDD INLET (SB), WASHINGTON MARGIN (WM), WEST OF JUAN DE FUCA (WJF), BARROW CANYON SHALLOW (BC), AND EAST HANNA SHOAL SHALLOW (EHS) SEDIMENTS TO SELECTED REFERENCE SEQUENCES. .......................................................................... 172 FIGURE 4.20. PHYLOGENETIC TREE SHOWING THE AFFILIATION OF 16S RRNA CLONE SEQUENCES FROM THE ALPHA, BETA, AND EPSILON PROTEOBACTERIA RETRIEVED FROM SHALLOW BUDD INLET (SB), WASHINGTON MARGIN (WM), WEST OF JUAN DE FUCA xvi (WJF), BARROW CANYON SHALLOW (BC), AND EAST HANNA SHOAL SHALLOW (EHS) SEDIMENTS To SELECTED REFERENCE SEQUENCES. ........................................ 174 FIGURE 4.21. PHYLOGENETIC TREE SHOWING THE AFFILIATION OF 168 RRNA CLONE SEQUENCES FROM THE BA CTEROIDETES, CHLOROBI, DEFERRIBACTERES, SPIROCHAETES, AND ACIDOBACTERIA RETRIEVED FROM SHALLOW BUDD INLET (SB), WASHINGTON MARGIN (WM), WEST OF JUAN DE FUCA (WJF), BARROW CANYON SHALLOW (BC), AND EAST HANNA SHOAL SHALLOW (EHS) SEDIMENTS To SELECTED REFERENCE SEQUENCES. .............................................................................................................. 176 FIGURE 4.22. PHYLOGENETIC TREE SHOWING THE AFFILIATION OF 16S RRNA CLONE SEQUENCES FROM THE CHLOROFLEXI, ACTINOBACTERJA, NITROSPIRA, ODl GROUP AND OPl 1 GROUP RETRIEVED FROM SHALLOW BUDD INLET (SB), WASHINGTON MARGIN (WM), WEST OF JUAN DE FUCA (WJF), BARROW CANYON SHALLOW (BC), AND EAST HANNA SHOAL SHALLOW (EHS) SEDIMENTS To SELECTED REFERENCE SEQUENCES. ................................................................................................................................. 178 FIGURE 4.23. PHYLOGENETIC TREE SHOWING THE AFFILIATION OF 16S RRNA CLONE SEQUENCES FROM THE FIRM/CUTES, FUSOBA CTERIA, AND FIBROBACTERES RETRIEVED FROM SHALLOW BUDD INLET (SB), WASHINGTON MARGIN (WM), WEST OF JUAN DE FUCA (WJF), BARROW CANYON SHALLOW (BC), AND EAST HANNA SHOAL SHALLOW (EHS) SEDIMENTS To SELECTED REFERENCE SEQUENCES. ........................................ 180 FIGURE 4.24. PHYLOGENETIC TREE SHOWING THE AFFILIATION OF 16S RRNA CLONE SEQUENCES FROM THE PLANCTOMYCETES, VERRUCOMICROBIA AND WS3 GROUP RETRIEVED FROM SHALLOW BUDD INLET (SB), WASHINGTON MARGIN (WM), WEST OF JUAN DE FUCA (WJF), BARROW CANYON SHALLOW (BC), AND EAST HANNA SHOAL SHALLOW (EHS) SEDIMENTS To SELECTED REFERENCE SEQUENCES. ........... 181 FIGURE 4.25. PHYLOGENETIC TREE SHOWING THE AFFILIATION OF 168 RRNA CLONE SEQUENCES FROM THE CYANOBACTERIA RETRIEVED FROM SHALLOW BUDD INLET (SB), WASHINGTON MARGIN (WM), WEST OF JUAN DE FUCA (WJF), BARROW CANYON SHALLOW (BC), AND EAST HANNA SHOAL SHALLOW (EHS) SEDIMENTS To SELECTED REFERENCE SEQUENCES. ........................................................................................... 183 FIGURE 4.26. RELATIONSHIP BETWEEN THE NUMBER OF OTUS AND THEIR ABUNDANCE, AS MEASURED BY THE NUMBER OF CLONES CORRESPONDING To EACH OTU, FOR 16S RRNA AND MRS CLONE LIBRARIES FROM BARROW CANYON SHALLOW (BC), EAST HANNA SHOAL SHALLOW (EHS), SHALLOW BUDD INLET (SB), WASHINGTON MARGIN (WM), AND WEST OF JUAN DE FUCA (WJF) SEDIMENTS ............................ 187 FIGURE 4.27. RAREFACTION CURVES FOR 16S RRNA AND MRS CLONE LIBRARIES FROM BARROW CANYON SHALLOW (BC), EAST HANNA SHOAL SHALLOW (EHS), SHALLOW BUDD INLET (SB), WASHINGTON MARGIN (WM), AND WEST OF JUAN DE FUCA (WJF) SEDIMENTS. .............................................................................................................. 188 xvii CHAPTER 1 INTRODUCTION Nitrogen is a component of fundamental building blocks of living organisms, such as nucleic and amino acids. Nitrogenous compounds experience multiple transformations, catalyzed primarily by microbes, which constitute the N cycle (Figure 1.1). One of the principal biotic transformations of nitrogen compounds within the N cycle is denitrification, the dissimilatory reduction of nitrate or nitrite to gaseous products, principally N2 and N20, with the concomitant conservation of energy (Tiedje, 1994). Denitrification is basically a bacterial respiratory process, although it is also found in Archaea and in mitochondria of certain fungi (Zumft, 1997). Within the bacteria, denitrifiers are found in diverse phylogenetic groups, although they are notably absent from the enterobacteria (Zumft, 1992). They are also Spread among diverse physiological groups, including organotrophs (organic energy source), lithotrophs (inorganic energy source), and phototrophs (light as energy source), although the most common energy sources are organic substrates. Most denitrifiers are aerobic bacteria, which have the alternative capacity to use N oxides as electron acceptors, when oxygen becomes limiting (Tiedje, 1988). Nitrification NO2 Assimilation - - - N K, NH, groups Assnmllatlon 2 _ °f prawns Nitrogen fixation / Deamination Aerobic Assimilatma— NH2 groups of proteins Deaminatio I Anaerobic Denitrification Figure 1.1. Diagram of the nitrogen cycle indicating processes and nitrogen compounds involved. Chemodenitrification (abiotic conversion of N03' to N2 by reaction with inorganic species), chemoxidation (abiotic conversion of NH; to N2 by reaction with inorganic species), and OLAND (oxygen-limited autotrophic nitrification-denitrification) (Brandes et al., 2007) have not been included for clarity. Denitrification is the main sink for fixed nitrogen and, therefore, a fundamental biogeochemical process (Knowles, 1982). Although its main significance is the completion of the nitrogen cycle, it also is of interest, from an anthropogenic point of View. First, it accounts generally for 20 to 70% of fertilizer losses, which is of concern since nitrogen is the most limiting nutrient to crop production. On the other hand, it is helpful in removing excess combined nitrogen in waste treatment and N-enriched coastal areas. Two of its intermediate products, N20 and NO, have been recognized to play a role in the destruction of the ozone layer and contributing to global warming, while the latter and NO2', are toxic and could produce local hazards (see Tiedje, 1988 and references therein). Denitrification in marine sediments is a major sink for nitrogen in the global nitrogen budget and is the main sink of oceanic fixed nitrogen (Codispoti, 1995). The continental margins account for half of the nitrogen cycling taking place in the oceans, although their area represents less than 20% of the world oceans (Walsh, 1991). One of the major Sites of oceanic nitrogen cycling are marine sediments, of which the Washington margin, in the Pacific Ocean, is a good example. It is a high productivity area, due to wind-driven coastal upwelling (Hickey, 1989), with a large flux of organic matter to the sediments, which results in very high denitrification rates of 3.7 pmol N cm' 2 s'1 in average (Devol, 1991). Growing attention has been given to the Arctic Ocean, with 50% of its area consisting of shelf seas, which account for about 25% of the world continental Shelves. High productivity has been detected in the Chukchi Sea in that area, related especially with the northward flow of water through the Bering Strait, driven by sea level difference (Woodgate et al., 2005). This area also presents high denitrification, with rates of the highly productive Barrow Canyon area in the Chukchi Sea comparable to those of the Washington margin (Devol et al., 2005). Although denitrification rates in deep sea sediments are significantly lower as compared to those in continental margin sediments, these rates have been revised upwards (Middelburg et al., 1996) and denitrifiers have been detected and isolated from this environment (Tamegai et al., 2007; Tarnegai et al., 1997), suggesting its importance for further study. The present day marine nitrogen budget seems to be unbalanced (Codispoti, 1995; Middelburg et al., 1996). The input of nitrogen to the ocean by atmospheric precipitation, riverine input and in situ nitrogen fixation seems tO be smaller than its removal by sedimentary denitrification, denitrification in the water column, permanent burial in sediments and organic nitrogen exports (Codispoti, 1995; Middelburg et al., 1996). We are currently in an interglacial period when higher denitrification rates are thought to occur than in glacial periods (Ganeshram et al., 1995), therefore supporting the possibility of an unbalanced nitrogen budget. One consequence would be a loss of fixed nitrogen by the ocean that could lead to lower photosynthesis and consequently to a lower oceanic uptake of carbon dioxide. This, in turn, would result in a higher carbon dioxide concentration in the atmosphere, which would enhance global warming (Codispoti, 1995; Middelburg et al., 1996). Despite the above argument, the nitrogen budget could still be in balance, as the estimates of sedimentary denitrification vary by more than an order of magnitude (Devol, 1991; Middelburg et al., 1996). This variability is due to extrapolation from different Site-Specific estimates and to varying results when direct or indirect measurements are used for estimation of global sedimentary denitrification (Devol, 1991). In order to accurately model the oceanic nitrogen budget, the range in sedimentary denitrification estimates needs to be narrowed. Better information on the distribution and diversity of denitrifiers and on the response of their activities to changes in environmental factors may provide more fundamental insight into interpreting and predicting denitrification activities and therefore help refine these estimates. On the other hand, other processes leading to loss of fixed nitrogen from sediments and water column have been recently described (Brandes et al., 2007), including chemodenitrification (Luther III et al., 1997) and anaerobic ammonium oxidation (anammox) (van de Graaf et al., 1995; Dalsgaard et al., 2005). It has been suggested that chemodenitrification, the reduction of nitrate to N2 by Mn”, could account for up to 90% of the N2 formed in continental margin sediments (Luther III et al., 1997), while anammox, the anaerobic ammonium oxidation with nitrite leading to the formation of N2, may account for 30 to 50% of the N2 produced in the oceans (Devol, 2003). Furthermore, 16S rRNA sequences associated to known anammox bacteria have been detected in a variety of marine sediments, in addition to other environments, suggesting that this process might be more widespread than previously thought (Penton et al., 2006). Dissimilatory nitrate reduction to ammonium (DNRA) is also widespread and more important than denitrification in environments with high carbon to electron acceptor ratios (Tiedje, 1994), however, it does not constitute a loss term for the marine nitrogen budget, as nitrate is reduced to ammonium, which stays in the system. The discovery of these additional processes leading to nitrogen loss could suggest an even further unbalanced marine nitrogen budget than suggested by the increase in marine denitrification estimations alone. However, nitrogen fixation estimates have also increased and this process seems to be more widespread than previously thought, therefore leaving open the possibility for a balanced marine nitrogen budget (Brandes et al., 2007). Denitrification consists of four reaction steps, each catalyzed by one or two different enzymes (Zumft, 1997; Philippot, 2002) (Figure 1.2). N03; —> N02- —> NO —> N20 ——> N2 Nar Nir Nor Nos Figure 1.2. Denitrification pathway Showing the nitrogen species and enzymes involved. Nar, nitrate reductase; Nir, nitrite reductase; Nor, nitric oxide reductase; Nos, nitrous oxide reductase. The key enzyme in denitrification is nitrite reductase, which catalyzes the reduction of nitrite to nitric oxide, because it is the first step where fixed nitrogen is lost to the biosphere (Zumft, 1997). For this reason, this enzymatic step has been the focus of physiological and biochemical studies on denitrification and several probes and antibodies have been designed to target the nitrite reductases (Ward, 1996; Bothe et al., 2000). There are two variants of the dissimilatory nitrite reductase. One contains copper and is coded by the gene nirK while the other type contains hemes c and d, and is coded by gene nirS (Zumft, 1997). Although normally these two nitrite reductases are mutually exclusive, the existence of both genes, nirS and nirK, has been detected in the genome of Paracoccus denitrificans Pd1222 (Hallin and Lindgren, 1999). Heme cd, nitrite reductase is a homodimer with a subunit mass of around 60 kDa. It is a tetraheme protein as each subtmit has a heme c and a heme d1. Copper—containing nitrite reductase is a trimer with three identical subunits of close to 40 kDa each and Cu as prosthetic metal (Zumft, 1997). Although these enzymes are functionally equivalent, they are structurally different and can therefore not be targeted with a general probe against both of them. Still, the transfer of the nirK gene from Pseudomonas aureofaciens to a Pseudomonas stutzeri strain with a mutated nirS gene led to the expression of a fimctional copper-containing nitrite reductase (Glockner et al., 1993). The heme cd, type seems to be numerically dominant in nature, while the copper type is present in a wider range of physiological groups, including, among others Gram-positive Spore formers, nitrogen fixing bacteria, and Archaea (Coyne et al., 1989; Coyne and Tiedje, 1990; Goregues etal., 2005). The taxonomic diversity of denitrifiers makes rRNA-based approaches to study the diversity and distribution of this functional guild inappropriate. Instead, the genes encoding the key enzymes involved in this pathway have been targeted in many studies as molecular markers. Although primers have been designed for genes encoding the nitrate (Flanagan et al., 1999; Cheneby et al., 2003), nitric oxide (Braker and Tiedje, 2003), and nitrous oxide (Scala and Kerkhof, 1998, Throback et al., 2004) reductases, the first and most targeted genes from the pathway have been the ones encoding the nitrite reductases (Braker et al., 1998; Hallin and Lindgren, 1999, Throback et al., 2004). Although nirK appears in bacteria of deeper systematic affiliation than nirS, it is more conserved and design of broad-range primers is somewhat easier (Bothe et al., 2000). However, environmental studies have often not been able to detect nirK or, as expected, detected a lower diversity as for nirS genes, especially in marine and other aquatic samples (Braker et al., 2000; Tarnegai et al., 2007, Santoro et a1, 2006; Nogales et al., 2002). Soil studies, on the other hand, Showed similar results for marshland soils, presenting a higher nirS than nirK diversity, however, nirS was not detected in upland soils, opposite to the general observation in aquatic environments (Priemé et al., 2002). This could be related to differential amplification efficiencies of the primer sets used or indicate a real difference in the distribution of these two genes in the environment (Priemé et al., 2002). Therefore, most denitrifier diversity studies in aquatic environments have focused on nirS, rather than nirK genes (Braker et al., 2001; Castro- Gonzalez et al., 2005; Hannig et al., 2006; Jayakumar et al., 2004). Regardless of the denitrification gene targeted, the denitrifier diversity seems to be extremely high in estuarine and continental Shelf marine sediments (Braker et al., 2000; Braker et al., 2001; Liu et al., 2003; Scala and Kerkhof, 1999; Tiquia et al., 2006), and lower in the water column (Hannig et al., 2006; Castro-Gonzalez et al., 2005; Oakley et al., 2007). However, all these denitrifier diversity studies reveal the presence of a large number of denitrification gene sequences unrelated to cultured denitrifiers. Environmental denitrification gene clone sequences form major clusters, generally including clone sequences from other environments, but with no overlap with isolated strains (Braker et al., 2000; Hannig et al., 2006; Jayakumar et al., 2004; Liu et al., 2003; Scala and Kerkhof, 1999; Oakley et al., 2007). On the other hand, isolation of denitrifiers from sediments and water column generally leads to the cultivation of known denitrifiers, such as Marinomonas Sp., Pseudomonas Sp. (often Pseudomonas stutzeri), and Halomonas Sp. (Braker et al., 2000; Oakley et al., 2007; Goregues et al., 2005), not closely related to the novel sequences found in the environment. Although only isolation of a denitrifier with one of the novel sequences would be a definite proof of its functionality, expression of some of these novel denitrification gene sequences has been detected in estuarine sediments from the river Colne for narG, napA, nirS, and nosZ (Nogales et al., 2002; Smith et al., 2007), suggesting that they might belong to active denitrifiers. Therefore, molecular methods are more appropriate for studying the overall diversity of denitrification genes. Several molecular approaches have been taken to detect denitrifier diversity in the environment (Philippot and Hallin, 2005). Most studies aiming to describe diversity have included the construction of clone libraries of the gene of interest. This method exhibits the highest resolution, allowing the detection of individual nucleotide differences between gene sequences. However, clone libraries generally only contain a limited ntunber of gene sequences, often representing only a fraction of the diversity present, as pristine marine sediments have been estimated to contain ~ 11,000 different genome equivalents (Torsvik et al., 2002). Therefore, to explore the general diversity of a community, based on the diversity of a specific gene, a fingerprinting technique, such as terminal restriction fragment length polymorphism (T-RFLP) (Liu et al., 1997) is often applied (Braker et al., 2001; Castro-Gonzalez et al., 2005; Hannig et al.; Scala and Kerkhof, 2000). This allows the fingerprinting of a community and the rapid comparison between different sites and samples, however, at a lower resolution than cloning of the genes. The full understanding of the distribution of denitrifiers in the environment requires, besides the description of their diversity, the determination of their abundance. Several PCR-based methods, such as most-probable-number and competitive PCR, have been used to quantify nirS (Michotey et al., 2000) and nirK (Qiu et al., 2004) from the environment. However, in recent years, real-time PCR (Heid et al., 1996) has been more widely used than these other techniques, due to its Simplicity and higher accuracy as it focuses on the logarithmic phase of product accumulation rather than the end product. Therefore, several denitrification genes, including narG, napA, nirS, nirK, and nosZ have been quantified in the environment using this technique (Smith et al., 2007; Henry et al., 2004; Henry et al., 2006; LOpez—Gutiérrez et al., 2004). Denitrifier distribution in marine sediments follows specific horizontal and vertical biogeographic patterns. A conserved denitrifier community structure has been Observed along the vertical redox potential gradient in marine sediments from the Pacific Northwest (Braker et al., 2001). Although nitrate and oxygen were present only in the sediment top 1 cm, the denitrifier community structure was conserved throughout the analyzed depth of 6.5 and 10 cm in Puget Sound and Washington continental shelf sediments, respectively. Furthermore, nirS genes have been cloned from these same 6.5 cm deep Puget Sound sediments (Braker et al., 2000). Bioturbation of the sediments by marine invertebrates has been suggested as an explanation for the presence of denitrification genes, which might correspond to functional denitrifiers, in sediments below the penetration depths of oxygen and nitrate, however, no sediments below the bioturbated zone had been analyzed in order to evaluate this hypothesis. On the other hand, large variations in denitrifier community structure have been detected on a horizontal scale, which increases in correlation with increasing geographic distance (Scala and Kerkhof, 2000). Puget Sound and Washington continental shelf sediments also exhibited differential nirS-based community structures at different Sites (Braker et al, 10 2000; 2001). These differences were more significant between these two geographic locations, than between sediments from different water depths at the continental Shelf. Different sediment depths, on the other hand, exhibited the lowest differentiation, as mentioned earlier. Increased knowledge on the distribution, activity and types of denitrifiers may help refine denitrification estimates and aid the better modeling of the marine N cycle. Objectives Based on the above rationale, the following objectives and hypothesis are proposed: Objective 1 To develop a real-time PCR assay for the quantification of Pseudomonas stutzeri nirS and evaluate its applicability in the environment. In order to better model denitrification in the environment, the quantification of the catalyst is helpful. Real-time PCR is a simple and accurate quantification method, which has been widely applied in the medical field, but only recently started to be used on environmental samples. Therefore, it needed to be tested for its applicability in the environment, in particular for the detection of a functional gene. P. stutzeri, a commonly isolated denitrifier from soil (Gamble et al., 1977) and marine environments (Ward and Cockcroft, 1993; Braker et al., 2000), may play a Significant role in global denitrification. The determination of its abundance in the environment will help elucidate its importance. 11 Objective 2 To elucidate the distribution and activity of denitrifiers in marine sediments, with respect to the bioturbation zones and redox gradients. As described earlier, the vertical distribution of denitrification genes in marine sediments exhibited a conserved community structure within the bioturbated layer, and the mixing of the sediments by marine invertebrates had been suggested as an explanation for this observation (Braker et al., 2001). However, sediments reaching below the mixed layer need to be analyzed in order to further support this hypothesis. In addition, denitrification capacity has been previously detected in sediments depleted of nitrate and oxygen for a long period of time (Jorgensen and Tiedje, 1993; Tiedje et al., 1982), suggesting that some denitrifiers can survive for extended periods even in the absence of their normal electron acceptors. Based on these observations, the following hypotheses will be tested: Hypothesis] Denitrifier populations are distributed homogeneously within bioturbated sediments, regardless of the redox potential. Hypothesis 2 The pattern of expression of denitrification genes is equal for denitrifiers at any depth within the bioturbated sediments, when subjected to equal environmental conditions. 12 Objective 3 To study the denitrifier diversity and distribution in a broad range of geographic locations with varying geochemical characteristics. As stated above, geographic location has been previously determined to exert a strong effect in the differentiation of denitrifier communities in the Pacific Northwest (Braker et al., 2000; 2001). Water depth within one geographic area, which is correlated with gradually changing biogeochemical conditions, seemed to exert a less Significant effect, while sediment depth was the least influential of the scales studied. However, a broader range of geographic locations with varying geochemical characteristics needed to be analyzed in order to further test the following hypothesis: Hypothesis Geographic location, water depth within one geographic area, and sediment depth within the bioturbated zone exert decreasing influence, in that order, on the differentiation of denitrifier communities. 13 References Bethe, H., G. Jost, M. Schloter, B. B. Ward, and K.-P. Witzel. 2000. Molecular analysis of ammonia oxidation and denitrification in natural environments. FEMS Microbiol. Rev. 24:673-690. Braker, G., A. Fesefeldt, and K.-P. Witzel. 1998. Development of PCR primer systems for amplification of nitrite reductase genes (nirK and nirS) to detect denitrifying bacteria in environmental samples. Appl. Environ. Microbiol. 64:3769-3775. Braker, G. and J. M. Tiedje. 2003. Nitric oxide reductase (norB) genes from pure cultures and environmental samples. Appl. Environ. Microbiol. 69:3476-3483. Braker, G., H. A. Ayala-deI-Rio, A. H. Devol, A. Fesefeldt, and J. M. Tiedje. 2001. Community structure of denitrifiers, Bacteria, and Archaea along redox gradients in Pacific Northwest marine sediments by terminal restriction fragment length polymorphism analysis of amplified nitrite reductase (nirS) and 16S rRNA genes. Appl. Environ. Microbiol. 67: 1 893—1 901. Braker, G., J. Zhou, L. Wu, A. H. Devol, J. M. Tiedje. 2000. Nitrite reductase genes (nirK and nirS) as functional markers to investigate diversity of denitrifying bacteria in Pacific Northwest marine sediment communities. Appl. Environ. Microbiol. 66: 2096— 2104. Brandes, J. A., A. H. Devol, and C. Deutsch. 2007. New developments in the marine nitrogen cycle. Chem. Rev. 107:577-589. Castro-Gonzalez, M., G. Braker, L. Farias, and O. Ulloa. 2005. Communities of nirS- type denitrifiers in the water column of the oxygen minimum zone in the eastern South Pacific. Environ. Microbiol. 7:1298-1306. Cheneby, D., S. Hallet, M. Mondon, F. Martin-Laurent, J. C. German, and L. Philippot. 2003. Microb. Ecol. 46:113-121. Codispoti, L. A. 1995. Is the ocean losing nitrate? Nature 376:724. Coyne, M. S., A. Arunakumari, B. A. Averill, and J. M. Tiedje. 1989. Immunological identification and distribution of dissimilatory heme ed; and nonheme copper nitrite reductase in denitrifying bacteria. Appl. Environ. Microbiol. 55:2924-2931. Coyne, M. S., and J. M. Tiedje. 1990. Distribution and diversity of dissimilatory NOz' reductases in denitrifying bacteria. In N. P. a. S. Revsbech, J. (ed.), Denitrification in soil and sediment. Plenum Press, New York, NY. Dalsgaard, T., B. Thamdrup, and D. E. Canfield. 2005. Anaerobic ammonium oxidation (anammox) in the marine environment. Res. Microbiol. 156:457—464. 14 Devol, A. H. 1991. Direct measurement of nitrogen gas fluxes from continental shelf sediments. Nature 349:3 19-321 . Devol, A. H. 2003. Solution to a marine mistery. Nature 422:575-576. Devol, A., B. Chang, and J. Christensen. 2005. Denitrification and sulfate reduction in Arctic continental margin sediments. Goldschmidt Conference Abstracts. A581. Flanagan, D. A., L. G. Gregory, J. P. Carter, A. Karakas-Sen, D. J. Richardson, and S. Spiro. 1999. Detection of genes for periplasmic nitrate reductase in nitrate-respiring bacteria and in community DNA. FEMS Microbiol. Lett. 177 :263-270. Gamble, T. N., M. R. Betlach, and J. M. Tiedje. 1977. Numerically dominant denitrifying bacteria from world soils. Appl. Environ. Microbiol. 33:926-939. Ganeshram, R. S., T. F. Pedersen, S. E. Calvert, and J. W. Murray. 1995. Large changes in oceanic nutrient inventories from glacial to interglacial periods. Nature 376:755-758. Glockner, A. B., A. Jungst, and W. G. Zumft. 1993. Copper-containing nitrite reductase from Pseudomonas aureofaciens is functional in a mutationally cytochrome cdl-free background (N irS’) of Pseudomonas stutzeri. Arch. Microbiol. 160:18-26. Goregues, C. M., V. D. Michotey, and P. C. Bonin. Molecular, biochemical, and physiological approaches for understanding the ecology of denitrification. Microb. Ecol. 49:198-208. Hallin, S. and P.-E. Lindgren. 1999. PCR detection of genes encoding nitrite reductase in denitrifying bacteria. Appl. Environ. Microbiol. 65:1652-1657. Hannig, M., G. Braker, J. Dippner, and K. Jiirgens. 2006. Linking denitrifier community structure and prevalent biogeochemical parameters in the pelagial of the central Baltic Proper (Baltic Sea). FEMS Microbiol. Ecol. 57:260-271. Heid, C. A., J. Stevens, K. J. Livak, and P. M. Williams. 1996. Real time quantitative PCR. Genome Res. 6:986-994. Henry, S., D. Bru, B. Stres, S. Hallet, and L. Philippot. 2006. Quantitative detection of the nosZ gene, encoding nitrous oxide reductase, and comparison of the abundances of 16S rRNA, narG, nirK, and nosZ genes in soils. Appl. Environ. Microbiol. 72:5181- 5189. Henry, S., E. Baudoin, J. C. Lépez-Gutiérrez, F. Martin-Laurent, A. Brauman, and L. Philippot. 2004. Quantification of denitrifying bacteria in soils by nirK gene targeted real-time PCR. J. Microbiol. Methods 59:327-335. 15 Hickey, B. M. 1989. Patterns and processes of circulation over the Washington continental shelf and Slope. In M. R. Landry and B. M. Hickey (eds), Coastal Oceanography of Washington and Oregon. Elsevier, New York, NY. Jayakumar, D. A., C. A. Francis, S. W. A. Naqvi, and B. B. Ward. 2004. Diversity of nitrite reductase genes (nirS) in the denitrifying water column of the coastal Arabian Sea. Aquat. Microb. Ecol. 34:69-78. Jargensen, K. S., and J. M. Tiedje. 1993. Survival of denitrifiers in nitrate-free, anaerobic environments. Appl. Environ. Microbiol. 59:3297-3305. Knowles, R. 1982. Denitrification. Microbiol. Rev. 46:43-70. Liu, W.-T., T. L. Marsh, H. Cheng, and L. J. Forney. 1997. Characterization of microbial diversity by determining terminal restriction fragment length polymorphisms of genes encoding 16S rRNA. Appl. Environ. Microbiol. 63:4516-4522. Liu, X., S. M. Tiquia, G. Holguin, L. Wu, S. C. Nold, A. H. Devol, K. Luo, A. V. Palumbo, J. M. Tiedje, and J. Zhou. 2003. Molecular diversity of denitrifying genes in continental margin sediments within the oxygen-deficient zone off the Pacific coast of Mexico. Appl. Environ. Microbiol. 69:3549-3560. Lépez-Gutiérrez, J. C., S. Henry, S. Hallet, F. Martin-Laurent, G. Catroux, and L. Philippot. 2004. Quantification of a novel group of nitrate-reducing bacteria in the environment by real-time PCR. J. Microbiol. Methods 57:399-407. Luther 111, G. W., B. Sundby, B. L. Lewis, P. J. Brendel, and N. Silverberg. 1997. Interactions of manganese with the nitrogen cycle: Alternative pathways to dinitrogen. Geochim. Cosmochim. Acta 61:4043-4052. Michotey, V., V. Méjean, and P. Bonin. 2000. Comparison of methods for quantification of cytochrome ch-denitrifying bacteria in environmental marine samples. Appl. Environ. Microbiol. 66:1564-1571 . Middelburg, J. J., K. Soetaert, P. M. J. Herman, and C. H. R. Heip. 1996. Denitrification in marine sediments: A model study. Global Biogeochem.Cycles. 10:661- 671. Nogales, B., K. N. Timmis, D. B. Nedwell and A. M. Osborn. 2002. Detection and diversity of expressed denitrification genes in estuarine sediments after reverse transcription-PCR amplification from mRNA. Appl. Environ. Microbiol. 68:5017-5025. Oakley, B. B., C. A. Francis, K. J. Roberts, C. A. Fuchsman, S. Srinivasan, and J. T. Staley. 2007. Analysis of nitrite reductase (nirK and nirS) genes and cultivation reveal depauperate community of denitrifying bacteria in the Black Sea suboxic zone. Environ. Microbiol. 9:118-130. 16 Penton, C. R., A. H. Devol, and J. M. Tiedje. 2006. Molecular evidence for the broad distribution of anaerobic ammonium—oxidizing bacteria in freshwater and marine sediments. Appl. Environ. Microbiol. 72:6829-6832. Philippot, L. 2002. Denitrifying genes in bacterial and Archaeal genomes. Biochim. Biophys. Acta 1577 :355-376. Philippot, L. and S. Hallin. 2005. Finding the missing link between diversity and activity using denitrifying bacteria as a model functional community. Curr. Opin. Microbiol. 8:234-239. Priemé, A., G. Braker and J. M. Tiedje. 2002. Diversity of nitrite reductase (nirK and nirS) gene fragments in forested upland and wetland soils. Appl. Environ. Microbiol. 68:1893-1900. Qiu, X.-Y., R. A. Hurt, L.-Y. Wu, C.-H. Chen, J. M. Tiedje, and J.-Z. Zhou. 2004. Detection and quantification of copper-denitrifying bacteria by quantitative competitive PCR. J. Microbiol. Methods 59: 199-210. Santoro, A. E., A. B. Boehm and C. A. Francis. 2006. Denitrifier community composition along a nitrate and salinity gradient in a coastal aquifer. Appl. Environ. Microbiol. 72: 2102-2109. Scala, D. J. and L. J. Kerkhof. 1998. Nitrous oxide reductase (nosZ) gene-specific PCR primers for detection of denitrifiers and three nosZ genes from marine sediments. FEMS Microbiol. Lett. 162:61-68. Scala, D. J. and L. J. Kerkhof. 1999. Diversity of nitrous oxide reductase (nosZ) genes in continental Shelf sediments. Appl. Environ. Microbiol. 65:1681-1687. Scala, D. J. and L. J. Kerkhof. 2000. Horizontal heterogeneity of denitrifying bacterial communities in marine sediments by terminal restriction fragment length polymorphism analysis. Appl. Environ. Microbiol. 66:1980-1986. Smith, C. J., D. B. Nedwell, L. F. Dong, and A. M. Osborn. 2007. Diversity and abundance of nitrate reductase genes (narG and napA), nitrite reductase genes (nirS and nrfA), and their transcripts in estuarine sediments. Appl. Environ. Microbiol. 73:3612- 3622. Tamegai, H., L. Li, N. Masui, and C. Kato. 1997. A denitrifying bacterium from the deep sea at 11,000 m depth. Extremophiles 1:207-211. Tamegai, H., R. Aoki, S. Arakawa, and C. Kato. 2007. Molecular analysis of the nitrogen cycle in deep-sea microorganisms from the Nankai Trough: genes for nitrification and denitrification from deep-sea environmental DNA. Extremophiles 11:269-275. 17 Throbfick, I. N., K. Enwall, A. Jarvis, and S. Hallin. 2004. Reassessing PCR primers targeting nirS, nirK and nosZ genes for community surveys of denitrifying bacteria with DGGE. FEMS Microbiol. Ecol. 49:401-417. Tiedje, J. M. 1988. Ecology of denitrification and dissimilatory nitrate reduction to ammonium. In A. J. B. Zehnder (ed.), Biology of anaerobic microorganisms. John Wiley & Sons,Inc, New York, NY. Tiedje, J. M. 1994. Denitrifiers. In R. W. Weaver, J. S. Angle, and P. S. Bottomley (eds), Methods of Soil Analysis, Part 2. Microbiological and Biochemical Properties. Soil Sci. Soc. Amer., Madison, WI. Tiedje, J. M., A. J. Sexstone, D. D. Myrold, and J. A. Robinson. 1982. Denitrification: ecological niches, competition and survival. Antonie van Leeuwenhoek. 48:569-583. Tiquia, S. M., S. Gurczynski, A. Zholi, and A. Devol. 2006. Diversity of biogeochemical cycling genes from Puget Sound sediments using DNA microarrays. Environmental Technology. 27: 1377-1389. Torsvik, V., L. Ovreas, and T. F. Thingstad. 2002. Prokaryotic diversity-magnitude, dynamics, and controlling factors. Science 296:1064-1066. van de Graaf, A., A. Mulder, P. de Bruijn, M. S. M. Jetten, L. A. Robertson, and J. Gijs Kuenen. 1995. Anaerobic oxidation of ammonium is a biologically mediated process. Appl. Environ. Microbiol. 61: 1246-1251 . Walsh, J. J. 1991. Importance of continental margins in the marine biogeochemical cycling of carbon and nitrogen. Nature 350:53-55. Ward, B. B. 1996. Nitrification and denitrification: probing the nitrogen cycle in aquatic environments. Microb. Ecol. 32:247-261. Ward, B. B. and A. R. Cockcroft. 1993. Immunofluorescence detection of the denitrifying strain Pseudomonas stutzeri (ATCC 14405) in seawater and intertidal sediment environments. Microb. Ecol. 25:233-246. Woodgate, R. A., K. Aagaard, and T. J. Weingartner. 2005. A year in the physical oceanography of the Chukchi Sea: Moored measurements from autumn 1990-1991. Deep-Sea Res. 52:31 16-3149. Zumft, W. G. 1992. The denitrifying prokaryotes. In A. Balows, H. G. Triiper, M. Dworkin, W. Harder, and K.-H. Schleifer (eds), The prokaryotes, 2nd ed. Springer- Verlag, New York, NY. Zumft, W. G. 1997. Cell biology and molecular basis of denitrification. Microbiol. Mol. Biol. Rev. 61:533-616. 18 CHAPTER 2 PSEUDOMONAS S T U T ZERI NITRITE REDUCTASE GENE ABUNDANCE IN ENVIRONMENTAL SAMPLES MEASURED BY REAL-TIME PCR Abstract We used real-time PCR to quantify the denitrifying nitrite reductase gene (nirS), a functional gene of biogeochemical Significance. The assay was tested in vitro and applied to environmental samples. The primer-probe set selected was Specific for nirS sequences that corresponded approximately to the Pseudomonas stutzeri Species. The assay was linear from 1 to 106 gene copies (r2=0.999). Variability at low gene concentrations did not allow detection of twofold differences in gene copy number at less than 100 copies. DNA spiking and cell-addition experiments gave predicted results, suggesting that this assay provides an accurate measure of P. stutzeri nirS abundance in environmental samples. Although P. stutzeri abundance was high in lake sediment and groundwater samples, we detected low or no abundance of this species in marine sediment samples from Puget Sound (Wash) and from the Washington ocean margin. These results suggest that P. stutzeri may not be a dominant marine denitrifier. Introduction Denitrification is one of the important biogeochemical processes in that it is the main Sink for fixed nitrogen (Knowles, 1982). In agriculture, it accounts for 20 to 30% of fertilizer losses (Firestone, 1982) and in marine environments it is thought to account for 19 up to 80% of the loss of the nitrogen load to coastal areas (Seitzinger, 1990). Two of its products, NO and N20, are involved in global warming and the destruction of the ozone layer. Denitrification is also used in waste treatment facilities to remove excess combined nitrogen (Tiedje, 1988). More accurate understanding and modeling of denitrification should be possible if the catalyst could be quantified. We have evaluated the application of real-time PCR to the quantification of the nitrite reductase gene in Pseudomonas stutzeri. P. stutzeri is a denitrifier commonly isolated from both soil (Gamble et al., 1977) and marine environments (Ward and Cockroft, 1993), and may be of general importance in global denitrification. Nitrite reductase catalyzes a key step in the nitrogen cycle, in that the reduction of nitrite (NO2_) to nitric oxide (NO) converts N to a form no longer available to most of the biota. This enzyme is found as two different variants. One contains copper and is encoded by nirK, while the other contains the hemes c and d; and is coded by nirS (Zumft, 1997). nirK is found in a wider range of physiological groups, while nirS appears to be more abundant in nature (Coyne et al., 1989). P. stutzeri contains nirS (KOrner et aL,1987) Several methods have been used to attempt to identify or quantify denitrifiers, including the design of specific PCR primers (Braker et' al., 1998; Hallin et al., 1999) or probes (Smith and Tiedje, 1992) for genes involved in denitrification or rDNA genes (Kerkhof, 1994), and immunofluorescent assays using polyclonal antibodies against specific denitrifying bacteria (Ward and Cockroft, 1993) or denitrification enzymes (Coyne et al., 1989). Recently, competitive PCR and most-probable-number PCR were 20 applied to the quantification of nirS—containing denitrifying bacteria (Michotey et al., 2000) The real-time PCR technique is based on the use of the 5' nuclease assay, first described by Holland et al. (Holland et al., 1991) and further improved by the use of fluorescent TaqMan methodology and the ABI Prism 7700 Sequence Detection System (PE Applied Biosystems, Foster City, CA) (Gibson et al., 1977; Heid et al., 1996). The system requires the design of a forward and a reverse primer, in addition to a probe that hybridizes between them. The probe is fluorescently labeled at both ends (Lee et al., 1993). The fluorescent dye at the 5' end serves as reporter and its emission spectra is quenched by the dye at the 3' end of the probe. During the elongation step of each PCR cycle, the DNA polymerase cleaves the annealed probe, using its 5' nuclease activity. Once separated from the quencher, the reporter fluorescence is detected, resulting in an increase in fluorescence emission. The fluorescence increases logarithmically as the PCR reaction proceeds, until a reagent becomes limiting. A threshold fluorescence intensity is defined within the logarithmic phase. The higher the amount of initial template DNA, the earlier the fluorescence will cross the defined threshold. Copy number of the initial target DNA is thereby determined by comparison to a standard curve. The advantage of the real-time PCR method over other PCR-based quantification methods is that it focuses on the logarithmic phase of product accumulation rather than on the end product abundance. This technique is therefore more accurate, Since it is less affected by amplification efficiency or depletion of a reagent. In addition, real-time PCR measures template abundance over a large dynamic range of around Six orders of magnitude (Heid et al., 1996). Finally, this method allows the Simultaneous analysis of 21 96 samples in a short time and reduces the risk of contamination, as no post-PCR manipulation is required. The main disadvantage of real-time PCR is the need for a special thermocycler and reagents that are expensive compared to the equipment utilized by other PCR-based quantification methods. Real-time PCR has been successfully applied in the medical field, for example in the quantification of various DNA and RNA viruses in patients (Gut et al., 1999, Josefsson et al., 1999; Kimura et al., 1999; Martell et al., 1999), in the detection of gene amplification (Bieche et al., 1998), mutations or chromosomal rearrangements (Dblken et al., 1998; Luthra et al., 1998) and in the quantification of gene expression (Bieche et al., 1999; Wang and Brown, 1999) or detection of various splice variants (Kafert et al., 1999). Recently, real-time PCR was applied to environmental samples in studies that quantified conidia of a human pathogenic mold in airborne samples (Haugland et al., 1999) and for determining the abundance of bacterioplankton in marine samples (Suzuki etaL,2000) In this paper we test real-time PCR for its use for quantifying an important functional gene in environmental samples. Materials and Methods Samples Soil samples were obtained from an agricultural plot at the Kellogg Biological Station (MI) (Asuming-Brempong, 1999). Groundwater samples were taken from the Schoolcrafi (MI) and Shiprock Uranium Mill Tailings Remedial Action (UMTRA) (NM) 22 sites. The uranium recovery process at the latter site used high concentrations of ammonia, some of which entered the groundwater and led to the accumulation of nitrate. The Site is located on an elevated terrace, along the south side of the San Juan River. Samples were taken on the terrace (samples 813 and 826), on the floodplain (samples 602, 603, and 619), and from a seep flowing from the base of the escarpment into the floodplain (sample 425). They present increasing concentrations of nitrate in the following order: 602, 826, 619, 425, 603, and 813 (Ivanova et al., 2000). P. stutzeri KC, a strain that hydrolyzes carbon tetrachloride (Criddle et al., 1990), was injected into the Schoolcraft aquifer for a bioremediation field test 15 days before the samples were taken from wells 2 m upstream, and 1 In and 2.5 m downstream from the injection site (MS, M11, and M19, respectively) (Hyndman et al., 2000). The freshwater sediment sample was collected from the surface 2 cm sediment from Wintergreen Lake (MI), a small hypereutrophic lake. The marine sediment samples were obtained from the Washington margin of the Pacific Ocean and from Puget Sound (WA). Sediment cores were Sliced into sections described later. Cultures Marine P. stutzeri isolates (strains A3-5, D7-6, D9-1, E4-2, and F9-2) were obtained by L. Wu from the Washington marine sediments (Braker et al., 2000). nirS clones were obtained by G. Braker by amplification of DNA extracted from Washington margin and Puget Sound sediments using Specific primers (Braker et al., 2000). All culture collection strains and marine isolates used in this study were grown in nutrient broth (DIFCO, Detroit, MI). E. coli transforrnants were grown in Luria- Bertani (LB) 23 broth (Sambrook et al., 1989) amended with kanamycin (50 ug/ml). Shewanella oneidensis MR-l, P. stutzeri KC and P. aeruginosa were grown at 30°C, while all other cultures were grown at 37°C. DNA extraction and quantitation Genomic DNA was extracted from late exponential phase cultures. Cells were harvested by centrifugation and resuspended in lysis buffer (50 mM Tris pH 8, 50 mM EDTA, 100 mM NaCl). The cells were incubated for 15 min at 37°C with lysozyme (3.5 mg/ml), achromopeptidase (70 ug/ml), and RNAse A (30 ug/ml). After two freeze-thaw treatments, the cell suspension was incubated for 5 min at 37°C with SDS (1%), followed by incubation with proteinase K (400 ug/ml) (l h at 60°C). The lysate was extracted twice with phenol/chloroform/isoamyl alcohol. After isopropanol precipitation, the DNA was resuspended in TE buffer (pH 8). Plasmid DNA was extracted from the E. coli transforrnants with the Wizard® Plus SV Minipreps DNA Purification System (Promega, Madison, WI), according to the manufacturer's instructions. The genomic DNA extractions from the Kellogg Biological Station soil, the Shiprock aquifer, and the Wintergreen Lake sediment were performed using the Ultra CleanTM Soil DNA Kit (MO BIO, Solana Beach, CA), following the manufacturer's instructions. Genomic DNA was extracted from the Schoolcraft groundwater samples by the method of van Elsas and Smalla (van Elsas and Smalla, 1995). This method was also used to extract genomic DNA from the Washington marine sediment samples, with an additional proteinase K treatment (50 pl of a 20 mg/ml) after the incubation with SDS. The protocol of Gray and 24 Herwig (Gray et al., 1996) was used to extract genomic DNA from the Puget Sound samples. The quality of the extracted DNA was analyzed by electrophoresis through a 0.8% agarose gel. DNA concentrations were measured by absorbance at 260 nm. P. stutzeri nirS gene copy number was estimated based on the P. stutzeri Zobell genome Size (4.29 Mbp) (Ginard et al., 1997), and on the assumption that only one copy of nirS is present per genome (Jfingst et al., 1991); i.e. 4.4 fg P. stutzeri DNA = 1 genome mm = 1 nirS copy. Primers and probe Nine P. stutzeri nirS sequences from isolates (Braker et al., 2000) and 52 non-P. stutzeri nirS sequences from marine isolates, clones, and unrelated Species (Braker et al., 2000) were compared to select conserved regions within the P. stutzeri nirS gene. The primers and probe were designed within conserved regions using the program PrimerExpress (PE Applied Biosystems, Foster City, CA) (Table 2.1). The probe was dually labelled with the fluorescent dyes FAM (6-carboxyfluorescein) and TAMRA (6- carboxy-tetramethyl-rhodamine) at the 5' and 3' ends, respectively, as recommended by the manufacturers. Primers and probe were synthesized by Integrated DNA Technologies (Coralville, IA). Real-time PCR The increase in fluorescence emission, due to the degradation of the probe by the DNA polymerase in each elongation step, was monitored during PCR amplification using 25 the 7700 Sequence Detector (PE Applied Biosystems, Foster City, CA). The fluorescence signal was normalized by dividing the emission of the reporter dye (FAM) by the emission of the passive reference dye ROX (6-carboxy-X-rhodamine). The parameter CT (threshold cycle) is the fractional cycle number at which the fluorescence emission crosses an arbitrarily defined threshold within the logarithmic increase phase (0.1 in our reactions). The higher the amount of initial template DNA, the earlier the fluorescence will cross the threshold, and the smaller will be the CT. The CT values obtained for each sample are compared with a standard curve to determine the initial copy number of target gene. The reaction mixture for real-time PCR consisted of IX TaqMan® Universal PCR Master Mix [containing AmpliTaq GoldTM DNA Polymerase, AmpErase® UNG (uracil- N-glycosylase), which degrades PCR carry-over products from previous reactions, dNTPS with dUTP, a passive reference (ROX), and Optimized buffer components] (PE Applied Biosystems), 300 nM forward primer (FP), 900 nM reverse primer (RP), and 525 nM fluorogenic probe. MicroAmp® optical caps and tubes were used (PE Applied Biosystems, Foster City, CA). A total volume of 30 ul was used for the optimization steps and 50 pl was used for the final reactions. PCR reaction conditions were as follows: 2 min at 50°C, 10 min at 95°C, then 40 cycles of 15 S at 95°C and 1 min at 60°C. Negative controls with no template DNA or no probe were run in each reaction. Specificity The DNA extracted from several strains, marine isolates (Braker et al., 2000) and E. coli transforrnants (Braker et al., 2000) was used as positive and negative controls to 26 test the specificity of the primer-probe set. Template DNA (18 ng) was added to each reaction tube. Sensitivity and detection limit Marine isolate E4-2 DNA was chosen as a standard for measuring the sensitivity of the primer-probe set and for generating standard curves in subsequent determinations. This isolate was previously identified as P. stutzeri based on 16S rDNA sequence identity and physiological characters (Braker et al., 2000). Serial dilutions (IO-fold) of P. stutzeri E4-2 DNA were prepared in herring Sperm DNA [1 pg-ml'l herring sperm DNA (Boehringer Mannheim, Indianapolis, IN) in water] as a carrier. All determinations were performed in triplicate and error bars are reported as 95% confidence intervals. Various calibration curves were constructed to determine the lower detection limit of this assay and our ability to discriminate 2-fold differences in template concentration. A dilution series of marine isolate E4-2 DNA was prepared in a solution of 1 pg-ml'l herring sperm DNA. Different volumes (2 pl, 4 pl, 10 pl, and 20 pl) of template DNA from this dilution series were added to individual reaction tubes and the difference was made up with water. The maximum allowable error (MAE) in order to distinguish a twofold difference in copy number was calculated using the following equation: ACT = m - log(2) where ACT is the difference between CT obtained from samples with a twofold difference in target copy number, and m is the slope of the standard curve. The maximum allowable error (MAE) is: 27 MAE = ACT / 2 MAE was calculated for each standard curve generated and the mean MAE and the 95% confidence interval were determined. Quantitation of nirS in environmental samples Reactions were performed using 100 ng template DNA. The template copy number was determined from CT values using a standard curve. Samples that exhibited CT values equal to or higher than the negative controls were considered as below the detection limit. All results were normalized to the quantity of community DNA. Accuracy In order to test for the presence of PCR inhibitors, 106 strain E4-2 genome copies (4.4 ng DNA) were added to 100 ng DNA extracted from one environmental sample from each Site. Samples with and without the addition of this positive control DNA were compared and the percent recovery of the added genome copy number was calculated. P. stutzeri abundance was also evaluated in microbial communities constructed from P. stutzeri KC, P. aeruginosa and E. coli JM109 in different proportions and added to l g of sterile quartz sand (Sigma, St. Louis, MO). Direct cell counts were obtained before mixing using a Petroff-Hauser counting chamber. P. stutzeri KC cells were added to the mixtures in various ratios (1, 1/10, 1/102, 1/103, 1/104, 1/105, 1/107, 1/103, and 0). Equal cell numbers of P. aeruginosa and E. coli JM109 were added to achieve 2 x 108 cells in each 300 pl mixture. After cell addition, the mixture was vortexed for 3 s, and DNA was extracted using the Ultra CleanTM Soil DNA Kit (MO BIO), following the manufacturer's 28 instructions. Triplicate samples were analyzed using 100 ng of template DNA. To avoid any problem of different extraction efficiencies, the P. stutzeri KC nirS gene copy number was expressed relative to total DNA extracted. P. stutzeri, P. aeruginosa and E. coli genome Sizes were recently measured and reported as 4.29 Mbp (Ginard et al., 1977), 5.9 Mbp (Ratnaningsih et al., 1990) and 4.6 Mbp (Blattner et al., 1997) respectively; their percent G+C was considered to be 63%, 67% and 50%, respectively (Krieg et al., 1984). To measure the accuracy of real-time estimations of P. stutzeri abundance in environmental samples, different numbers of P. stutzeri cells were added to Wintergreen Lake sediment and KBS soil samples. The number of total cells in the soil or sediment samples was determined by direct counts after staining with 5-(4,6-dichlorotriazine-2-yl) aminofluorescein (DTAF) (Bloem, 1995). Serially diluted cells (108, 107, and 10'5 cells) were added to 0.5 g sediment or soil samples. After vortexing for 5 S, total DNA was extracted as above. In order to normalize these values to the total community DNA, the average genome size of soil or sediment bacteria was considered equal to the E. coli genome size, or 4.6 Mbp (Blattner et al., 1997). Results Specificity DNA extracted from a range of denitrifying isolates was used to test the specificity of the primer—probe combination (Table 2.2). Logarithmic increase in fluorescence intensity was readily detected by real-time PCR in 17 of 21 P. stutzeri strains. However, 29 we were not able to detect by real-time PCR the P. stutzeri strains corresponding to genomovars 4, 5, 7, and one strain in genomovar 1. Sequence analysis of nirS of the strains from genomovars 4, 5, and 7 indicated a higher Similarity (80%, 81%, and 82%, respectively) with the nitrite reductase gene of P. aeruginosa than with P. stutzeri. We identified several missmatches between these strains and the primer-probe combination used in this study (5 to 9 for the primers and 4 to 5 for the probe). No non-P. stutzeri strain or clone gave a real-time PCR signal. Sensitivity and detection limit We tested the sensitivity of the real-time detection system using a dilution series of P. stutzeri DNA (Fig. 2.1A). The threshold value for this and all subsequent analyses was chosen to be 0.1. This value falls within the range of logarithmic fluorescence increase, yet avoids the signal from the no template control (open circle). Nearly all of our no template controls Showed some increase in fluorescence intensity similar to that in Fig. 2.1A. Since this increase is not logarithmic and similar signals were detected in control reactions without DNA polymerase (data not shown), this fluorescence increase is likely due to probe degradation. The data obtained were used to draw a standard curve relating CT values to the added mass of P. stutzeri DNA and number of gene copies (Fig. 2.1B). A linear response was observed over more than six orders of magnitude, ranging from 14 to 4.05 x 106 nirS gene copies (r2 = 0.999, Fig. 2.13). We constructed a series of calibration curves to study the detection limit of the system and our ability to differentiate similar P. stutzeri DNA concentrations. Different volumes (ranging from 2 pl to 20 pl) of a dilution series of template DNA (ranging from 30 0.1 to 1000 nirS gene copies - pl") were added to the PCR reaction. Given the 96 well capacity, the analyses were done as a low and high concentration set (Fig. 2.2A and 2.2B, respectively). Although the curves remained linear down to 1 nirS copy (20 pl of 0.05 nirS gene copies - pl") the variability associated with CT values from low copy numbers preclude our ability to distinguish concentrations in samples with Similar amounts. Only when the copy number was 100 or greater were we able to reliably differentiate a 2- fold difference in P. stutzeri nirS concentration. The increase in variability with cycle number is apparent when the upper 95% confidence interval from all determinations is plotted against CT (Fig. 2.3). Based on the Slope of each individual standard curve, we calculated the maximal error in the CT value that would Still allow detection of a two-fold difference in gene copy number (MAE). The average of the different MAE is equal to 0.53 (Fig. 2.3). The corresponding 95% confidence interval is too small to be observed (Fig. 2.3). Accuracy P. stutzeri KC, E. coli JM109, and P. aeruginosa cells were added to sterile quartz sand in various known amounts. P. stutzeri KC contains the targeted nirS gene while P. aeruginosa has a nirS 67% similar in nucleotide sequence. The correlation between the calculated and measured values was extremely high (slope = 0.98, r2 = 0.992) (Fig. 2.48). The lowest proportions of nirS (1/107 and 1/108) overlapped with the background in the real—time PCR measurements. To further test the accuracy of the system, a Similar experiment was performed adding known amounts of P. stutzeri cells to different soil and sediment samples. 31 Measured values of nirS were correlated well (slope = 0.97, r2 = 0.971) with the values calculated from direct cell counts (Fig. 2.5). Analyses of environmental samples DNA extracted from environmental samples exhibited a wide range of P.stutzeri nirS gene abundance measured by real-time PCR (Table 2.3). The groundwater and freshwater sediment samples had the highest P. stutzeri nirS copies. Marine sediments consistently displayed low P. stutzeri nirS abundance. In order to test for the presence of any PCR inhibitors in the environmental samples, 106 genome copies of strain E4-2 were added to each sample. Five of the Six samples yielded the expected amount of nirS (Fig. 2.6). The KBS soil sample Showed slightly less nirS copies than expected. 32 Table 2.1. Primer and probe sequences compared to example positive and negative control nirS gene sequences Example nirS Forward primerb Probeb Reverse primerb sequence 5'ACAAGGAGCACAAC FAM- 5'CGCGTCGGCCCAGA3' TGGAAGGT3' S'GGCAACCTGTTCGTCAAG ACCCA3'-TAMRA 5'ACAAGGAGCACAAC 5'GGCAACCTGTTCGTCAAG 5'CGCGTCGGCCCAGA3' Posrtrve control, T G G A AG GT3' ACCCA3' P. stutzeri Zobell la 5'ATCCGQAGIACG_CCT 5'GGCTCGCTGTTCATCAAGA S'QCACAGCIGQATCTT Negative contro GGAAcgr CCCA3' P. aeruginosa ”Underlined bases represent missmatches with the primer-probe set. (Corresponding to P. stutzeri Zobell nirS positions 1260 to 1281, 1310 to 1332, and 1350 to 1363 (forward primer, probe, and reverse primer, respectively). 33 Table 2.2. Specificity of real-time PCR reactions for Pseudomonas stutzeri DNA Species, strain, or Reference Denitrifi- nirS‘ Real-time clonel (genonovar) or sourceb cation‘ PCR" Positive controls P- stutzeri (1) ATCC 17539 + + + P- stutzeri (l) ATCC 27951 + + + P. stutzeri (l) ATCC 17593 + + + P. stutzeri (l) ATCC 17594 + + + P. stutzeri (1) CCUGl 1256 + - P. stutzeri (2) ATCC 17591 + + + P. stutzeri (2) ATCC 17587 + + + P. stutzeri Zobell (2) ATCC 14405 + + + P. stutzeri (2) ATCC 17592 + + + P. stutzeri (2) ATCC 17595 + + + P. stutzeri (3) ATCC 50227 + + + P. stutzeri l9SMN4 (4) DSM 6084 + + - P. stutzeri DNSP21 (5) DSM 6082 + + - P. stutzeri (7) DSM 50238 + + _ P. stutzeri JM300 (8) DSM 10701 + + + P. stutzeri KC ATCC 55595 + + + P. stutzeri A3-5 L. Wu + + + P. stutzeri D7-6 L. Wu + + + P. stutzeri D9-1 L. Wu + + + P. stutzeri E4-2 L. Wu + + + P. stutzeri F 9-2 L. Wu + + + Negative controls E. coli K12 DSM 498 - - - Paracoccus denitrificans ATCC 17741 - Azospirillum brazilense DSM 1690 - Shewanella oneidensis MR-l ATCC 700550 - - - Pseudomonas aeruginosa ATCC 15692 + + - Pseudomonas balearica DSM 6083 + + - Marine isolate CIO-ls L. Wu + + - Marine isolate D4-14 L. Wu + + - nirS clone A4 G. Braker + + - nirS clone B6 G. Braker + + - nirS clone A12 G. Braker + + - nirS clone B76 G. Braker + + - ”Marine isolates and nirS clones were obtained from Puget Sound marine sediment (7). ”ATCC, American Type Culture Collection; DSM, Deutsche Sammlung von Mikroorganismen; CC UG, Culture Collection, University of Goteborg; L. Wu, Oak Ridge National Laboratory, Oak Ridge, Tenn; G. Braker, MSU, East Lansing, Mich. cPresence of denitrification and nirS from literature sources, except for the characterized genomovars, which were determined in this study by PCR. ”4r, logarithmic amplification readily detected; -, no amplification detected. 34 Figure 2.1. Generation of standard curve. (A) Increase of fluorescence intensity with cycle number for serially diluted P. stutzeri DNA. Symbols, from lefi to right: V , 17.8 ng; e , 5.9 ng; A , 0.59 ng; O , 59 pg; I , 5.9 pg; V , 0.59 pg; 0 , 59 fg; o , no template control. CT, Cycle at which the fluorescence intensity crosses an arbitrary threshold value. (B) Standard curve. Values represent means i 95% confidence interval (n=3). 35 Fluorescence intensity (ARn) Threshold cycle (CT) 101 - 10° - 104 40 Cycle number 35 - 30 - 25 - 20'- 15 - 10 8:0.999 y= -3.46x + 38.84 100 l f I I 101 1O2 1O3 104 105 106 107 Number of P. stutzeri nirS gene copies I I I I I 104 10'3 10'2 104 10° 10‘ ng P. stutzeri DNA 36 Figure 2.2. Limit of detection of the P. stutzeri nirS gene using real-time PCR. The denoted volumes of serially diluted P. stutzeri DNA were added to different reactions. (A), low concentrations, (B) high concentrations. Values represent means i 95% confidence interval (n=3). 37 Threshold cycle (CT) Threshold cycle (CT) 45 - A. 40 ‘ Volume added .~~ to reaction 35 - ‘§£‘ i . ~ 2 pl ‘ 4 pl 30 - ‘ 10 pl 20 pl 25 - o -1 1 10 100 P. stutzeri nirS gene copies-pl DNA'1 45 - B. 40 - Volume added 35 I to reaction 4 pl 10 pl 25 - 20 111 1 10 100 1000 P. stutzeri nirS gene copies-pl DNA'1 38 I- 10 O a... 0 Tu 8‘ o E .15 6~ ‘ o ‘c’ o 4~ '0 I0: 8 o 2" 3?: 0.53 o: o- a“) S. 3 I I I 10 20 30 40 Threshold cycle (CT) Figure 2.3. Relationship between threshold cycle (CT) and error. The horizontal line represents the maximum allowable error to discriminate between samples containing a twofold difference in P. stutzeri nirS cOpy number. Value Shown represents the mean from 14 independent standard curves. The 95% confidence interval is Shown as a dashed line. 39 109 —. <1 5 B. 108 107 i '8‘ S 51 2 cl .‘Q 106 .g' E‘ ' Ti . i 105 u.‘ 10! 10 1o- 10- 105 10' 1 Calculated P. stutzeri nirS gene coples x pg DNA" 104 103 102 10‘ 10° P. stutzeri nirS gene copies x pg DNA'1 1 1/10 1/102 1/103 1/104 1/105 1/107 1/108 Ratio of P. stutzeri cells in mixture Figure 2.4. Quantification of nirS in artificial mixtures of P. stutzeri KC, P. aeruginosa, and E. coli cells. (A) Comparison between P. stutzeri nirS gene copies - pg DNA" measured by real-time PCR (gray bars) and calculated values based on cell counts (black bars). (B) Correlation between calculated and measured values (slope = 0.98, r2 = 0.992). Error bars represent 95% confidence intervals (n=3) for both figures. 40 r2=0.971 1o8 4 y=0.97x + 0.60 '< z a U) :3. X (D .02 'U Q. 93 8 E E g 8, 107- 5g 2 't 0) Is 5 "f 106 . . Q 106 107 108 Calculated P. stutzeri nirS gene copies x pg DNA'1 Figure 2.5. Quantification of nirS in KBS soil (0) and Wintergreen Lake sediment (A) samples spiked with P. stutzeri cells. The line Shows the correlation between P. stutzeri nirS gene copies - pg DNA'l measured by real-time PCR and calculated values. Error bars represent 95% confidence intervals (n=3). 41 Table 2.3. P. stutzeri nirS gene abundance in various habitats. Habitat Sample No. of copies of P. stutzeri nirS P. stutzeri nirS DNA (p g no. or DNA/pg of community DNAf of community DNA) type Mean Upper 95% Mean Upper 95% C1 C1 Soil“ 2 .4 KBS (Mich) ‘1): 1&1} 4.331? 10x BDL BDL 100x BDL BDL Groundwater” Schoolcraft $4181 1 113 D116” 4131):}? Bioremediation Site M19 2' 0x 107 ' 8x9 (Mich.)° - X Shipmc" UMTRAC Site 425 1.921103 3.721103 8.421103 1621102 (N- Mex-1 602 3.121105 8.821104 1.4 3.92110'I 603 2.221105 2.121104 9.72110'1 9321102 619 2.01107 9.4x105 90 4.1 813 5.0x106 4.521105 22 2 826 5.021105 1.621105 2.2 6.92110" Freshwater sediment” Wintergreen Lake (MICII.) 2.2x106 3.1x105 9.7 1.3 Marine sediment‘ Pacific Ocean (Wash) 0-0.5 cm BDL BDL 0.5-1 cm BDL BDL 1-2 cm 1.021102 2.011102 4.42110'4 8.8x10“' 2-3 cm BDL BDL 3-5 cm BDL BDL 5—10 cm BDL BDL Puget Sound (Wash) 1-1.5 cm BDL BDL 1.52 cm 2.121104 2.421103 9.12110'2 1.02110'2 6-6.5 cm 7.0x102 2.021102 3.12110'3 8.82110“ “Soils were treated with 0, l, 10, and 100x the normal field rate (1.1 kg ha") of the herbicide 2,4-dichlorophenoxyacetate (1). ”Samples were collected 2 m upstream (M8), 1 m downstream (M11) and 2.5 m downstream (M19) from the well where P. stutzeri KC was injected (21). ° UMTRA, Uranium Mill Tailings Remedial Action ° The surface 2 cm containing the active denitrifiers were analyzed ' Depths within the sediment core are reported. I Results are from triplicate samples when 95% confidence interval (CI) is indicated and from single samples, otherwise. ‘ BDL, below detection limit. 42 107 106 .. P. stutzeri nirS copy number 103 — KBS SCH SHI WIN WAS PUG Samples Figure 2.6. P. stutzeri nirS copy number measured by real-time PCR without (black bars) and with (gray bars) the addition of 10'5 P. stutzeri nirS copy numbers to DNA extracted from different environmental samples. KBS, soil from Kellogg Biological Station (Ml); SCH, groundwater from Schoolcraft Bioremediation Site (MI); SHI, groundwater from Shiprock (NM); WIN, freshwater sediment from Wintergreen Lake (MI); WAS, marine sediment from Washington margin (WA); PUG, marine sediment from Puget Sound (WA). 43 Discussion Good detection methods share four features: specificity, sensitivity, precision, and accuracy. Real-time PCR appears to satisfy these requirements. The primer-probe set we designed for the P. stutzeri nirS gene amplified a group of strains that generally corresponded to the P. stutzeri species (Table 2.2). Except for one strain, all the representatives of the two more commonly isolated genomovars, l and 2, were readily amplified. The sequence Similarity of the nirS gene of the P. stutzeri strains in genomovars 4, 5, and 7 to the P. aeruginosa nitrite reductase gene explains the inability to detect them by real—time PCR. In addition to cultured strains, we selected a range of isolates and cloned nirS sequences from the Pacific Northwest marine environment for the primer-probe design. This makes our system especially appropriate for application in this environment. Although there may be cross-reactivity with related DNA in the natural microbial communities, none of our habitat-specific negative controls gave a positive reaction. Furthermore, the phylogenetically closest relative, P. balearica, was not detected. The need for the probe in real-time PCR requires the identification of three specific regions in the DNA sequence, rather than two, which provides for the high specificity. This however, can make the design of an appropriate primer-probe set more difficult. The 16S rRNA gene would be an alternative target that is more conserved but would sample a larger organismal group that may not correlate with function. The method was linear over more than Six orders of magnitude and sensitive down to 1 gene copy, Similar to the results obtained in other studies (Haugland et al., 1999; Kimura et al., 1999). However, the high variability associated with low target copies limits precision near the detection limit (Fig. 2.2). We are able to detect the presence of 44 only 1 copy, but not precisely, e.g. we cannot discriminate a 2-fold difference at 100 copies or less. The theoretical upper 95% confidence interval to allow the discrimination of a two-fold difference in copy number is 0.53 (Fig. 2.3). Confidence intervals below this value are achieved more frequently at lower CT values, equivalent to higher number of target copies. The error values associated with the measurements of P. stutzeri nirS abundance in various environmental samples is consistent with this characterization of precision, i.e. high concentrations have an error value at least one order of magnitude smaller than the measured value, while the error was in the same order of magnitude when nirS copy number was small (Table 2.3). The method proved to be extremely accurate. When P. stutzeri was mixed in various proportions with E. coli and P. aeruginosa cells, it could be detected when present at 100% to 0.001% in the mixture, presenting a high correlation between expected and measured results over this whole range. Furthermore, spiked samples gave the expected results. Environmental samples analyzed with real-time PCR displayed a wide range of P. stutzeri nirS abundances. P. stutzeri was highly abundant in freshwater sediment from Wintergreen Lake (MI), which agrees with a study by Gamble et al. (1977), who identified P. stutzeri as the dominant denitrifier in these sediments. P. stutzeri was either absent or present at extremely low population densities in marine sediments from the Washington margin and Puget Sound (WA). Ward and Cockcroft (Ward and Cockroft, 1993) measured the abundance of P. stutzeri in the water column of Monterrey Bay, CA, and found that it represented only 0.02 to 0.08% of the total bacterial community. Our findings indicate that P. stutzeri is also not abundant in marine sediments. This is in 45 agreement with the studies by Braker et a1. (2000), who observed nirS heterogeneity and the lack of a dominant group in nirS clones isolated from these Sites. Furthermore, these results are also consistent with studies done on other denitrification genes. Scala and Kerkhof (1999) Observed a high diversity among nitrous oxide reductase (nosZ) genes in marine sediments, with no overlap between environmental nosZ sequences and cultured denitrifiers. Groundwater samples from the Schoolcraft Bioremediation site 041) followed expected trends, corresponding with injection of substrates and P. stutzeri strain KC into a contamination plume. We detected no P. stutzeri upgradient from the inoculation site, and found the highest P. stutzeri abundance just downgradient from the inoculation Site. P. stutzeri was also prevalent at Shiprock perhaps resulting from the high nitrate contamination at that Site. Real-time PCR is a specific, sensitive, precise and accurate method for quantifying a gene or organism group in a broad range of environmental sample types. It remains to be seen whether functional gene sequences are conserved enough in a habitat so that a reasonable number of probe-primer sets can provide useful quantitative information for components of a group or process. This same method and primer-probe regions should be effective in quantifying mRNA and hence in determining which group of organisms or gene families are active in denitrification. Acknowledgements This work was supported by Department of Energy grants DE-FG02-98ER62535 and DE-FG02-97ER62469. We kindly thank Dr. Jorge Lalucat for providing the P. stutzeri strains with genomovar classifications, Dr. Allan Devol for collecting the Puget Sound and Pacific Ocean sediment samples, Dr. Phillip Long and the UMTRA personnel 46 for the Shiprock groundwater samples, Liyou Wu for the marine isolates, Dr. Gesche Braker for the nirS clones and for providing us DNA extracted from the Pacific Ocean sediments, and Katrina Linning for DNA extracted from the Schoolcraft samples. Oak Ridge National Laboratory is managed by University of Tennessee-Battelle LLC for the Department of Energy under contract DE-AC05-000R22725. 47 References Asuming-Brempong, S. 1999. The effect of 2,4-dichlorophenoxyacetate selection on microbial communities in microcosm and field studies and the impact on ecosystem function. Ph.D. thesis. Michigan State University, East Lansing. Bieche, I., I. Laurendeau, S. Tozlu, M. Olivi, D. Vidaud, R. Lidereau, and M. Vidaud. 1999. Quantitation of MYC gene expression in sporadic breast tumors with a real-time reverse transcription-PCR assay. Cancer Res. 59:2759-2765. Bieche, 1., M. Olivi, M.-H. Champeme, D. Vidaud, R. Lidereau, and M. Vidaud. 1998. Novel approach to quantitative polymerase chain reaction using real-time detection: application to the detection of gene amplification in breast cancer. Int. J. Cancer 78:661-666. Blattner, F. R., G.3. Plunkett, C. A. Bloch, N. T. Perna, V. Burland, M. Riley, J. Collado-Vides, J. D. Glasner, C. K. Rode, G. F. Mayhew, J. Gregor, N. W. Davis, H. A. Kirkpatrick, M. A. Goeden, D. J. Rose, B. Man, and Y. Shao. 1997. The complete genome sequence of Escherichia coli K-12. Science 277: 1453-1474. Bloem, J. 1995. Fluorescent staining of microbes for total direct counts, p.1-12. In A.D.L. Akkermans, J .D. van Elsas, and F.J. de Brujin (ed.), Molecular microbial ecology manual, Kluwer Academic Publishers, Dordrecht, The Netherlands. Braker, G., A. Fesefeldt, and K.-P. Witzel. 1998. Development of PCR primer systems for amplification of nitrite reductase genes (nirK and nirS) to detect denitrifying bacteria in environmental samples. Appl. Environ. Microbiol. 64:3769-3775. Braker, G., J. Zhou, L. Wu, A. H. Devol, and J. M. Tiedje. 2000. Nitrite reductase genes (nirK and nirS) as functional markers to investigate diversity of denitrifying bacteria in Pacific Northwest marine sediment communities. Appl. Environ. Microbiol. 66:2096-2104. Coyne, M. S., A. Arunakumari, B. A. Averill, and J. M. Tiedje. 1989. Immunological identification and distribution of dissimilatory heme cdl and nonheme copper nitrite reductase in denitrifying bacteria. Appl. Environ. Microbiol. 55:2924-2931. Criddle, C. S., J. T. DeWitt, D. Grbic-Galic, and P. McCarty. 1990. Transformation of carbon tetrachloride by Pseudomonas Sp. strain KC under denitrification conditions. Appl. Environ. Microbiol. 56:3240-3246. Diilken, L., F. Schiller, and G. Diilken. 1998. Quantitative detection oft(14;18)-positive cells by real-time quantitative PCR using fluorogenic probes. BioTechniques 25:1058- 1064. 48 Firestone, M. K. 1982. Biological denitrification, p.289-326. In F.J. Stevenson (ed.), Nitrogen in agricultural soils. Agronomy monograph 22., American Society for Agronomy, Madison, Wis. Gamble, T. N., M. R. Betlach, and J. M. Tiedje. 1977. Numerically dominant denitrifying bacteria from world soils. Appl. Environ. Microbiol. 33:926-939. Gibson, U. E. M., C. A. Heid, and P. M. Williams. 1996. A novel method for real time quantitative RT-PCR. Genome Research 6:995-1001. Ginard, M., J. Lalucat, B. Tiimmler, and U. Riimling. 1997. Genome organization of Pseudomonas stutzeri and resulting taxonomic and evolutionary considerations. Int. J. Syst. Bacteriol. 47: 1 32-143. Gray, J. P. and R. P. Herwig. 1996. Phylogenetic analysis of the bacterial communities in marine sediments. Appl. Environ. Microbiol. 62:4049-4059. Gut, M., C. M. Leutenegger, J. B. Huder, N. C. Pedersen, and H. Lutz. 1999. One- tube fluorogenic reverse transcription-polymerase chain reaction for the quantitation of feline coronaviruses. J. Virol. Methods 77:37-46. Hallin, S., Lindgren, and P.-E. 1999. PCR detection of genes encoding nitrite reductase in denitrifying bacteria. Appl. Environ. Microbiol. 65:1652-1657. Haugland, R. A., S. J. Vesper, and L. J. Wymer. 1999. Quantitative measurement of Stachybotrys chartarum conidia using real time detection of PCR products with the TaqManTM fluorogenic probe system. Mol. Cell. Probes 13:329-340. Heid, C. A., J. Stevens, K. J. Livak, and P. M. Williams. 1996. Real time quantitative. PCR. Genome Res. 6:986-994. Holland, P. M., R. D. Abramson, R. Watson, and D. H. Gelfand. 1991. Detection of Specific polymerase chain reaction product by utilizing the 5'-3'exonuclease activity of T hermus aquaticus DNA polymerase. Proc. Natl. Acad. Sci. 88:7276-7280. Hyndman, D. W., M. J. Dybas, D. Wiggert, X. Zhao, R. Wallace, T. Voice, A. Chan, M. S. Phanikumar, and C. S. Criddle. 2000. Hydraulic characterization and design of a full-scale biocurtain. Ground Water 38:462-474. Ivanova, I. A., J. R. Stephen, Y.-J. Chang, J. Briiggemann, P. E. Long, J. P. McKinley, G. A. Kowalchuk, D. C. White, and S. J. Macnaughton. 2000. A survey of 168 rRNA and amoA genes related to autotrophic ammonia-oxidizing bacteria of the [3- subdivision of the class proteobacteria in contaminated groundwater. Can. J. Microbiol. 46:1012-1020. Josefsson, A., K. .Livak, and U. Gyllenstein. 1999. Detection and quantitation of human papillomavirus by using the fluorescent 5' exonuclease assay. J. Clin. Microbiol. 37:490-496. 49 Jiingst, A., S. Wakabayashi, H. Matsubara, and W. G. Zumft. 1991. The nirSTBM region coding for cytochrome cdl-dependent nitrite respiration of Pseudomonas stutzeri consists of a cluster of monO-, di-, and tetraheme proteins. FEBS Lett. 279 :205-209. Kafert, S., J. Krauter, A. Ganser, and M. Eder. 1999. Differential quantitation of alternatively spliced messenger RNAS using isoform-Specific real-time RT-PCR. Anal. Biochem. 269:210-213. Kerkhof, L. 1994. A species-specific probe and a PCR assay for the marine bacterium, Pseudomonas stutzeri strain Zobell. Microb. Ecol. 27 :201-212. Kimura, H., M. Morita, Y. Yabuta, K. Kuzushima, K. Kato, S. Kojima, T. Matsuyama, and T. Morishima. 1999. Quantitative analysis of Epstein-Barr virus load by using a real-time PCR assay. J. Clin. Microbiol. 37 :132-136. Knowles, R. 1982. Denitrification. Microbiol. Rev. 46:43-70. Ktirner, H., K. Frunzke, K. Dohler, and W. G. Zumft. 1987. Immunochemical patterns of distribution of nitrous oxide reductase and nitrite reductase (cytochrome cdl) among denitrifying pseudomonads. Arch. Microbiol. 148:20-24. Krieg, N. R. and J. G. Holt (Eds.). 1984. Bergey's Manual of Systematic Bacteriology. Williams & Wilkins, Baltimore, MD. Lee, L. G., C. R. Connell, and W. Bloch. 1993. Allelic discrimination by nick- translation PCR with fluorogenic probes. Nucleic Acids Res. 21:3761-3766. Luthra, R., J. A. McBride, F. Cabanillas, and A. Sarris. 1998. Novel 5' exonuclease- based real-time PCR assay for the detection of t(14;l8)(q32;q21) in patients with follicular lymphoma. Am. J. Pathol. 153263-68. Martel], M., J. Gémez, J. I. Esteban, S. Sauleda, J. Quer, B. Cabot, R. Esteban, and J. Guardia. 1999. High-throughput real-time reverse transcription-PCR quantitation of hepatitis C virus RNA. J. Clin. Microbiol. 37:327-332. Michotey, V., V. Méjean, and P. Bonin. 2000. Comparison of methods for quantification of cytochrome cdl-denitrifying bacteria In environmental marine samples. Appl. Environ. Microbiol. 66: 1564- 1571. Ratnaningsih, E., S. Dharmsthiti, V. Krishnapillai, A. Morgan, M. Sinclair, and B. W. Holloway. 1990. A combined physical and genetic map of Pseudomonas aeruginosa PAO. J. Gen. Microbiol. 136:2351-2357. Sambrook, J., E. F. Fritsch, and T. Maniatis. 1989. Molecular cloning: a laboratory manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. Scala, D. J. and L. J. Kerkhof. 1999. Diversity of nitrous oxide reductase (nosZ) genes in continental Shelf sediments. Appl. Environ. Microbiol. 65 :1681-1687. 50 Seitzinger, S. P. 1990. Denitrification in aquatic sediments, p.301-322. In N.P. Revsbech and J. Sorensen (ed.), Denitrification in soil and sediment, FEMS Symposium, Plenum Press, New York. Smith, G. B. and J. M. Tiedje. 1992. Isolation and characterization of a nitrite reductase gene and its use as a probe for denitrifying bacteria. Appl. Environ. Microbiol. 58:376- 384. Suzuki, M. T., L. T. Taylor, and E. F. Delong. 2000. Quantitative analysis of small- subunit rRNA genes in mixed microbial populations Via 5'-nuclease assays. Appl. Environ. Microbiol. 66:4605-4614. Tiedje, J. M. 1988. Ecology of denitrification and dissimilatory nitrate reduction to ammonium, p.179-244. In A.J.B. Zehnder (ed.), Biology of anaerobic microorganisms, John Wiley & Sons,Inc, New York, NY. van Elsas, J. D. and K. Smalla. 1995. Extraction of microbial community DNA from soils, p.1-11. In A.D.L. Akkerrnans, J.D. van Elsas, and F .J. de Brujin (ed.), Molecular microbial ecology manual, Kluwer Academic Publishers, Dordrecht, The Netherlands. Wang, T. and M. J. Brown. 1999. mRNA quantification by real time TaqMan polymerase chain reaction: validation and comparison with RNase protection. Anal. Biochem. 269: 1 98-201. Ward, B. B. and A. R. Cockcroft. 1993. Immunofluorescence detection of the denitrifying strain Pseudomonas stutzeri (ATCC 14405) in seawater and intertidal sediment environments. Microb. Ecol. 25:233-246. Zumft, W. G. 1997. Cell biology and molecular basis of denitrification. Microbiol. Mol. Biol. Rev. 61:533-616. 51 CHAPTER 3 DENITRIFIER COMMUNITY STRUCTURE AND ACTIVITY WITH RESPECT TO REDOX GRADIENTS AND BIOTURBATION Abstract Denitrifiers have previously been detected at sediment depths below the penetration depths of oxygen and nitrate, their normal electron acceptors. We were therefore interested in studying the vertical distribution of denitrifiers, with respect to redox gradients and presence of macrobenthic fauna, which could redistribute bacteria within the sediments and allow the penetration of overlying water deep into the sediments. Three highly bioturbated Sites in Puget Sound, Washington, were compared to a Washington Margin area with lower bioturbation. Despite the shallow oxygen and nitrate penetration at the studied sites (<2cm), the heme cd; nitrite reductase gene (nirS) containing community presented a conserved structure, as determined by T-RFLP, within the bioturbated sediments (to at least 37 cm and ~15 cm below the sediment surface depth in Puget Sound and Washington Margin samples, respectively), with a Sharp change in community structure in deep unbioturbated sediments. PCA ordination and cluster analysis grouped the sediments based on sampling location and mixing, with unmixed sediments from all areas forming a separate group. Quantification by real-time PCR of clusters of nirS clones retrieved from these same sediments followed this same pattern, with significant reductions of abundance in deep unmixed sediments. On the other hand, P. stutzeri was generally present at highly constant abundance throughout the sediment depths within as well as below the bioturbated zone. Denitrification capacity, determined 52 as nitrous oxide production in microcosms after nitrate amendment, was detected throughout the mixed layers at all Sites, confirming the presence of denitrifiers capable of denitrification in sediments below the depth where nitrate is present in situ. Introduction Many near shore and continental shelf and slope areas are characterized by high carbon flux to the sediments, due to their high surface productivity and Shallow water depth, leading to a high sedimentary respiration rate (Brandes and Devol, 1995). Puget Sound and Washington continental margin sediments are an example of this phenomenon. The high respiration rate is responsible for a rather shallow penetration depth of nitrate, generally in the order of 1 cm (Braker et al., 2000; Brandes and Devol, 1995). Therefore, active denitrification is expected to occur in situ only in the top 1 cm, as the presence of an N-oxide is generally required for the expression of the denitrification genes (Zumft, 1997). Braker et al., however, were able to clone nirS gene sequences from 6.5 cm deep sediments from Puget Sound (Braker et al., 2000) and detected a highly Similar denitrifier community structure, as determined by T-RFLP, in the top 6.5 and 10 cm of Puget Sound and Washington continental shelf sediments, respectively (Braker et al., 2001). They suggested that sediment mixing by marine invertebrates could have been responsible for the conserved community structure at different depths despite the unfavorable electron acceptor gradients. Wadden Sea sediments also presented a similar microbial community structure down to 4.75 cm when major phylogenetic groups were studied by fluorescent in situ hybridization (FISH), regardless of the redox gradients (Llobet-Brosa et al., 1998). Furthermore, in Tokyo Bay 53 sediments, only minor bacterial groups changed in the top 10 cm, while the major population structure remained stable, as revealed by T-RFLP and quinone profiling (Urakawa et al., 2000). None of these studies, however, included deep sediments lying below the bioturbated zone. Besides detection of denitrification genes in nitrate depleted sediments, denitrification capacity has also been detected in environments devoid of nitrate for a long period of time, c. g. nitrate-free sludge, eutrophic lake sediments and river sediments (Jorgensen and Tiedje, 1993; Tiedje et al., 1982). Oxygen was also absent from these environments, hence, no aerobic respiration could take place. A low level of fermentation was suggested as an explanation for the long-term survival of these denitrifiers in the absence of nitrate and oxygen (Jorgensen and Tiedje, 1993). Puget Sound sediments support a dense benthic population, responsible for a high sediment mixing rate of 43 cm2 y". Surface mixed layers of up to 22 cm have been observed in these sediments by the analysis of excess me activity profiles (Carpenter et al., 1985). On the other hand, in Washington slope sediments, though variable, surface sediment mixing coefficients are markedly lower than in Puget Sound (0.25-9.8 cm2 y'l), with average depths of surface mixed layers of 6.2 :E 3.5 cm (Carpenter et al., 1982). Furthermore, the sediment mixing coefficients in this area suffer a particularly sharp decrease below a water depth of 800 to 1400 m (<1 cm2 y'l). The low dissolved 02 concentration in waters over the Slope from about 600 to 1400 m probably accounts for this decrease due to the inhibition of macrobenthic activity (Carpenter and Peterson, 1989). This clear difference in sediment mixing makes Puget Sound and the low 02 zone of the Washington Slope ideal sites for comparison of sediments under different 54 bioturbation intensities. These Sites were therefore selected to test the following hypotheses based on the above rationale: ° Denitrifier populations are distributed homogeneously within bioturbated sediments, regardless of the redox potential. 0 The pattern of expression of denitrification genes is equal for denitrifiers at any depth within the bioturbated sediments, when subjected to equal environmental conditions. In order to test the hypotheses, sediments from mixed and unmixed layers were analyzed in regard to denitrifier community structure and diversity by means of nirS clone library construction and T-RFLP, abundance of certain nirS phylotypes determined by real-time PCR, and denitrification capacity after nitrate addition to microcosms. Materials and Methods Study area and sediment sampling Marine sediment samples were collected from three stations in Puget Sound, Washington (Shallow Budd Inlet, Turning Basin, and Carr Inlet) and from the continental slope at the Washington margin of the Pacific Ocean (Figure 3.1, Table 3.1). The Puget Sound and Washington margin samples were retrieved in July 2000 aboard the R/V Clifford A. Barnes and in July 2001 aboard the R/V T.G. Thompson, respectively. One core of at least 38 cm length was obtained from each site in Puget Sound using a Soutar box corer and subsequent collection of subcores with 7.5- and lO-cm cast acrylic tubes (Kristensen et al., 1999). A multicorer was utilized to retrieve the sample at the Washington margin. Deep sediment samples were collected at Turning Basin (69 cm 55 sediment depth), Carr Inlet (135 cm sediment depth) and Washington margin (62 cm) using a gravity corer. Soutar box corer subcores and multicorer cores were sectioned into 0.5 cm Slices and the sediments from the desired depth intervals (Table 3.2) were kept at -20°C for DNA extraction. Intact cores from each station were kept at 4°C for microcosm preparation. Chemical analyses of sediment samples The overlying waters and porewaters were sampled with a whole core squeezing apparatus and analyzed for oxygen and nitrate plus nitrite concentrations as described by Brandes and Devol (Brandes and Devol, 1995). Sediment cores were sectioned and centrifuged at 1000 X g for 20 min. to separate the pore water for ammonium (Strickland and Parsons, 1972), sulfate (Tabatabai, 1974), iron (Stookey, 1970) and manganese (Brewer and Spencer, 1974) determination. The sulfate reduction rate in the sediments was measured by the method of Fossing and Jorgensen (Fossing and Jorgensen, 1989). The percentage of organic carbon was obtained from other studies in the sampled area (Devol, pers. comm.) and was determined by the method of Hedges and Stern (Hedges and Stern, 1984). DNA extraction and purification Genomic DNA was extracted from 0.5 g of sediment from each sample using the Ultra Clean Soil DNA kit (MO BIO, Solana Beach, CA), following the manufacturer’s instructions. For Real Time PCR, the DNA was further purified by gel extraction from a 0.8% Sea Plaque agarose gel with B-agarase (Cambrex Bio Science, Rockland, ME), as 56 indicated by the manufacturer, followed by concentration with Microcon YM-IOO columns (Millipore, Bedford, MA). Genomic DNA was extracted as described in Chapter 2 (also, Griintzig et al., 2001) from marine P. stutzeri isolate E4-2 (Braker et al., 2000) grown in nutrient broth (Difco, Detroit, M1) at 37°C. Plasmid DNA was extracted from Escherichia coli transforrnants grown in Luria-Bertani broth (Sambrook et al., 1989) at 37°C using the QIAprep Spin Miniprep Kit (QIAGEN, Valencia, CA). The quality of the extracted DNA was analyzed by electrophoresis on a 0.8% or 1% agarose gel for genomic or plasmid DNA, respectively, and the concentration was determined by absorbance at 260 nm. T-RFLP analysis of nirS gene Terminal restriction fragment length polymorphism (T-RFLP) analysis of the sediment samples was performed as described by Braker et a1. (2000; 2001) with modifications. The nirSlF primer labeled at the 5’ end with 6-carboxyfluorescein (Integrated DNA Technologies, Coralville, IA) and nirS6R primer (Braker et al., 1998) were used to amplify ~890 bp of the nirS gene in five replicate 50 pl PCR reactions for each environmental DNA sample. The PCR reactions contained 150 ng environmental DNA, 1 pM of each primer, 200 pM of each deoxyribonucleoside triphosphate, 2.5 mM MgCl2, 0.4 pg pl'l bovine serum albumin (Roche Molecular Biochemicals, Indianapolis, IN), 1.5 U T aq polymerase (Promega, Madison, WI), and 1X buffer provided with the enzyme. After an initial denaturation step of 3 min at 95°C, a touchdown PCR was performed as described by Braker et al. (2000). The replicate PCR reactions were pooled, concentrated to approximately 60 pl with a Speedvac (Savant, Holbrook, NY) and the 57 amplified 890 bp nirS fragments were purified from a 2% preparatory agarose gel with the QIAquick Gel Extraction kit (QIAGEN, Valencia, CA) eluting in 30 pl sterile-filtered cell culture tested water (Sigma, St. Louis, MO). Aliquots (10 pl) were digested with 5U of restriction enzymes MspI and HhaI (New England Biolabs, Beverly, MA) in Single enzyme reactions for 14 h at 37°C in the manufacturer’s recommended reaction buffers. The restriction endonucleases were inactivated by incubating the reaction mixture at 65°C for 25 min. MspI and HhaI were used because Braker et al. (2001) have shown that they yielded the highest number and most even distribution of T-RFS at these Sites. Two different methods were applied for the separation and identification of the generated terminal restriction fragments (T-RFS). The digested DNA samples from Puget Sound sediments were concentrated to 2 pl and analyzed on a 373 ABI Stretch automated DNA sequencer (Applied Biosystems Instruments, Foster City, CA) by gel electrophoresis, as described by Braker et al. (2001). The T-RFS generated from the Washington margin sediment, on the other hand, were separated by capillary electrophoresis with an Applied Biosystems PRISM 3100 Genetic Analyzer (Applied Biosystems, Foster City, CA). This system loads the sample by electrokinetic injection, which is highly affected by the presence of small ions in the sample, as they will be injected preferentially, due to their high charge-to-mass ratio, interfering with the DNA uptake. Therefore, an additional desalinization of the inactivated restriction digest was performed, by adding sterile-filtered cell culture tested water (Sigma, St. Louis, MO) up to a final volume of 500 pl, before concentration and desalinization on a Microcon YM- 10 column ((Millipore, Bedford, MA). The desalinized sample was firrther concentrated to approximately 12 pl with a Speedvac (Savant, Holbrook, NY) and 4 pl was mixed with 58 forrnamide and internal standard and loaded onto the capillary electrophoresis system with an injection time and voltage of 30 s and 3 kV, respectively. These conditions revealed Similar results to those previously Obtained from Carr Inlet by gel electrophoresis. Capillary electrophoresis reported longer T-RF Sizes as the fragment size increased; this shift was normalized during data analysis to equate with the data from gel electrophoresis. The lengths of the T-RFS were determined by comparison with the internal standard using GeneScan 3.7 (Applied Biosystems, Foster City, CA) software. Analysis of nirS T-RFLP profiles Electropherograms from each of the analyzed samples were compared with Genotyper 3.7 software (Applied Biosystems, Foster City, CA) and peaks, over a threshold of 50 fluorescence units in at least one sample, representing the same T-RF were manually aligned. The alignment was based primarily on fragment length, but the pattern of peaks was also taken into consideration (Blackwood et al., 2003). An average fi'agment length was determined for each T-RF (Dunbar et al., 2001). Only T-RFS from 55 to 600 bp length were included in the analysis, to avoid the detection of the primers and Size determination uncertainties. The presence and height of peaks was used to characterize the communities. In order to normalize the results, the relative abundance of peaks was determined for each sample by dividing the height of each peak by the total fluorescence intensity of that sample. The results were converted to percentages and represented as histograms. The profiles resulting from the single-enzyme reactions were analyzed independently, as suggested by Dunbar et al. (Dunbar et al., 2000). 59 The T-RFLP profiles from sections within and between cores were compared by Principal Component Analysis (PCA) after calculating the Hellinger distance between profiles a31ackwood et al., 2003). The Hellinger distance is determined by calculating the Euclidean distance between profiles after square root transformation of relative peak heights. This transformation allows the use of PCA on community composition data containing many zeros as a result of representing a long gradient, with loss and appearance of several Species (Legendre and Gallagher, 2001). A dendrogram was constructed to visualize Similarities between profiles by the application of hierarchical agglomerative cluster analysis to the Hellinger distances between profiles. Ward’s method was chosen for group linking (Ward, 1963). PCA and cluster analysis were performed using the Community Analysis Package software version 3.01 (Pisces Conservation Ltd., Hampshire, United Kingdom). An in silico digestion of the nirS clone sequences retrieved in this study and of 68 additional nirS sequences from strains and environmental clones from GenBank was performed for HhaI and MspI with a custom made Perl script. The theoretical T-RFLP profile produced by the environmental clones retrieved in this study was compared to the actual T-RFLP profiles. Cloning of nirS sequences A clone library of nirS sequences was constructed from Shallow Budd Inlet sediments corresponding to a depth interval from 2 to 2.5 cm. This Specific depth was chosen due to its high richness in nirS sequences based on T-RFLP. Environmental DNA extracted from these sediments was PCR amplified as described above for T-RFLP analysis with some modifications. Unlabelled primers nirSlF and nirS6R were used in 10 60 replicate PCR reactions. In addition, only 20 cycles, instead of 25 cycles, were performed in the stable phase of the touchdown PCR to reduce PCR-induced artifacts and biases (Qiu et al., 2001; Acinas et al., 2005). The replicate PCR reactions were pooled, concentrated to approximately 140 pl with a Speedvac (Savant, Holbrook, NY) and the amplified 890 bp nirS fragments were purified from a 1.2% preparatory agarose gel with the QIAquick Gel Extraction kit (QIAGEN, Valencia, CA) following the manufacturer’s instructions. The purified fragments were cloned using the TA cloning kit (Invitrogen, Carlsbad, CA) and 188 white insert-bearing clones were randomly selected for sequencing. These clones were grown on LB freezing buffer (Sambrook and Russell, 2001) with 50 pg ml'l kanamycin and inserts were sequenced with vector primer Ml3F using an ABI 1730 Genetic Analyzer or an ABI Prism 3700 DNA Analyzer (Applied Biosystems, Foster City, CA). Phylogenetic analysis of cloned nirS sequences The nirS sequences obtained in this study were compared to the GenBank database by using BLAST to remove sequences with no homology to nirS. In addition, GenBank sequences closely similar to the cloned sequences were identified. The sequences were aligned with a Specialized Hidden Markov Model aligner utilized by the Functional Gene Pipeline/Repository of the RDP. Closely Similar sequences chosen from GenBank were added to the alignment by profile alignment in Clustal X (Thompson et al., 1997). The aligned sequences were translated and the alignment manually corrected in BioEdit (Hall, 1999). A phylogenetic tree was inferred by the neighbour-joining method (Saitou and Nei, 1987) with Mega 2.1 (Kumar et al., 2001) using Poisson correction (Nei and Kumar, 2000) and 100 bootstraps. 61 Quantification of nirS from Specific clusters and strains by real-time PCR Five clusters of highly related nirS sequences as well as the nirS gene from P. stutzeri were quantified in the studied sediments by real-time PCR. Syerreen methodology was used to amplify four of the nirS clusters, while TaqMan methodology was used to amplify cluster 2 and P. stutzeri nirS. ARB (Ludwig et al., 2004) and PrimerExpress (Applied Biosystems, Foster City, CA) software were used to design primers and probes specific for the studied clusters (Table 3.4). A previously described primer-probe set was applied in the quantification of P. stutzeri nirS (Griintzig et al., 2001; Chapter 2). The probes used for TaqMan methodology were dually labeled with the fluorescent dyes 6-carboxyfluorescein (FAM) and 6-carboxytetramethyl-rhodamine (TAMRA) at the 5’ and 3’ ends, respectively. Primers and probes were synthesized by Integrated DNA Technologies (Coralville, IA). Syerreen real-time PCR assays were carried out on an ABI Prism 7900HT Sequence Detection System (Applied Biosystems, Foster City, CA) in 10 pl reactions using the Syerreen PCR Core Reagents kit (Applied Biosystems, Foster City, CA) following the manufacturer’s instructions with some modifications. The reaction mixture consisted of 0.3 pM of each primer, lXSyerreen PCR buffer, 2.5 mM MgCl2, 0.2 pM of each deoxyribonucleoside triphosphate, 0.25 U AmpliTaq Gold DNA Polymerase, 0.1 U AmpErase UNG, and 10 ng environmental DNA. TaqMan real-time PCR assays were performed on a 7700 Sequence Detector System (Applied Biosystems, Foster City, CA) in 30 pl reactions containing 1X TaqMan Universal PCR Master Mix (Applied Biosystems, Foster City, CA), 0.3 pM forward primer, 0.3 pM or 0.9 pM reverse primer for cluster 2 or P. stutzeri, respectively, 0.6 pM 62 or 0.525 pM of the fluorogenic probe for cluster 2 or P. stutzeri, respectively, and 10 ng environmental DNA. The PCR cycling used for all reactions was the same as described in Chapter 2 (Griintzig et al., 2001). A melting curve analysis was performed from 60°C to 95°C on the Syerreen assay products to determine if only one product was formed during the real-time PCR amplification. Negative controls with no template DNA were run in each reaction. Additionally, controls with no probe were prepared for TaqMan assays. Data were analyzed with the SDS 2.1 software (Applied Biosystems, Foster City, CA). The specificity of the primer sets was tested using several clones retrieved in this study as positive and negative controls. Plasmid DNA (1 pg) extracted from the corresponding E. coli transfonnants was added as template DNA in this case. The nirS copy number of Specific clusters and of P. stutzeri was determined by comparison to a standard curve. Plasmid DNA from the following clones was used to generate standard curves for nirS clusters: NIRS4A C07 (cluster 1), NIRS4A F08 (cluster 2), NIRS4A A02 (cluster 3), NIRS4A E10 (cluster 5), NIRS4A G11 (cluster 6). Genomic DNA from marine isolate E4-2, previously identified as P. stutzeri (Braker et al., 2000), was used to construct standard curves for P. stuzeri nirS quantification. Standard curves were run in triplicate or quadruplicate, while duplicate reactions were performed for environmental DNA. Samples that exhibited CT values (cycle number at which the fluorescence emission crosses a defined threshold) equal to or higher than the negative controls were considered as below the detection limit. All results were normalized to the quantity of community DNA. 63 Denitrification capacity in sediments In order to determine denitrification capacity, microcosms were prepared from sediment samples from five different depths from each location (Carr Inlet: 0-2, 2-4, 10- 12, 34-38, and 135 cm. Shallow Budd Inlet: 0-2, 2-4, 10-12, 24-28, and 34-38 cm. Turning Basin: 0-2, 2-4, 10-12, 34-38, and 69 cm. Washington Margin: 0-2, 2-4, 10-12, 34-38, and 62 cm.). A homogeneous slurry was prepared for each Site and depth and further divided to set up individual microcosms, while a sample was conserved as unamended control. Each microcosm consisted of 50 ml Slurry and 110 ml headspace. The slurry contained 10 g sediment and artificial seawater (Kemp et al., 1993) to make up a final volume of 50 ml. Salinity was adjusted in Puget Sound microcosms to 28 ppt, the salinity of the area’s water. Three sets of microcosms were prepared with either 40 pM NaNO3, 400 pM NaNO3, or 160 pM O2. Acetylene (10%) was added to each microcosm to inhibit the reduction of nitrous oxide (Sorensen, 1978). All media and bottles used were flushed with He to displace O2. Microcosms were incubated at 20°C in the dark shaking at 220 RPM. A control not subjected to shaking was also prepared, as well as a control without added acetylene. Three microcosms of each type were incubated for 1, 5, and 26 h, respectively, after which a gas sample was taken and the sediment stored at -80°C for further analysis. Nitrous oxide concentrations in the headspace samples were determined on a Hewlett Packard 5890 Series 11 gas chromatograph (Hewlett Packard, Palo Alto, CA) with an electron capture detector by comparison to a standard curve. The partition coefficient of nitrous oxide (2.1 gas/water) was included in the determination of total micromoles of nitrous oxide produced. 64 Nitrous oxide production rates were determined for the incubations with 400 pM NaNO3, which presented a nearly linear accumulation of N20. T-RFLP analysis of selected microcosm samples was performed as described above. Results Biogeochemical characteristics of the sediments As expected for near Shore and continental shelf and Slope sediments with high sedimentary respiration rate (Brandes and Devol, 1995) oxygen and nitrate exhibited a narrow penetration depth into the sediments (Figure 3.2), being depleted at Similar depths ranging from 0.5 to 2 cm, consistent with previous measurements in the same areas (Brandes and Devol, 1995; Braker et al., 2000). Shallow Budd Inlet showed the highest oxygen concentration, reaching 435 pM, and extremely low nitrate concentrations, never higher than 1.7 pM, due to the extensive growth of algae in the overlying water at that Site. Turning Basing was characterized by a high sulfate reduction rate in the sediments (Table 3.1), twice as high as the next highest measured value in Shallow Budd Inlet and consistent with the consumption of sulfate observed in the pore water profiles (Figure 3.3). Nitrate was also extremely low at Turning Basin, never reaching concentrations higher than 0.7 pM. Carr Inlet sediments, on the other hand, had much higher nitrate concentrations than the other Sites in Puget Sound, reaching 14.8 pM. Besides individual differences, the three studied Sites in Puget Sound presented a dense benthic macrofauna, with invertebrates being Observed down to at least 38 cm depth, consistent with the high mixing rate of 43 cm2 y'1 and surface mixed layers of up to 22 cm observed in sediments 65 from other Puget Sound areas by analysis of excess 21OPb activity profiles (Carpenter and Peterson, 1989). Macrofauna was absent, on the other hand, in deep samples retrieved with the gravity corer (Turning Basin, 69 cm; Carr Inlet, 135 cm). The Washington Margin sample, in contrast, lies within a zone with a sharp decrease in sediment mixing coefficients, probably accounted for by the low dissolved oxygen concentration in waters over the slope from about 600 to 1400 m, which leads to the inhibition of macrobenthic activity (Carpenter and Peterson, 1989). This was corroborated by our observations of the sediment cores of this Site, showing noticeably less macrobenthic fauna present only down to about 15 cm in the sediments. Measured pore water oxygen concentrations, as expected, never exceeded 12 pM, more than an order of magnitude smaller than the measured concentration in any of the Puget Sound Sites (Figure 3.2). Nitrate, on the other hand, reached 44.6 pM, 3-fold higher than the highest concentration detected in Carr Inlet sediments. Denitrifier community structure by T-RFLP of nirS genes The denitrifier community structure changes along redox gradients were studied by analyzing heme cd, nitrite reductase gene (nirS) T-RF S from different depth sediments at the four sites. Restriction with Mspl yielded a higher number of T-RFS in the studied samples than HhaI and was therefore chosen for the subsequent data presentation. However, analysis performed with HhaI led to comparable results and further supported the conclusions. In Puget Sound samples, the pattern detected by T-RFLP was highly Similar in the sediment samples down to 37 cm depth within each Site (Figure 3.4), with T-RF richness from 22 to 41, 39 to 47, and 21 to 37 for Turning Basin, Shallow Budd Inlet and Carr 66 Inlet, respectively. A sharp decrease in T-RF richness was detected in deep samples retrieved below the bioturbated sediments in Turning Basin (5 T-RFS at 69 cm depth) and Carr Inlet (7 T-RF S at 135 cm depth). No samples below the bioturbated zone were retrieved from Shallow Budd Inlet. In contrast, Washington Margin sediments, presented a conserved nirS community structure only down to 10 cm in the sediments, with T-RF richness between 40 and 46. A Sharp decrease in T-RF richness was already detected at this site for the 36-37 cm deep sample (12 T-RF S), which lies below the bioturbated zone, as well as for the even deeper sample from 62 cm depth (15 T-RFS). Two T-RFs, 97 bp and 140 bp, were common to all sites and depths, and in Significant relative abundance in all cases (13.8 to 69.6% for T-RF of 97 bp and 1.0% to 22.9% for T-RF of 140 bp). Only 15 other T-RFS were detected in all areas, but not at all studied depths. The remaining T- RFS were absent in at least one site, with 37 T-RFS being characteristic of one sampling location only. Although some T-RFS were detected in sediments within and below the bioturbated zone, most T-RF S were detected only within the mixed sediments being absent in the unmixed zone. However, certain T-RFS were particular only to the deep unmixed sediments at one site (75 bp and 208 bp at 62 cm deep Washington Margin sediments, and 91 bp and 218 hp at 135 cm deep Carr Inlet sediments), although some were detected also in Shallow sediments in other areas. T-RF 116 bp was detected in shallow and deep sediments of Washington Margin, Turning Basin and Shallow Budd Inlet, however its relative abundance was Significantly higher in sediments below the bioturbated zone. Cluster analysis of the T-RFLP profiles grouped the samples according to Site and depth (Figure 3.5), with samples from unmixed sediments from all sites forming a 67 separate cluster. Sediments from the bioturbated depths at each location formed distinct clusters. This pattern was also observed when an ordination technique (principal component analysis, PCA) was applied to the T-RFLP profiles (Figure 3.6), in order to reduce the dimensionality of the data and summarize common correlations among variables (T-RFS) into fewer variables (components) (Dollhopf et al., 2004). The first two dimensions explained 47% of the total variability (26% and 21% in dimensions 1 and 2, respectively). T-RFLP profiles were grouped according to sampling location, however samples from unmixed sediments were grouped together, regardless of the site they were retrieved from, in agreement with the cluster analysis. PCA allows the identification of the main T-RFS responsible for the grouping of the samples, based on the size and direction of the eigenvectors representing each T-RF. Eigenvectors in the Vicinity of a sample, are most likely affecting it, and the size of the vector is a measure of the magnitude of the effect. T-RFS present only in one Site or presenting a significantly higher relative abundance at that site, were mainly responsible for the dispersion of the samples from different locations (data not Shown). So, T-RFS 236 bp and 62 bp, present only or at Significantly higher abundance, respectively, in Washington Margin samples, were mainly responsible for the separation of the samples from this site. A set of additional T-RF S detected only in Washington Margin, but with low relative abundance, also had a significant effect in the separation of these samples, e.g. 268, 350, and 428 bp. Carr Inlet samples were mainly affected by T-RFS 87 and 406 bp, present only at this Site, as well as by a number of additional T-RFS of low relative abundance, e.g. 410, 393, and 386 bp. Shallow Budd Inlet and Turning Basin were ordinated closely together, in 68 concordance with cluster analysis. The main T-RFS responsible for their ordination were T-RFS present only at both of these Sites, e.g. 80, 110, 111, 178, and 221 bp. Phylogenetic analysis of nirS clone sequences Out of 188 sequenced clones, 137 were identified as nitrite reductase heme cdl (nirS) sequences longer than 600 bp and used, after translation, to build a phylogenetic tree by neighbor-joining analysis including also closely related sequences from public databases (Figures 3.7, 3.8, and 3.9). A wide range of sequence divergence between nirS clones was observed, with clones being from 49.6 to 100% Similar. Six individual clusters of highly Similar nirS sequences, generating only one or two hypothetical T-RFS by in silica digestion with the same restriction enzymes used in the T-RFLP analysis, were identified (Table 3.3). These clusters were later targeted for quantification by real- time PCR (Table 3.4). Most nirS clones were only distantly related to any cultured denitrifier, with the exception of clones grouped in cluster 4 (Table 3.3) being 80.4 to 83.3% Similar to Raseabacter denitrificans and clone NIRS4B A12 being 77.7 and 78.8% similar to Paracoccus pantatrophus and Paracoccus denitrificans PD1222, respectively. Two large clusters of sequences (subtrees A and B) contained almost all nirS clones retrieved in this study (96% of sequences), as well as other environmental clones from related and unrelated areas, but no cultured denitrifier with the exception of Azaarcus talulyticus 2F B6. All clones were at least 75% similar to environmental nirS clones from other studies with the relatedness reaching values of up to 98.5% Similarity for NIRS4B F09 and pA33, a sediment clone retrieved from a different location in Puget Sound (Braker et al., 2000). Clones in subtrees A and B are most closely related to sequences found in these previously studied Puget Sound sediments as well as Washington Margin 69 sediments (Braker et al., 2000), in a Baltic Sea cyanobacterial aggregate (Tuomainen et al., 2003), in the Baltic Sea water column (Hannig et al., 2006), and in sediments from Huntington Beach, Califomia (Santoro et al., 2006). Furthermore, they are also related (up to 97.1% Similarity) to an mRNA sequence (clone ANIS-54) detected in sediments of the River Colne estuary in the United Kingdom (Nogales et al., 2002). Arabian Sea water column clones (Jayakumar et al., 2004) and an Eastern South Pacific water column clone from the oxygen minimum zone (Castro-Gonzalez et al., 2005) clustered with clones in subtree B, however more distantly. 1n silica digestion of these clones with the same restriction enzymes was compared to the measured T-RFLP pattern at Shallow Budd Inlet (Figure 3.10) (see Figure 4.17 for HhaI results). A shift between in silica and real T-RF Sizes of up to 6 bp was consistently detected, coincident with the Shift observed between samples analyzed by gel and capillary electrophoresis (see Materials and Methods). Most of the clones (89%) generated a hypothetical T-RF of 102 bp, corresponding to the main T-RF at all depths (97 bp), thus indicating that this abundant T-RF is not representing one dominant nirS sequence, but rather a group of different sequences with a coincident restriction site. Only three other hypothetical T-RFS were generated, 68, 140, and 172 bp, corresponding to T- RF S of 62, 136, and 169 bp, respectively. The latter T-RF (hypothetical T-RF of 172 bp) was generated by 12 clones in comparison to 1 and 2 clones generating hypothetical T- RFS of 140 and 68 bp, respectively, in accordance with the higher relative abundance of that T-RF in the observed profile. Hypothetical T-RF 140 bp was generated by only one clone (NIRS4B A12); hence its low abundance might have interfered with its detection by T-RF LP, although it was present in other studied areas (data not Shown). On the other 70 hand, although partial restriction by the enzyme might be responsible for the detection of certain additional T-RFS in the T-RFLP profile compared to the clone library, the T-RF sizes between 62 and 97 bp could not be explained by partial restriction of the clones generating a hypothetical 68 bp fragment. Therefore, it is assumed that the lack of identification of clones corresponding to the additional T-RF sizes is principally due to an incomplete coverage by the clone library of the nirS diversity at the studied site. Quantification of clone clusters by real-time PCR Specific primer sets (Syerreen methodology) or primer-probe sets (TaqMan methodology) (Table 3.4) were designed to quantify the clusters identified on the phylogenetic tree of nirS sequences from Shallow Budd Inlet sediments (Table 3.3) by real-time PCR. The goal was to better describe the abundance in the environment of some of the groups of novel sequences unrelated to cultured denitrifiers, as T-RFLP is only considered semiquantitative (Liu et al., 1997), due to the possible PCR bias towards different targets. These particular clusters were chosen as they generated only one or two individual hypothetical T-RFS by in silica digestion and were related closely enough to allow the design of primers against the whole group, but dissimilar from the other clones to avoid cross amplification of different sequences. All clusters were amplified using Syerreen methodology, except for cluster 2, which was quantified by TaqMan real-time PCR, as the primer set used for Syerreen real-time PCR led to non-specific amplification of cluster 3 clones and was therefore replaced. Also, sets of primers designed against cluster 4, did not amplify any clone in these clusters and were therefore discarded. In addition, the previously designed primer-probe set against P. stutzeri nirS (Grttntzig et al., 2001) was also used. The detection limit was extremely low for each 71 primer or primer-probe set in all real-time PCR amplifications, always below seven nirS copies per 10 ng community DNA, with the exception of the quantification of cluster 2 nirS, which presented a detection limit of 65 copies per 10 ng community DNA. In Shallow Budd Inlet, as well as in the geographically close Turning Basin Site, cluster 2 was the most abundant nirS cluster (4 X 106 nirS copies per pg community DNA at Shallow Budd Inlet, 24-25 cm) (Figure 3.11) as expected by the high number of clones corresponding to this cluster (28 clones). Clusters l and 3 presented about an order of magnitude lower abundances than cluster 2, consistent with the lower number of clones identified in these clusters. The lowest abundances corresponded to clusters 5 and 6, represented by only 4 and 3 clones, respectively. In correlation with the T-RF LP profiles, the abundance of each cluster did not vary Significantly at different depths within the mixed sediments (down to 37 cm), however the nirS copy number diminished 100- to 1000-fold in the deep Turning Basin sediments (69 cm). Significantly lower abundances of clusters 1, 3 and 6 were detected in Carr Inlet sediments (Figure 3.12 and Table 4.4), compared to Shallow Budd Inlet and Turning Basin. Furthermore, cluster 2 was below detection limit at all depths, except at 10-10.5 cm, at this site. However, consistent with the results at the other Puget Sound Sites, the abundance of clusters 1 and 3 did not vary significantly within the mixed sediments and was below detection limit at 135 cm depth. Cluster 6, on the other hand, presented a low relative abundance and was below detection limit already at 10-10.5 cm depth. Much lower abundances of the quantified clusters were detected in Washington Margin sediments (Figure 3.12), often being below the detection limit. This is in agreement with the T-RFLP results, which identified a more distantly related community at this Site. Interestingly, P. stutzeri nirS was detected at a relatively 72 high abundance of up to 1.5 X 106 nirS copies per pg community DNA in Washington Margin sediments. This is equivalent to 6.6 ng P. stutzeri DNA per pg community DNA, i.e., 0.66% P. stutzeri DNA in total community DNA. In addition, its abundance in Carr Inlet (up to 1.4 X 105 nirS copies per pg community DNA) was Significantly higher than the abundance of the quantified clusters (Table 4.4). On the other hand, P. stutzeri nirS in Shallow Budd Inlet (2.7 x 104 nirS copies per pg community DNA) and Turning Basin (3.2 X 105 nirS copies per pg community DNA) was generally less abundant than the nirS clusters identified in the Shallow Budd Inlet clone library. In silica digestion of P. stutzeri ZoBell nirS with Mspl generates a hypothetical T-RF of 141bp. A T-RF of 137 bp, which could correspond to P. stutzeri was observed at significant relative abundance in the T-RFLP profiles of Washington Margin (up to 14.1%) and Carr Inlet (up to 6.6%) (Figure 3.4), but at lower abundance in Shallow Budd Inlet and Turning Basin, in accordance with the real-time PCR results. Besides being more abundant in marine sediments than previously measured (Griintzig et al., 2001), a high abundance of P. stutzeri nirS was also detected in deep unmixed sediments from Turning Basin and Washington Margin, similar to the abundance detected in shallow sediments. This again is consistent with the presence of T-RF 137 bp in deep Washington Margin sediments. Denitrification capacity of sediments studied in microcosms Microcosms of sediments from each location and five different depths were prepared to study their initial denitrification capacity, measured as nitrous oxide production after nitrate addition. Puget Sound sediments from all sites and down to 38 cm depth were able to denitrify immediately, as observed by significant nitrous oxide production after NaNO3 amendment (Figures 3.13, 3.14, and 3.15). On the other hand, in 73 Washington Margin sediments, Significant denitrification capacity was detected down to 12 cm depth only, with sediments from 34 to 38 cm depth showing a minor accumulation of nitrous oxide (Figure 3.16). No nitrous oxide accumulation was detected in microcosms amended with oxygen, as expected, neither in controls lacking acetylene. Addition of 400 pM NaNO3 led to an accumulation of nitrous oxide close to linearity in all cases. Therefore, it was used to calculate nitrous oxide production rates in the microcosms (Table 3.5). The highest rate was measured in Shallow Budd Inlet sediments from 2 to 4 cm depth (0.133 N20 h'l microcosm'l). NO lag time was detected before initiation of nitrous oxide production, though a slight increase in production rate was in some cases observed for the deepest mixed sediments devoid of nitrate, after addition of 400 pM NaNO3 (Shallow Budd Inlet, 24-28 cm; Carr Inlet, 2-4 cm, 10-12 cm, 34-38 cm). The accumulation rate of nitrous oxide reached a plateau after 5 h in microcosms amended with 40 pM NaNOg, suggesting a complete consumption of the added nitrate after this time. Carr Inlet top sediment samples (0-2 and 2-4 cm) amended with 40 pM NaNO3 accumulated about 1 pmol nitrous oxide after 26 h incubation. This represents a complete conversion of the added N-NO3' to N-N2O. However, deeper samples within the mixed zone of this Site and samples from the other Sites generally reached a plateau at values below 0.45 pmol, indicating only partial conversion to nitrous oxide, maybe due to concomitant occurrence of dissimilatory nitrate reduction to ammonium (DNRA). In order to study the changes in community structure that occurred during incubation with nitrate, T-RFLP analysis was done on the sediment from selected microcosms from Shallow Budd Inlet, as these sediments had the highest accumulation of nitrous oxide. The analyzed microcosms had been amended with 400 pM NaNO3, as 74 nitrate seemed not have been depleted even after 26 h incubation with this initial concentration. Microcosms from the five studied depths (0-2, 2-4, 10-12, 24-28, and 34- 38 cm) incubated for 26 h were analyzed, as well as microcosms from 2-4 cm depth incubated for different times (0, 5 and 26 h). The results were compared to the T-RFLP profiles generated from the original sediments retrieved from Shallow Budd Inlet (Figure 3.17). A noteable change in community structure was detected in all microcosm samples compared to the original sediments, with some T-RFS disappearing, appearing, or changing in relative abundance after incubation. Besides the disappearance of some T- RFS there was a general increase in richness after incubation (52 to 59 T-RFs after 26 h incubation, compared to 39 to 47 T-RFS in the original sediments). Though different from the original sediments, the community structure of incubated sediments from different depths seemed highly Similar. Furthermore, no significant variations were detected with different incubation times. This is further supported by PCA ordination of these T-RFLP profiles (Figure 3.18), which Shows a clear separation of profiles from original sediments and from microcosm sediments along the first dimension, which explained 59% of the total variability. The second dimension explained an additional 11% of the variability. Samples incubated for 0, 5 and 26 h were grouped closely together, indicating that the main changes in community structure had not occurred during the incubation time, but rather due to the original set up of the microcosms. Subcores used for DNA extraction were immediately sectioned and samples stored at -20°C, while subcores used for microcosm set up were kept at 4°C for 3 days after retrieval, maybe leading to changes in the community structure. In addition, the treatment of sediments for microcosm set up (e.g. dilution with artificial seawater, mixing, etc.) might have further affected the 75 denitrifier community structure. T-RFS present only in the original sediments, in the incubated sediments, or whose relative abundance changed noteably, were mainly responsible for the separation of the samples. Several T-RFS present only in the original samples (e.g. 72, 79, 80, 89, 108, 110, 111 bp) had a strong influence in the separation of the original samples from the rest. In addition, T-RF S 84 and 97 bp, the latter representing most of the novel nirS clones retrieved from this environment, also strongly influenced the separation, as their relative abundance diminished after incubation. On the other hand, T—RF 103 bp, present only in the microcosms and absent in the original sediments, and T- RFS 137, 167 and 169 bp, present at Significantly higher abundance in the microcosms, had a strong effect on the separation of the microcosm samples from the rest. Interestingly, T-RF 137 bp coincides with P. stutzeri, a commonly isolated bacteria, and T-RF 169 bp coincides with twelve of the retrieved clones from this Site, with Raseabacter denitrificans, Azaarcus talulyticus, and the marine denitrifying isolate C10- 1, previously retrieved from sediments at the Washington margin (Braker et al., 2000). 76 vv IX! who'd m H‘- n g/ Gm! Pnimll - C I I t \ arr n e «I.» Margin .- 0 \ j . w, I .. “IQ “ Turning Basin 3‘ a Q 1» ,,_ it» .. 420‘ "3' “’2' :3 m Shallow BUdd Inlet Figure 3.1. Locations of sampling stations at Puget Sound and the Washington margin slope. 77 Table 3.1. Locations and characteristics of sampling stations. Sulfate Water Organic . . reduction Station Location depth carbon 1:3? $318133), rate lrrigation‘ (m) (%) (mmoles m'2 day") Shallow 47°04.99'N high Budd Inlet 122°54.16'W 3 2'5 '9'15 26'2 '2 (38 cm) Turning 47°03.13'N high Basin 122°54.45'W '2 3'2 ”'80 2"2 25 (38 cm) 47°17.21'N high Carr Inlet 12204295,“, 84 1.8 16.80 29.6 7.5 (38 cm) Washington 46°25.57'N medium Margin 12404150.“, 1138 3.1 3.65 34.386 0.04 (15 cm) °Numbers in parenthesis indicate lowest depth at which macrobenthic fauna was still detected. 78 Table 3.2. Sediment samples analyzed from different sites. Sediment depth analyzed (cm) Analysis Shallow Budd Washington method Turning Basin Carr Inlet Inleta Margin T-RF LP 005 005 0-0.5 0-0.5 1-1.5 1-1.5 1-1.5 1-1.5 2-2.5 2-2.5 2-2.5 2-2.5 10-10.5 10-10.5 10-10.5 10-10.5 24-25 36-37 36-37 36-37 36-37 69 135 62 Cloning 2-2.5 Real-time PCR 0-0.5 0-0.5 0-0.5 0-0.5 0.5-1 0.5-1 0.5-1 0.5-1 1-1.5 1-1.5 1-1.5 l-l.5 1.5-2 1.5-2 1.5-2 1.5-2 2-2.5 2-2.5 2-2.5 225 10-10.5 10-10.5 10-10.5 10-10.5 24-25 36-37 36-37 36-37 36-37 69 135 62 Microcosms 0-2 0-2 0-2 0-2 2-4 2-4 2-4 2-4 10-12 10-12 10-12 10-12 24-28 34-38 34-38 34-38 34-38 69 135 62 "No deep sample (retrieved with gravity corer) was obtained at this Site. 79 Figure 3.2. Pore water profiles for O2 and N03' for Puget Sound (Shallow Budd Inlet, Turning Basin, and Carr Inlet) and Washington Margin sediments. 80 Moron)? ? 1' L 1 1° 1.2 '.‘ 1° 1.3 20 500 02 (”Mi 0 100 200 390. 400 .2- 0 2l 4- 64 81 1° ‘ Shallow Budd Inlet Depth (mm) 12‘ 141 N03'(pM)0 2 4 6 6 10 12 14 16 18 20 02(pM) q 20 40 60 80 100 120 140169 Depth (mm) 12 Turning Basin 14 , , , , . . , , 1103'me 2 4 6 8 1o 12 14 16 18 180 20 021“"10 50 100 150 __200 250 .4 1 r 1 -2 Depth (mm) 14 ‘ Carr Inlet 1103111141) 0 1o 20 30 40 02011111) 0 - so 100 150 Depth (mm) 12 Washington Margin 14 , . . 81 50 200 +0, + N03- Figure 3.3. Pore water profiles for Fe(II), 8042', Mn(II), and NH4+ for Puget Sound (Shallow Budd Inlet, Turning Basin, and Carr Inlet) and Washington Margin sediments. 82 renown), 30421111141), Mn(ll)(pM)0 1o 20 30 4o NH"(pM) o 50 100 150 200 250 I) Shallow Budd ’5‘ 8 15 20 . , . . FO(II)(1.IM), 3042111101), Mn(uxnmo s 10 15 211 215 so NH4+IPMI 0 so 100 150 200 250 300 350 L Turning Basin 10 Depth (cm) 15 L A IL L ‘1 A A AA A A V V v r v v y 77 20- . . . rotunda),NH4ttnM),so42'(mM) 0 1o 20 30 40 so so 70 so so 100 Mn (II) (11M) 0 so 100 150 200 250 300 oi. A 5 " E 8. Carr Inlet 5 1o 1% D 15- 20 ~ . . Fe(l|)(pM), SO42'(mM), Mn(II)(pM) L 1.0 210 :10 40 50 NH 4'01") 0 so 100 150 200 250 '8 Washington 3' + Fe(ll) Margin 3 + MM“) 0 '0' + NH4 —0— 30‘2' 83 0% 20% 40% 60% 80% 100% g, 00.5 1:162 1:365 g 1-1-5 1370 1:175 = 2'25 I79 1:80 210-105 " 3 36-37 E1184 E1187 :5: 62 12189 12191 A 3 MS I94 G97 . E § 1-1.5 5102 121105 E F 2-2.5 C1108 821110 10-10.5 E, 5 36-37 C1111 111116 E 135 13134 5137 2 3. I140 I154 C - g 3 ‘13:: I167 @169 '5 E 2.25 C1175 I178 :21 E10-10.5 I199 113208 0 = 24-25 ' o 2 36-37 @218 [3221 In E1228 I236 g 005 I265 I275 a :12: 111344 111376 E10403 [1403 13406 g 36-37 8416 I474 " 59 C1484 C1598 Relative abundance of T-RFs (%) Figure 3.4. T-RFLP profiles obtained by amplification of nirS genes from Puget Sound and Washington Margin marine sediments from various depths. Presented T-RFS were generated after restriction with Mspl. Numbers in legend indicate T-RF length in base pairs for fragments representing more than 1.5% relative abundance. 84 2.58 1.82 1.28 0.538 TB 58 CI 135 WM 62 WM 315-3? Cl 383? CI 1010.5 CI 22.5 CI 1-1.5 CI 00.5 WM 101 0.5 WM 225 WM 1-1.5 WM 005 TB 383? TB 0-0.5 TB 1010.5 TB 225 TB 1-1.5 SB 36-37 SB 24-25 SB 1-1.5 SB 10-1 0.5 SB 22.5 SB 00.5 Figure 3.5. Dendrogram obtained by cluster analysis of T-RFLP profiles. Hierarchical clustering analysis was performed applying Ward’s method on the Hellinger distances between T-RFLP profiles from Puget Sound and Washington Margin sediments from various depths. TB, Turning Basin; SB, Shallow Budd Inlet; CI, Carr Inlet; WM, Washington Margin. Numbers next to sample names indicate sediment depth in cm. 85 8 4 ‘ A 6 ~ A ‘ 4 d ‘ C; 1 ’A x . 1:\_ , 2 ‘ x deep sediments 0 TB 9‘. A n I SB 2 -B I: 5 -2 0 2 4 6 8 1b 4 CI "" -2 - x WM 1: In .4 4 *5 >51 —8 — x Dim 1 (26%) Figure 3.6. PCA ordination of T-RFLPS of nirS genes from Puget Sound and Washington Margin sediments. TB, Turning Basin; SB, Shallow Budd Inlet; CI, Carr Inlet; WM, Washington Margin. Percentage of variance explained by each dimension is presented in brackets. Oval indicates the deep samples from below the bioturbated zone. Dim, dimension. 86 ‘ Subtree A J L——- Subtree B i ‘— Huntington Beach sed. clone th 46 (00159571) 7 L— Thiabaciilus denitrificans (NC 007404) J~]—— Arabian Sea water col. clone GB40—4F (AY336818) l I 99[ “" Pseudomonas aeruginosa PAO1 (AEOO4488) . Pseudomonas fluorescens (AF 1 97466) ' rl—— Marine denitrifying isolate 89-12 (AJ2483Q3) gél (— - Marine denitrifying isolate 04 14 (AJ248395) 73;, 1* Marinabacter aquaeoleiVTB (NZ AALGOtOOOOOZ) . 100'—— Marine denitrifying isolate C10-1 (AJ248394) j I 4122- * Cupriavidus necatar (X91394) - T?“ ' Ralstonia metallidurans CH34 (NZ AAAI03000011) | l 5L '— Huntington Beach sed. clone we so (00159648) (1091:..— NIRS4A E11 _ Huntington Beach sed clone th 4C (DQ159496) 3 ”7 " Huntington Beach sed. clone hbD 5A (00159611) . 7— Huntington Beach sed. clone hbD 2F (00159599) .2 100.3 59 ll - Huntington Beach sed. clone hbD 80 (00159624) l ' ~ - - —~ Silicibacterpameroyi ass-3 (CP000032) 100 f Paracoccus denitrificans PD1222 (U05002) .__.__1 _;fi 1 903 ‘ Paracoccus pantatraphus (AJ4014S2) 1 i; 7 NIRS48 A12 ‘ 53 J1 "'77 Raseabacter denitrificans (AJ224911) a}, - Huntington Beach sed done has 213 (DQ159527) 8100 Cluster 4 (4 clones) fi—r “Tm Hydrogenobacter thermophilus (A8210046) 7 . 7 J _ Arabian Sea water col. clone GB40-5A (AY336820) tQQjTT—T— Puget Sound sed. clone p816 (AJ2484ZO) 100?* -— * * -- -—— Wash. margin sed. clone wA20 (AJ248428) 100; Puget Sound sed. clone pB6 (N248419) _ _, j _ __ j ' Puget Sound sed. clone pA17 (AJ248407) I 1oo' l Baltic Sea water col. clone M150—85 (00072183) 100 Baltic Sea water col. clone MOO-54 (00072203) 67 i— Thauera selenatis AX (AY078264) J ‘1 Thauera aramatica K172 (AY078256) I 1 100:‘ j ’ Pseudomonas stutzeriZoBell (X56813) 95 U Marine denitrifying isolate E4-2 (AJ248398) i 100‘ Alca/igenes faecalis A15 (AJ224913) Azaspirillum braSi/ense Sp7 (AJ224912) spa—4 01 531 Figure 3.7. Phylogenetic tree showing the affiliation of nirS clone sequences retrieved from Shallow Budd Inlet, Puget Sound, sediments to selected reference sequences. The tree was generated by neighbor-joining analysis of approximately 215 aminoacids with Azaspirillum brasilense Sp7 as the outgroup. Values at branch points indicate the percentage of 100 replicate trees supporting that branch. Bootstrap values below 50% were omitted. Names of clones retrieved in this study begin with NIRS4A and NIRS4B. Clusters of highly Similar nirS clone sequences are identified by numbers, with quantity of clones in the cluster indicated in brackets. For graphical purposes, detail of subtree A and subtree B is presented in Figure 3.8 and 3.9, respectively. The scale bar represents 0.1 substitutions per sequence position. sed, sediment; Wash, Washington; col., column. 87 r" Cluster 2 (28 clones) l‘: NIRS4A H01 NIRS4A 003 NIRS4A 802 . Li— NIRS4A 003 H){__ gNIRS4A H07 NIRS4A H09 NIRS4B (311 NIRS4B 011 62L)— NIRS4A 012 50 _‘ NIRS4A H02 lH+ ~ NIRS4A 010 ;— NIRS4AG01 )J Cluster 3 $11 clones) NIRS4B E 0 ) 98)— NIRS4A 608 9 9 )3") NIRS4B 008 NIRS4A 005 I NIRS4B H10 NIRS4B c302 NIRS4B 501 L— NIRS4B E02 70 River Colne estuary sed. clone ANlS-54 (AJ440470) JE Puget Sound sed. clone pA12 (AJ248405) )_ NIRS4A Goz : NIRS4A F11 NIRS4AA11 : NIRS4A F05 1 50: NIRS4A G10 H— NIRS4B 004 NIRS4B 512 NIRS4B c304 NIRS4A H03 NIRS4A 505 NIRS4B F06 NIRS4B co3 'uNlRS4AA05 60 L- NIRS4B 007 83 r NIRS4B B08 [90“ NIRS4B F04 , NIRS4A F10 . ._W_ NIRS4A E05 ! FL;- NIRS4B 011 L Baltic Sea cyanob. aggregate clone BS1270 (AJ457196) i NIRS4B 010 )F’NIRS4B 002 934 NIRS4AC10 F l ‘ NIRS4A 006 i NIRS4A 001 J, m- NIRS4A H10 1 63 71)) “NIRS4A C08 ) 832r NIRS4A E01 l 52 NIRS4A F12 ; NIRS4A 809 i “—96 L I— NIRS4A F02 _ ___ . : 109,5 ti” NIRS4A BO4 ' NIRS4B C09 L47" — NIRS4B C10 '. I“ Puget Sound sed. clone p346 (AJ248422) r, 65 “ “7) r“ NIRS4AG12 100 79g; Puget Sou nd sed. clone p849 (AJ248423) Puget Sound sed clone p866 (AJ248424) %_______ Baltic Sea water col. clone M60- 76 (00072204) ‘ NIRS4A 604 50 _-___ "fl 7 Huntington Beach sed. clone th 86 (00159557) F—a—m— ~ 4 010 Figure 3.8. Detail of Subtree A as described in Figure 3.7. sed, sediment; col., column; cyanob., cyanobacterial. 88 56I—' — Huntington Beach sed. clone th 66 (00159579) 2, ‘——— Huntington Beach sed. clone th 7G (00159582) ' “ "‘“‘ Huntington Beach sed. clone hbD 4C (00159607) 50 0““ Arabian Sea water col. clone 6857-78 (AY336858) :57 3 P— ES. Pacif. OMZ water col. clone AC100-135 (AJ811517) * "‘7 Arabian Sea water col. clone V483—3EE (AY336925) Huntington Beach sed. clone hbD 2G (00159600) - ‘- Arabian Sea water col. clone V483-8H (AY336912) ‘ Azoarcus talulyticus 2FB6 (AY078272) _ 85 l—‘ Puget Sound sed. clone p820 (AJ248421) ) , gl 1 “'7'“ ' Baltic Sea water col. clone M60-145 (00072215) 95 l T___ Arabian Sea water col. clone V483-4E (AY336904) f 51909 L“ Arabian Sea water col. clone (3840-6A (AY336822) l 52 " “7 Cluster 6 (3 clones) ’ 551 ' Puget Sound sed. clone pA5 (AJ248403) -- Baltic Sea water col. clone M60-165 (00072223) I Huntington Beach sed. clone th 8F (00159664) ’ F _._ NIRS4A F07 ) 96l NIRS4A 808 L. 84 fl NIRS4A 011 ) [-991 *- NIRS4A 806 a; L— Huntington Beach sed. clone th 9F (00159587) )— r— NIRS4A 810 100 Huntington Beach sed. clone th 6F (00159546) 64 52) Wash. margin sed. clone wA15 (AJ248427) ” 60 l' NIRS48 F09 100 L Puget Sound sed. clone pA33 (AJ248412) — ‘~ Baltic Sea water col. clone M60-156 (00072220) ” 7 Wash. margin sed. clone wF16 (AJ248437) 10 Huntington Beach sed. clone th 4C (00159534) 50; 7 Cluster 5 (4 clones) )Ew) Huntington Beach sed. clone th 70 (00159549) A 100 Huntington Beach sed. clone th 1E (00159522) as! 57 - NIRS4A H05 | NIRS4A C01 82 F NIRS4A E12 Cluster1 (21 clones) Figure 3.9. Detail of Subtree B as described in Figure 3.7. sed., sediment; Wash., Washington; 001., column; E.S. Pacif. OMZ, Eastern South Pacific Oxygen Minimum Zone. 89 ) 062(68) i I72 ' 074 D 79 El 80 m 84 In 89 El 91 94 097(102) D102 0105 8108 E 110 II] 111 [0116 0137 B 140 I167 0169(172) i 376 0% 20% 40% 60% 80% 100% 9416 I474 Relative abundance of T-RFs (%) 9598 1-1.5 2-2.5 10-10.5 24-25 Depth in sediments (cm) 36-37 Figure 3.10. Comparison of T-RF LP profiles obtained by amplification of nirS genes from Shallow Budd Inlet, Puget Sound, sediments from various depths to distribution of T-RF sizes obtained by in silico digestion of nirS clones retrieved from that same site (sediment depth 2-2.5 cm). Presented T-RFs (real and hypothetical) were generated by restriction with Mspl. Numbers in legend indicate real T-RF lengths in base pairs for fragments representing more than 1% relative abundance. Values in brackets indicate corresponding T-RF size generated by in silico digestion of clones. One clone generated a hypothetical T-RF of 140 bp (plain grey), however the corresponding real T-RF (136bp) was not detected at this site and was not included in the legend. 90 Table 3.3. Description of clusters identified in nirS phylogenetic tree. . . In silica In silica Minimum Cluster Clones in cluster similarity generated generated name (0/ )a T-RFs with T-RFs with 0 Mspl HhaIb NIRS4A NIRS4B A03 A09 Al 1 B03, C07, C09, ’ ’ ’ Cluster 1 004, 1303, 509, C01’ C05’ E04’ 98 102 7° (20” 23" G07 G09 H08 E06, E08, F05, (1) ’ ’ F07, G05, G08 A02, A05, A06, A06, A09, A10, A07, A10, BOZ, BOS, C02, C04, B1 1, C04, C06, 36 (21), 234 Chm“ 2 C06, 012, 007, 009, 805, 1307, 90 102 (7) D08, F08, H06 F02, F12, G09, H1 1 A02, A04, E02, A08, D02, F01 , 36 (7), 234 Chm” 3 F03, F06, F09 F08, 606 98 '02 (4) Cluster 4 307, D09, H1 1 E09 97 172 590 Cluster 5 E10 A04, C07, El 1 98 102 70 Cluster 6 011 003, 010 98 102 47 aSimilarity is inferred from Poisson-corrected distances between amino acid sequences. bNumber of clones generating a specific T-RF is indicated in parenthesis, when more than one T-RF size is observed. 91 Table 3.4. Primers and probes used to quantify the nirS gene of specific clusters and strains by real-time PCR. Target Primer or probea Sequence (5’-3’) Positionc Method Cluster 1 nirSRTl F ATGGATCCCAAGTTTGGTCC 1159-1178 Syerreen nirS RTl R TCGAGAACGCGTACGACT’I‘T 1295-1276 Syerreen Cluster 2 nirSTaq2F GGATAGCACGCACCGCTATT 1023-1042 TaqMan nirSTanpr CATGACGGCCGCGAACAAGTCC 1044-1065 TaqMan nirSTaq2R TCGATCACCGCGATC'ITGT 1085-1067 TaqMan Cluster 3 nirSRT3F CGATGTGACTGAGATI‘CCTCATC 1115-1138 Syerreen nirSRT3R AGGCATGCTCGGGATTTTC 1273-1255 Syerreen Cluster 5 nirSRTSF AATCTCGAAGCACTGGTGGAT 1099-1119 Syerreen nirSRTS R GGATGACCGTCGGGATCA 1247-123 0 Syerreen Cluster 6 nirSRT6F GACCGTAAACTGGCTGCTCTT 1093-11 13 Syerreen nirSRT6R ATGGCCCTCAGGGTCTGTT 1245-1227 Syerreen P. stutzeri nirSTaqutFb ACAAGGAGCACAACTGGAAGGT 1259-1280 T aqMan nirSTaqutprb GGCAACCTGTTCGTCAAGACCCA 1 309-133 1 TaqMan nirSTaqutRb CGCGTCGGCCCAGA 1362-1349 TaqMan a F, forward primer; R, reverse primer; pr, probe. b As in Chapter 2 and Grilntzig et al., 200]. ° Positions in the m'rS gene of Pseudomonas stutzeri ZoBell EMBL X56813. 92 Shallow Budd Inlet nirS abundance (copies per microgram community DNA) 1.E+O1 1.E+02 1.E+03 1.E+O4 1.E+05 1.E+06 1.E+07 5)---. __fl,, E 10 _, ___.-.-——- —-~— 9. 5 15~~vi~ ’ s \ I \ 1: 20-#’1 E g 25 -———~ - — “ ‘” “ .99 3 3° ‘ / / 7 35 -———-— ‘ -—~ teem-r it 1131 40 Turning Basin nirS abundance (copies per microgram community DNA) 1.E+01 1.E+02 1.E+03 1.E+O4 1.E+05 1.E+06 1.E+07 0 J ‘1 1O _H____.._~.. E 20 - 2. 5 3 a 0 ,3 40 _ -I—Cluster1 ‘E -Ei—Cluster2 ° 50 — — .§ +C|uster3 E 50 — +Cluster5 70 _ —A——Cluster6 +P.stutzeri 80 Figure 3.11. Cluster specific and P. stutzeri nirS gene abundance in Shallow Budd Inlet and Turning Basin sediments from different depths as measured by real-time PCR. 93 Carr Inlet nirS abundance (copies per microgram community DNA) 1.E+O1 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06 1.E+07 0 1991.999, .. l 10' E 3 I: 20 / // 3 ¢ 3 30 5 9/ J/ 0 .§ 40 '0 \ o \\ tn \ 130 140 Washington Margin nirS abundance (copies per microgram community DNA) 1.E+01 1.E+02 1.E+03 1.E+O4 1.E+05 1.E+06 1.E+07 0 10 — g 20 - .: § 30 - 1: +Cluster1 g 40 - ~8—Cluster2 % 50 +Cluster3 3 —x—Cluster5 so . - —A—Cluster6 +P.stutzeri 70 Figure 3.12. Cluster specific and P. stutzeri nirS gene abundance in Carr Inlet and Washington Margin sediments from different depths as measured by real-time PCR. Error bars represent standard deviation of duplicate assays. Missing values, represented as discontinuities in plots, were below detection limit. Note that all values were below detection limit for 135 cm deep Carr Inlet sediments. 94 3.5 Shallow Budd Inlet N 01 l N 4 + SAhighN --O-- SAlowN "O" SA02 + SBhighN --><-- SBlowN ---><-- 8802 —O— SChighN —-O-- SClowN ~O-- SCOZ ——)K— SDhighN --)K-- SDIowN --)K-- 8002 —I— SEhighN —-l-- SElowN -----l SE02 .3 U1 1 Nitrous oxide (micromoles I microcosm) 0.5 - 0 10 20 30 Time (h) Figure 3.13. Nitrous oxide production in microcosms of Shallow Budd Inlet (Puget Sound) sediments from various depths after addition of different electron acceptors. Sediment depth is indicated by letters A through E (A, 0-2 cm; B, 2-4 cm; C, 10-12 cm; D, 24-28 cm; B, 34-38 cm). Different microcosm sets were amended with 400 M NaNO3 (highN), 4O uM NaNO3 (lowN) or 160 uM 02 (02). 95 2.5 Carr Inlet A 2 — E at O 0 2 .2 E 9, 1.5 — 2 O . E + CAhighN 2 --o- - CAlowN ‘é’ -----o 0A02 § ,f - -><- - CBlowN g ,5.” ---><-- 0B02 m ,9?" + CChighN g 'l'i’. "C'- CCIOWN E 1’ ----e 0002 z + CDhighN - -)K- - CDlowN --->i<-- 0002 + CEhighN - ---- 0151ch ---n- 0502 Figure 3.14. Nitrous oxide production in microcosms of Carr Inlet (Puget Sound) sediments from various depths after addition of different electron acceptors. Sediment depth is indicated by letters A through E (A, 0-2 cm; B, 2-4 cm; C, 10-12 cm; D, 34-38 cm; E, 135 cm). Different microcosm sets were amended with 400 M NaNO; (highN), 40 M NaN03 (lowN) or 160 uM 02 (02). 96 3.5 Turning Basin 2.5 - +TAhighN --O-- TAlowN We TA02 +TBhighN -->(-- TBlowN --><-- T802 —O—TChighN --O-- TClowN -----0 TC02 +TDhighN --)K-- TDIowN ---)l<- T002 +TEhighN —-l-— TElowN ---l-- TE02 1.5 ~ Nitrous oxide (micromoles I microcosm) Figure 3.15. Nitrous oxide production in microcosms of Turning Basin (Puget Sound) sediments from various depths after addition of different electron acceptors. Sediment depth is indicated by letters A through E (A, 0-2 cm; B, 2-4 cm; C, 10-12 cm; D, 34-38 cm; E, 69 cm). Different microcosm sets were amended with 400 M NaNO3 (highN), 4O uM NaNO3 (lowN) or 160 M 02 (O2). 97 2.5 Washington Margin A 2 - E (D o O 2 .2 E m 1.5 1 2 o E B .2 E. 0 17 :2 x o 0) 3 .5 z 0.5 — >6 —'-‘-'—“‘" '-'—'.-"_-,-.—v‘-‘x / 9999999 ’i— 99999999999 1" ....... 07. o- - , —-—-:---—; ------ 0 10 20 Time (h) +WAhighN --O-- WAlowN MO WA02 ——X——WBhighN --><-— WBlowN --X-- W802 +WChighN —-O-- WClowN -----0 W002 +WDhighN --X-- WDlowN --X-- W002 —I—WEhighN —-I-- WElowN ---I- WE02 Figure 3.16. Nitrous oxide production in microcosms of Washington Margin sediments from various depths after addition of different electron acceptors. Sediment depth is indicated by letters A through E (A, 0-2 cm; B, 2-4 cm; C, 10-12 cm; D, 34-38 cm; E, 62 cm). Different microcosm sets were amended with 400 pM NaNO3 (highN), 40 uM NaNO3 (lowN) or 160 uM 02 (02). 98 Table 3.5. Nitrous oxide production rates in microcosms amended with 400 pM NaN03. Rates of N20 production Depth (cm) (pmoles N20 h'l microcosm'l) SB CI TB WC 0-2 0.129 0.0612 0.102 0.0509 2-4 0.133 0.0432 0.0471 0.0386 10-12 0.119 0.0528 0.077 0.0233 24-28 0.0915 ND ND ND 34-38 0.0126 0.0304 0.0768 0.0026 62 ND ND ND 0.0003 69 ND ND 0.0004 ND 135 ND 0.0001 ND ND ND: not determined. 99 Original sediment A 041.5 E62 1:172 E 1.1_5 E74 5179 f; 2-2.5 580 8384 E10-10.5 '89 391 Q g 24.25 094 097 ,3 999., 13102 I103 Amended 0105 I108 0110 0111 .: 0-2 a D116 0137 s 24 i .. A 0140 3:167 S 5 10-12 l g - )0169 0189 1: 4. g 2 28 1.221 I228 3“” 1,0275 I329 l0344 0376 :33 0" )0406 D416 § g 5" I461 I474 _‘.=’ "'5 26 h 0598 0% 20% 40% 60% 80% 100% Relative abundance of T-RFs (%) Figure 3.17. T-RF LP profiles obtained by amplification of nirS genes from Shallow Budd Inlet, Puget Sound, sediments compared to profiles obtained after incubation of sediments from the same site with 400 uM NaNO; in microcosms. Relative abundances of T-RFs in microcosms from different sediment depths incubated for 26 h, as well as from sediments of 2-4 cm depth incubated for different times, are represented. T-RFs were generated after restriction with Mspl. Numbers in legend indicate T-RF length in base pairs for fragments representing more than 1% relative abundance. 100 10 8 9 36-37. cm 6 - 0 Original sediment § 4 — 34-38 cm 5 A El Amended with NaN03 N 2 ~ o n 2499 cm (time curve) E 5 h DD A a 0 ‘ 0_2 cm 2-4 cm /26h 24-25 cm 0'0-5 cm A Amended with NaN03 _2 _ 91042 cm 1.1 .5 0m (different sediment ’2-25 cm depths) _4 _ 10-105 cm '6 l l r -10 -5 0 5 10 Dim 1 (59%) Figure 3.18. PCA ordination of T-RFLPs of nirS genes from Shallow Budd Inlet, Puget Sound, sediments not subjected and subjected to incubation with 400 [1M NaNO3 in microcosms. In addition to the untreated sediments from several depths, sediments from 2 to 4 cm depth incubated for different times, as well as sediments from different depths incubated for 26 h, were included in the analysis. Labels next to the points in the graph indicate sediment depth or incubation time. 101 Discussion Bioturbation of the sediments by the activity of benthic macrofauna has three major implications: modification of sediment texture, bio-irrigation, and dispersal of solid particles (Meysman et al., 2006), notably first recognized by Charles Darwin who observed the reworking activity of earthworms in soil (Darwin, C., 1881). The dispersal of solid particles, which include not only non-living particles, but also some of biological nature, e.g. microbes, was suggested as an explanation for the conserved denitrifier community structure observed in sediments regardless of the redox gradients (Braker er al., 2001). I further tested this hypothesis, by analyzing sediments lying below the bioturbated zone and compared sites with significantly different bioturbation intensities. As expected, a conservation of the denitrifier community structure, revealed by T-RFLP of the nirS gene, was detected within the mixed sediments from all areas, with a sharp decrease in richness of 3- to 8-fold in deep unmixed sediments (Figure 3.4). This decrease in richness is not related, however, only to the disappearance of T-RFs in deep samples, but also to a replacement by others, as some particular T-RFs are detected only in deep samples, thus indicating the presence of microbes better adapted to the conditions of these subsurface sediments. Furthermore, the predation by macrobenthic fauna on bacteria in the bioturbated sediments, an effect absent in the deep unbioturbated sediments, might be an additional factor leading to the differentiation of the communities in these two depth zones, as well as the mucus secreted by these animals when gliding, which constitutes a source of organic carbon and has been demonstrated to stimulate microbial growth (Reise, 2002). The detected shifi in community structure is consistent with the observation of clear differences in the bacterial types isolated from‘surface 102 Wadden Sea sediments when compared to those from 50 and 100 cm deep layers, and almost no overlap with those from > 200 cm deep sediments (Kepke et al., 2005), a trend also observed by DGGE (Wilms et al., 2006). Furthermore, microarray analysis of various functional genes, including nirS, in Puget Sound sediments also showed a conserved community structure down to 25 cm, significantly different from the one observed below 50 cm depth (Tiquia et al., 2006). The highly conserved community structure down to 38 and 10 cm in Puget Sound and Washington Margin sediments, respectively, was directly correlated to the presence of macrobenthic fauna down to at least those depths, but absent in the deeper analyzed samples. Burrows of benthic infauna have also been observed in previous studies extending down to at least 35 cm in natural sediments and microcosms (Satoh et al., 2007) and can even reach depths of around 1 m for some polychaetes (Reise, 2002). These burrows not only lead to the physical mixing of the sediments, but also to the irrigation of the sediments with overlying water. Microelectrode measurements in a sediment microcosm allowed the detection of oxygen down to a depth of 35 cm within a burrow (Satoh et al., 2007). This can therefore. explain the presence of denitrifiers as deep as 38 cm in Puget Sound sediments, and furthermore, justify their ability to start denitrifying without lag time when incubated with nitrate amendment (Figures 3.13, 3.14, 3.15, and 3.16), as nitrate might have penetrated deep into the sediments through burrows, leading to concentrations in the surrounding sediments similar to the surface sediments, differentiating them from the nitrate-depleted bulk sediments. PCA ordination as well as cluster analysis grouped the mixed sediments according to site and separate from the deep sediments (Figures 3.5 and 3.6). Braker et al. had also 103 observed site-specific grouping of nirS and 16S genes from sediments retrieved from the same area (Braker et al., 2001). Furthermore, the most geographically distant and biogeochemically divergent site (Washington Margin) was also the most distantly related in PCA and cluster analysis, while the two closest locations in Puget Sound with similar nitrate profiles (Turning Basin and Shallow Budd Inlet) were grouped closer together. This pattern was also observed by T-RF LP analysis of another denitrification gene (nosZ) in continental shelf sediments from the Atlantic Ocean, with increasing variability when going from a centimeter to a meter and further to a kilometer scale (Scala and Kerkhof, 2000) Although nirS clones retrieved from Shallow Budd Inlet presented a sequence divergence of up to 49.6% being also only distantly related to cultured denitrifiers, their similarity to other environmental clones never fell below 75% with a significant fraction (65%) being more than 90% similar to these previously retrieved nirS sequences (Figures 3.7, 3.8, and 3.9). This indicates an ever increasing database of environmental nirS sequences of different environments, leading to a higher coverage of the nirS diversity in environmental samples. However, these environmental sequences are still, with a few exceptions, only distantly related to cultured denitrifiers, which leads one to question their functionality in the environment. The close similarity of several Puget Sound clones to an mRNA nirS sequence (clone ANIS-54) from the River Colne estuary (up to 97.1% similarity) (Nogales er al., 2002) might be an indication of the expression of these sequences in nature, giving some support to the idea that they code for functional nitrite reductases. 104 A nirS clone (pA33) previously detected in sediments at a different location in Puget Sound (Braker er al., 2000) was, not surprisingly, the closest environmental nirS sequence to a clone retrieved in this study (NIRS4B F09). Although quite closely related (98.5% similarity), pA33 as well as four additional previously retrieved Puget Sound clones (pA16, pA63, p816, and pB20) (Braker et al., 2000) were not detected by real- time PCR quantification with specific primers designed against them at the four sites studied here (data not shown), indicating the absence of more closely related sequences that could be detected by the highly specific real-time PCR method. Therefore, a more site-specific occurrence of extremely similar sequences is inferred. Although most clones were only distantly related to cultured strains, four clones (cluster 4) were up to 83.3% similar to Raseabacter denitrificans. The Raseabacter clade is ofien found associated with dinoflagellates, as it has the ability to degrade dimethylsulfonopropionate (DMSP) produced by the latter (Miller and Belas, 2004). Interestingly, red tide was occurring at the time of sampling of Shallow Budd Inlet sediments, leading to the speculation that cluster 4 nirS sequences might be associated with Raseabacter clade members able to catabolize DMSP. Furthermore, nirS and 16S phylogenetic trees, including also clones from other areas exhibited the same association of Raseabacter denitrificans with clones retrieved from Shallow Budd Inlet (Chapter 4). The commonly isolated marine denitrifier P. stutzeri was detected in all studied areas by real-time PCR, being most abundant in Washington Margin sediments (0.66% P. stutzeri DNA in total community DNA), compared to 1 and 2 orders of magnitude lower abundance in Puget Sound sediments (Figures 3.1] and 3.12). Its abundance in Washington Margin sediments was 10-fold and several orders of magnitude higher than 105 previously determined abundances in Monterey Bay water column (Ward and Cockcrofi, 1993) and sediments from other sites in Puget Sound and the coast of Washington (Griintzig et al., 2001), respectively. This suggests that, though apparently cosmopolitan, this bacterium presents a rather wide range of abundances even in geographically close locations. On the other hand, quantification of clusters of novel nirS sequences by real-time PCR showed significantly higher abundances in sediments from which these sequences were originally retrieved (Shallow Budd Inlet) and the closely located Turning Basin than in Carr Inlet and Washington Margin sediments, giving further support to the above stated idea of site-specific nirS sequences being present at these locations. Each cluster of novel nirS sequences presented generally a similar abundance as measured by real-time PCR throughout the mixed sediments, with a sharp decrease in unmixed sediments, indicating that their relative abundance in the bacterial community, and not only in the denitrifier community as observed by T-RF LP, does not change with depth within the mixed zone. Interestingly, P. stutzeri nirS abundance determined by real-time PCR did not diminish in deep samples from Washington Margin and Turning Basin lying below the bioturbation zone, therefore, its relative abundance in the bacterial community seems stable throughout the studied sediment depths, regardless of mixing. As several other denitrifiers, represented for example by the novel nirS clusters, seem to diminish in deep unmixed sediments, the fraction of nirS DNA corresponding to P. stutzeri in the total nirS DNA assemblage will increase. This is consistent with the increase of T-RF 137 bp, which correlates with the expected size for P. stutzeri-nirS, in deep Washington Margin samples (Figure 3.4). Therefore, P. stutzeri seems to be able to 106 survive and possibly also to thrive in sediments depleted of nitrate and oxygen for a long period of time. Though not known to be fermenters, a very low level of fermentation has been detected in Pseudomonas isolates depleted of nitrate and oxygen for several months (Jorgensen and Tiedje, 1993) and could account for the maintenance of these denitrifiers in the absence of their normally used electron acceptors. Furthermore, iron reduction has been recently detected in certain Pseudomonas strains (Naganuma er al., 2006), constituting another alternative metabolicpathway for these denitrifiers. In addition to the conserved denitrifier community structure in mixed sediments based on semi-quantitative (T-RF LP) and quantitative (real-time PCR) distribution of nirS sequences, the denitrification capacity of the sediments throughout the mixed layer was confirmed by nitrous oxide accumulation after nitrate addition (Figures 3.13, 3.14, 3.15, and 3.16). No lag time was observed in all cases, suggesting that the denitrification enzymes might have already been expressed in situ, even in sediments below the depth in which nitrate is detected in bulk sediments. As suggested above, nitrate might have penetrated deep into the sediments through macrobenthic fauna] burrows, leading to denitrification activity in the sediments surrounding them. Satoh et al., observed a higher number of 16S rRNA gene copy numbers of ammonia-oxidizing bacteria and Nitraspira- like nitrite-oxidizing bacteria in sediments surrounding infaunal burrows thaniin bulk sediments (Satoh et al., 2007). Furthermore, ammonium consumption was also higher at the burrow wall, indicating that these sites were hot-spots for nitrification, due to the penetration of oxygen through the burrows, although no oxygen was detected in the surrounding bulk sediments. Therefore, besides the effect of burrows allowing the direct penetration of nitrate into deeper sediments, the indirect effect of stimulation of 107 nitrification through the penetration of oxygen, has also been shown to increase the denitrification rates, through coupled nitrification-denitrification (Dollhopf er al., 2005). Although the presence of an N-oxide, as well as anaerobiosis, is normally required for denitrification gene expression (Zumft, 1997), nirS gene expression has also been observed in the absence of N-oxides in P. aeruginosa PAOl and P. stutzeri JM300 (Ka er al., 1997). Therefore, some denitrification gene expression could also occur in nitrate- depleted bulk sediments, maybe also related to the periodic exposure to nitrate, and explain the fast response of denitrifiers to nitrate amendment. No denitrification capacity was, however, detected in sediments below the bioturbated zone, indicating that either no denitrifiers with all necessary enzymes to convert N03' to N20 are present in these sediments, or that longer incubation times are required before detecting N20 accumulation. Stoichiometric conversion to N20 of 40 M NaNO; was only observed for Carr Inlet top sediments, with less than 45% recovery of the added N-NO3’ as N-N20 in the other samples. This might be due to the conversion of nitrate to ammonium by DNRA, which could be important in these sediments as this process is favored under high available carbon to electron acceptor ratios (Tiedje, 1994). On the other hand, technical problems during the determination of denitrification capacity, such as the imperfect inhibition of the nitrous oxide reductase by acetylene in reduced sediment layers (Binnerup er al. , 1992), could have led to underestimation of denitrification. Surprisingly, the setup of microcosms for the determination of denitrification capacity led to a noteable change in the denitrifier community structure, detectable already before addition of nitrate to the sediments in the microcosm incubated for 0 h 108 (Figure 3.17). Natural and microcosm sediments were clearly separated by PCA ordination, while sediment depth and incubation time did not exert such a significant effect on the community structure (Figure 3.18). Although spatial variations between subcores used for original determination of denitrifier community structure by T-RF LP and microcosm studies can not be ruled out, it seems improbable that the significant changes detected are a reflection of the cm scale separation of the subcores, as previous studies have observed only minor variations in community structure at this scale (Scala and Kerkhof, 2000; Berardesco er al., 1998). Therefore, it seems that the manipulation of the sediments used for microcosm setup in addition to the time frame between retrieval and processing of the cores was responsible for the observed changes. Particularly, keeping the cores at 4°C for 3 days, 15°C below their in situ temperature, might have had a significant effect, as seasonal variability of the denitrifier community has been determined to be significant in sediments (Scala and Kerkhof, 2000), which might be linked to changes in temperature, besides other factors. Furthermore, the disturbances exerted on the sediments during microcosm setup, such as the dilution with artificial seawater and mixing, might have also affected the bacterial community. Although the surface of the bulk sediments was briefly exposed to oxygen during the transfer to the microcosms, the large bulk of reduced sediments was never exposed to it. Therefore, the effect of oxygen on the community can be considered minimal. Interestingly, the abundance of the T-RF representing most of the novel nirS genes retrieved from this environment (97 bp) was lower in the microcosm sediments than in the original sediments. On the other hand, the abundance of the T-RF of 137 bp, coincident with P. stutzeri, and the T-RF of 169 bp, coincident with twelve of the clones, Raseabacter 109 denitrificans, Azaarcus talulyticus, and a marine denitrifier (C10-1) isolated from sediments at the Washington margin (Braker er al., 2000) showed the opposite trend, being higher in the microcosm sediments. Therefore, manipulation of sediments early on after retrieval might have already affected community structure favoring commonly isolated bacteria, such as P. stutzeri (Ward and Cockcrofi, 1993). In conclusion, denitrifier populations, using the nirS gene as marker, were distributed homogeneously, regardless of the redox potential, within the bioturbated sediments, with a sharp change in their community structure in deep unmixed sediments. Denitrification capacity, indicative of potentially active denitrifiers, was, as well, detected throughout the mixed layer, highlighting the importance of considering them when studying denitrification and its main players in the environment. 110 References Acinas S. G., R. Sarma-Rupavtarm, V. Klepac-Ceraj, and M. F. P012. 2005. PCR- induced sequence artifacts and bias: insights from comparison of two 16S rRNA clone libraries constructed from the same sample. Appl. Environ. Microbiol. 71:8966-8969. Berardesco, G., S. Dyhrman, E. Gallagher, and M. P. Shiaris. 1998. Spatial and temporal variation of phenanthrene-degrading bacteria in intertidal sediments. Appl. Environ. Microbiol. 64:2560-2565. Binnerup, S. J., K. Jensen, N. P. Revsbech, M. H. Jensen, and J. Sarensen. 1992. Denitrification, dissimilatory reduction of nitrate to ammonium, and nitrification in a bioturbated estuarine sediment as measured with ”N and microsensor techniques. Appl. Environ. Microbiol. 582303-313. Blackwood, C. B, T. Marsh, S.-H. Kim, and E. A. Paul. 2003. Terminal restriction fragment length polymorphism data analysis for quantitative comparison of microbial communities. Appl. Environ. Microbiol. 69:926-932. Braker, G., H. A. Ayala-del-Rio, A. H. Devol, A. Fesefeldt, and J. M. Tiedje. 2001. Community structure of denitrifiers, Bacteria, and Archaea along redox gradients in Pacific Northwest marine sediments by terminal restriction fragment length polymorphism analysis of amplified nitrite reductase (nirS) and 16S rRNA genes. Appl. Environ. Microbiol. 67 : 1893-1901 . Braker, G., J. Zhou, L. Wu, A. H. Devol, J. M. Tiedje. 2000. Nitrite reductase genes (nirK and nirS) as functional markers to investigate diversity of denitrifying bacteria in Pacific Northwest marine sediment communities. Appl. Environ. Microbiol. 66: 2096— 2104. Braker, G., A. Fesefeldt, and K.-P. Witzel. 1998. Development of PCR primer systems for amplification of nitrite reductase genes (nirK and nirS) to detect denitrifying bacteria in environmental samples. Appl. Environ. Microbiol. 64:3769-3775. Braker, G., J. Zhou, L. Wu, A. H. Devol, and J. M. Tiedje. 2000. Nitrite reductase genes (nirK and nirS) as functional markers to investigate diversity of denitrifying bacteria in Pacific Northwest marine sediment communities. Appl. Environ. Microbiol. 66:2096-2104. Brandes, J. A., and A. H. Devol. 1995. Simultaneous nitrate and oxygen respiration in coastal sediments: Evidence for discrete diagenesis. Journal of Marine Research. 53:771- 797. Brewer, P. G. and D. W. Spencer. 1974. Colorimetric determination of manganese in anoxic waters. Limnology and Oceanography. 16:107-110. 111 Carpenter, R., and M. L. Peterson. 1989. Chemical cycling in Washington's coastal zone, p. 367-509. In M. R. Landry and B. M. Hickey (ed.), Coastal oceanography of Washington and Oregon. Elsevier, Amsterdam. Carpenter, R., M. L. Peterson, and J. T. Bennett. 1982. 2lon-derived sediment accumulation and mixing rates for the Washington continental slope. Mar. Geol. 48:135- 164. Carpenter, R., M. L. Peterson, and J. T. Bennett. 1985. 2”’Pb-derived sediment accumulation and mixing rates for the greater Puget Sound region. Mar. Geol. 64:291- 312. Castro-Gonzalez, M., G. Braker, L. Farias, and O. Ulloa. 2005. Communities .of nirS- type denitrifiers in the water column of the oxygen minimum zone in the eastern South Pacific. Environ. Microbiol. 7: 1298-1306. Darwin, C. 1881. The Formation of Vegetable Mould Through the Action of Worms With Observation of Their Habits. John Murray. Dollhopf, S. L., H. L. Ayala-del-Rio, and J. M. Tiedje. 2004. Artificial neural network and multivariate statistical analysis of microbial community fingerprints. In Kowalchuk, G. G., de Brujin, F. J., Head, 1. M., Akkermans, A. D., and van Elsas, J. D. (eds.). Molecular methods in microbial ecology manual. Second edition. Springer Verlag. Dollhopf, S. L., J.-H. Hyun, A. C. Smith, H. J. Adams, S. O’Brien, and J. E. Kostka. 2005. Quantification of ammonia-oxidizing bacteria and factors controlling nitrification in salt marsh sediments. Appl. Environ. Microbiol. 71:240-246. Dunbar, J., L. O. Ticknor, and C. R. Kuske. 2000. Assessment of microbial diversity in four southwestern United States soils by 16S rRNA gene terminal restriction fragment analysis. Appl. Environ. Microbiol. 66:2943-2950. Dunbar, J., L. O. Ticknor, and C. R. Kuske. 2001. Phylogenetic specificity and reproducibility and new method for analysis of terminal restriction fragment profiles of 16S rRNA genes from bacterial communities. Appl. Environ. Microbiol. 67 :190-197. Fossing, H. and B. B. Jorgensen. 1989. Measurement of bacterial sulfate-reduction in sediments: evaluation of single-step chromium reduction method. Biogeochemistry. 8:205-222. Griintzig, V., S. C. Nold, J. Zhou, and J. M. Tiedje. 2001. Pseudomonas stutzeri nitrite reductase gene abundance in environmental samples measured by real-time PCR. Appl. Environ. Microbiol. 67:760-768. Hall, T. A. 1999. A user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucl. Acids Symp. Ser. 41:95-98. ' . 112 Hannig, M., G. Braker, J. Dippner, and K. Jiirgens. 2006. Linking denitrifier community structure and prevalent biogeochemical parameters in the pelagial of the central Baltic Proper (Baltic Sea). FEMS Microbiol. Ecol. 57:260-271. Hedges, J. and J. Stern. 1984. Carbon and nitrogen determinations of carbonate containing solids. Limnology and Oceanography. 29:657-663. Jayakumar, D. A., C. A. Francis, S. W. A. Naqvi, and B. B. Ward. 2004. Diversity of nitrite reductase genes (nirS) in the denitrifying water coltunn of the coastal Arabian Sea. Aquat. Microb. Ecol. 34:69-78. Jargensen, K. S., and J. M. Tiedje. 1993. Survival of denitrifiers in nitrate-free, anaerobic environments. Appl. Environ. Microbiol. 59:3297-3305. Ka, J.-O, J. Urbance, R. W. Ye, T.-T. Ahn, and J. M. Tiedje. 1997. Diversity of oxygen and N-oxide regulation of nitrite reductases in denitrifying bacteria. FEMS Microbiol. Lett. 156:55-60. Kemp, P. F., B. F .Sherr, E. B. Sherr, J. J. Cole. 1993. Handbook of methods in aquatic microbial ecology. Lewis Publishers. Kiipke, B., R. Wilms, B. Engelen, H. Cypionka, and H. Sass. 2005. Microbial diversity in coastal subsurface sediments: a cultivation approach using various electron acceptors and substrate gradients. Appl. Environ. Microbiol. 71:7819-7830. Kristensen, E., A. H. Devol, and H. E. Hartnett. 1999. Organic matter diagenesis in sediments on the continental shelf and slope of the eastern tropical and temperate North Pacific. Cont. ShelfRes. 19:1331-1351. Kumar, S., K. Tamura, I. B. Jakobsen, and M. Nei. 2001. MEGA2: molecular evolutionary genetics analysis software. Bioinformatics. 17:1244-1245. Legendre, P. and E. D. Gallagher. 2001. Ecologically meaningful transformations for ordination of species data. Oecologia. 129:271-280. Liu, W.-T., T. L. Marsh, H. Cheng, and L. J. Forney. 1997. Characterization of microbial diversity by determining terminal restriction fragment length polymorphisms of genes encoding 16S rRNA. Appl. Environ. Microbiol. 63:4516-4522. Llobet-Brossa, E., R. Rosella-Mora, and R. Amann. 1998. Microbial community composition of Wadden Sea sediments as revealed by fluorescence in situ hybridization. Appl. Environ. Microbiol. 64:2691-2696. Ludwig, L., O. Strunk, R. Westram, L. Richter, H. Meier, Yadhukumar, A. Buchner, T. Lai, S. Steppi, G. Jobb, W. Ffirster, I. Brettske, S. Gerber, A. W. Ginhart, 0. Gross, S. Grumann, S. Hermann, R. Jost, A. Kiinig, T. Liss, R. LiiBmann, M. May, B. Nonhoff, B. Reichel, R. Strehlow, A. Stamatakis, N. 113 Stuckmann, A. Vilbig, M. Lenke, T. Ludwig, A. Bode and K.-H. Schleifer. 2004. ARB: a software environment for sequence data. Nucleic Acids Research. 32:1363-1371. Meysman, F. J. R., J. J. Middelburg, and C. H. R. Heip. 2006. Bioturbation: a fresh look at Darwin’s last idea. Trends in Ecology and Evolution. 21:688-695. Miller, T. R. and R. Belas. 2004. Dimethylsulfoniopropionate metabolism by Pfiesteria- associated Raseabacter spp. Appl. Environ. Microbiol. 70:3383-3391. Naganuma, T., M. Sato, D., Hoshii, Y. Amano-Murakami, T. Iwatsuki, and K. W. Mandernack. 2006. Isolation and characterization of Pseudomonas strains capable of Fe(III) reduction with reference to redox response regulator genes. Geomicrobiology Journal. 23:145-155. Nei, M. and S. Kumar. 2000. Molecular evolution and phylogenetics. Oxford University Press, New York. Nogales, B., K. N. Timmis, D. B. Nedwell and A. M. Osborn. 2002. Detection and diversity of expressed denitrification genes in estuarine sediments afier reverse transcription-PCR amplification from mRNA. Appl. Environ. Microbiol. 68:5017-5025. Qiu, X. Y., L.-Y. Wu, H. Huang, P. E. McDonel, A. V. Palumbo, J. M. Tiedje, and J.-Z. Zhou. 2001. Evaluation of PCR-generated chimeras, mutations, and heteroduplexes with 16S rRNA gene-based cloning. Appl. Environ. Microbiol. 67 : 880-887. Reise, K. 2002. Sediment mediated species interaction in coastal waters. J. Sea Res. 48:127-141. Sambrook, J., E. F. Fritsch, and T. Maniatis. 1989. Molecular cloning: a laboratory manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. Sambrook, J. and D. W. Russell. 2001. Molecular cloning: a laboratory manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N. Y. Saitou, N. and M. Nei. 1987. The neighbour joining method: a new tool for reconstructing phylogenetic trees. Mol. Biol. Evol. 4:406-425. Santoro, A. E., A. B. Boehm and C. A. Francis. 2006. Denitrifier community composition along a nitrate and salinity gradient in a coastal aquifer. Appl. Environ. Microbiol. 72: 2102-2109. Satoh, H., Y. Nakamura, and S. Okabe. 2007. Influences of infaunal burrows on the community structure and activity of ammonia-oxidizing bacteria in intertidal sediments. Appl. Environ. Microbiol. 73: 1 341-1348. Scala, D. J. and L. J. Kerkhof. 2000. Horizontal heterogeneity of denitrifying bacterial communities in marine sediments by terminal restriction fragment length polymorphism analysis. Appl. Environ. Microbiol. 66:1980-1986. 114 Serensen, J. 1978. Capacity for denitrification and reduction of nitrate to ammonia in a coastal marine sediment. Appl. Environ. Microbiol. 35:301-305. Stookey, L. L. 1970. Ferrozine —a new spectrophotometric reagent for iron. Analytical Chemistry. 42:779-781. Strickland, J. D. H. and T. R. Parsons. 1972. A practical handbook of seawater analysis. Fisheries Research Board of Canada, Ottawa, Ontario. Tabatabai, M. A. 1974. A rapid method for determination of sulfate in water samples. Environ. Letters 7 :237-243. ' Thompson, J. D., T. J. Gibson, F. Plewniak, F. Jeanmougin, and D. G. Higgins. 1997. The ClustalX windows interface: flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucleic Acids Research. 24:4876-4882. Tiedje, J. M. 1994. Denitrifiers. pp. 245-267. Methods of Soil Analysis, Part 2. Microbiological and Biochemical Properties. Soil Sci. Soc. Amer., Madison, Wisconsin. Tiedje, J. M., A. J. Sexstone, D. D. Myrold, and J. A. Robinson. 1982. Denitrification: ecological niches, competition and survival. Antonie van Leeuwenhoek. 48:569-583. Tiquia, S. M., S. Gurczynski, A. Zholi, and A. Devol. 2006. Diversity of biogeochemical cycling genes from Puget Sound sediments using DNA microarrays. Enviromnental Technology. 27:1377-1389. Tuomainen, J. M., S. Hietanen, J. Kuparinen, P. J. Martikainen, and K. Servomaa. 2003. Baltic Sea cyanobacterial bloom contains denitrification and nitrification genes, but has negligible denitrification activity. FEMS Microbiol. Ecol. 45:83-96. Urakawa, H., T. Yoshida, M. Nishimura, and K. Ohwada. 2000. Characterization of depth-related population variation in microbial communities of a coastal marine sediment using 168 rDNA-based approaches and quinone profiling. Environ. Microbiol. 2:542- 554. Ward, J. H. 1963. Hierarchical grouping to optimize an objective function. Journal of the American Statistical Association. 58:236-244. Ward, B. B. and A. R. Cockcroft. 1993. Immunofluorescence detection of the denitrifying strain Pseudomonas stutzeri (ATCC 14405) in seawater and intertidal sediment environments. Microb. Ecol. 25:233-246. Wilms, R., H. Sass, B. Kiipke, J. Kiister, H. Cypionka, and B. Engelen. 2006. Specific bacterial, archaeal, and eukaryotic communities in tidal-flat sediments along avertical profile of several meters. Appl. Environ. Microbiol. 72:2756-2764. Zumft, W. G. 1997. Cell biology and molecular basis of denitrification. Microbiol. Mol. Biol. Rev. 61:533-616. 115 CHAPTER 4 DENITRIFIER AND BACTERIAL DIVERSITY AND DISTRIBUTION PATTERNS IN SEDIMENTS FROM THE ARCTIC AND THE PACIFIC NORTHWEST Abstract Denitrifier and bacterial communities were studied in sediments from nine geographically distant sites with different geochemical characteristics, including the Arctic (high and low productivity areas) and the Pacific Northwest (Puget Sound, continental slope and abyssal sea floor). High diversity and richness were observed in all sites for the denitrifying as well as the bacterial community, based on heme cal) nitrite reductase (nirS) and 16S rRNA gene clone libraries, respectively, with the highest values measured at the Pacific Northwest continental slope. The different sites seem to harbor distinct microbial communities, as indicated by the statistically significant differences observed in the clone libraries from different sites, with geographic distance as the main determinant in differentiation of denitrifier, but not bacterial communities. This was further confirmed by station-specific clustering on phylogenetic trees and by low similarity and estimated richness of shared OTUS, often close to zero, indicating little or no overlap in membership between communities at different sites. Station-specific clustering of different depth denitrifier communities was also observed based on nirS T- RFLP profiles, with clusters from close geographic locations combining, in turn, into larger clusters. These observations support the hypotheses that differences in denitrifier community structure are influenced more strongly by geographic location than by 116 differing environmental factors within one area, often related to water depth, which in turn have a stronger influence than sediment depth. Introduction In recent years, interest has grown regarding the study of the distribution of bacteria in the environment. Marine sediments have been shown to contain a particularly high abundance and diversity of microorganisms (Whitman er al., 1998; Torsvik et al., 2002), which is only starting to be discovered. F urthermore, the distribution of microorganisms seems to follow specific laws, similar to the ones governing the distribution of macroorganisms, leading to the frequent observation of biogeographic patterns (Green and Bohannan, 2006; Hughes Martiny er al., 2006; Ramette and Tiedje, 2007). Although not necessarily applicable to all microorganisms, some endemism has been observed when the distribution of certain taxonomic groups or of bacteria in general has been studied, often depending on the level of taxonomic resolution chosen (Cho and Tiedje, 2000; Ramette and Tiedje, 2007; Homer-Devine er al., 2004). The distribution of denitrifying bacteria, a functional group that can not be defined by a corresponding taxonomic group, also showed specific patterns in the environment, when the diversity of denitrification genes was studied. So, horizontal heterogeneity at different distance scales in marine sediments and along environmental gradients in a beach aquifer was observed for nitrous oxide reductase (nosZ) (Scala and Kerkhof, 2000; Scala and Kerkhof, 1999) and nitrite reductase (nirS and nirK) (Santoro et al., 2006) genes, respectively. Furthermore, Pacific Northwest marine sediment samples exhibited differential nirS-based community structures, with geographic location, water depth and 117 sediment depth exerting decreasing influence in the differentiation of the communities, in that order (Braker et al, 2000; Braker er al., 2001). Our previous work (Chapter 3) with sediments from the same area also suggested this pattern, with the exception that sediment depth played a more significant role when communities above and below the bioturbation depth were compared. In order to further describe the distribution of denitrifiers and the influence the above stated factors have on their differentiation, a broad range of sediment samples from various locations and biogeochemical characteristics was analyzed. The Arctic Ocean comprises 25% of the total continental shelf areas, the main sites for nitrogen cycling taking part in the oceans (Walsh, 1991), and climate models suggest that the effects of global warming will be magnified in this area (Houghton et al., 2001). Sediment denitrification rates in some highly productive regions on the Arctic margins reach values similar to the highly productive Pacific Northwest area, while they are lower in less productive regions (Devol er al., 2005). Therefore, this region constitutes an interesting extreme environment to study the denitrifier community structure and compare it to the geographically distant and biogeochemically different, yet similar in certain aspects, Pacific Northwest area. Deep-sea sediments are still poorly explored, however, a high bacterial diversity, including denitrifiers, is being unraveled in this environment (Tamegai et al., 2007) and functional denitrifiers have been isolated even from 11,000 m deep sediments at the Mariana Trench (Tamegai er al, 1997). This makes them interesting targets for denitrifier community structure studies and comparison to continental shelf and slope sediments. Therefore, in addition to the highly productive previously studied Pacific Northwest 118 sediments from Puget Sound and Washington Margin (Chapter 3), abyssal sea floor sediments from the same area, however located west of the Juan de Fuca ridge, a significant physical barrier for dispersion of sediment organisms, as well as Arctic sediment samples from different depths within high and low productivity zones were analyzed and their denitrifier and bacterial community structures determined and compared between different sites, based on nirS and 16S rRNA genes, respectively. Materials and Methods Study area and sediment sampling Marine sediment samples were collected from five stations in the Pacific Northwest and four in the Arctic (Figure 4.1, Table 4.1). The Pacific Northwest stations included three in Puget Sound (Shallow Budd Inlet, Turning Basin, and Carr Inlet), one on the continental slope (Washington Margin) and one on the abyssal sea floor west of the Juan de Fuca ridge (West of Juan de Fuca) in the Pacific Ocean. The Arctic stations included a shallow and a deep sample from transects along Barrow Canyon and east of Hanna Shoal in the Chukchi Sea (Barrow Canyon Shallow, Barrow Canyon Deep, East Hanna Shoal Shallow, East Hanna Shoal Deep). Puget Sound and Washington Margin samples were retrieved and stored as described previously (Chapter 3). The West of Juan de Fuca sample was collected aboard the R/V T.G. Thompson in July 2001. Barrow Canyon and East Hanna Shoal samples were retrieved during a cruise on the US Coast Guard Ice Breaker USCGC Healy in July 2002 and May-June 2004, respectively. Sediment cores were obtained with a multicorer and sectioned into 0.5 to 2 cm slices. Sediments from the desired depths (Table 4.2) were kept at -20°C or centrifuged at 12,500 X g to separate the 119 bulk sediments from pore water and stored at -80°C, for West of Juan de Fuca and Arctic samples, respectively. Chemical analyses of sediment samples Chemical determinations for Puget Sound and Washington Margin samples were performed as described previously (Chapter 3). In the West of Juan de Fuca and Arctic sediments, a microelectrode was used for oxygen concentration determinations. The porewater separated after centrifugation of the sediments at 12,500 X g was used for nitrate (Armstrong et al., 1967), ammonitun (Strickland and Parsons, 1972), sulfate (Tabatabai, 1974), iron (Stookey, 1970) and manganese (Brewer and Spencer, 1974) determinations. The sulfate reduction rate in the sediments was measured by the method of Fossing and Jargensen (Fossing and Jorgensen, 1989). The percentage of organic carbon was obtained from other studies in the sampled area (Devol, pers. comm.) and was determined by the method of Hedges and Stern (Hedges and Stern, 1984). DNA extraction and purification Genomic and plasmid DNA extractions, purifications, and quality determinations were performed as described previously (Chapter 3). T-RFLP analysis of nirS gene Terminal restriction fragment length polymorphism (T-RFLP) analysis of the nirS gene was performed on sediment samples from selected depths from all sites (Tables 3.2 and 4.2), following the same procedure described in Chapter 3. Terminal reStriction fragments (T-RFs) generated from West of Juan de Fuca and Arctic samples were analyzed by capillary electrophoresis. 120 PCA ordination and cluster analysis of the T-RF LP profiles, as well as comparison to the in silica digestion of nirS clones retrieved in this study was performed as described previously (Chapter 3), with the exception that the Sarensen distance, which considers only presence and absence of T—RFs, was applied in the cluster analysis. Canonical correspondence analysis (CCA) (ter Braak, 1986) was applied on the T-RFLP results to obtain an ordination of the samples (sites) based on their biological components (T-RFs) by correspondence analysis, but optimized in terms of the influence of environmental variables. One T-RFLP profile from each site was used for the analysis, corresponding to a sediment depth of 1-1.5, 1-2, or 0-2 cm, depending on the sample depths available from each site (Tables 3.2 and 4.2). Nine environmental variables measured in all areas were included in the analysis (oxygen, nitrate, ammonium, water depth, latitude, longitude, temperature, salinity and organic carbon) (Table 4.1; Figures 4.2, 4.3, and 4.4). Overlying water values were used for oxygen, nitrate and ammonium. CCA was performed using the Environmental Community Analysis software version 1.37 (Pisces Conservation Ltd., Hampshire, United Kingdom). In addition, to describe the nirS-containing community at each site and depth, richness and diversity were determined based on the T-RFLP profiles. The total number of T-RFs for each sample was presented as its richness, while the diversity was calculated with the Shannon index H ’(Hill et al., 2003): H ’=-Zp,- ln 1),, where p,- is the proportion of the ith T-RF in the community. An in silica digestion of the nirS clone sequences retrieved in this study was performed for Hhal and Mspl with a custom made Perl script. The theoretical T-RFLP 121 profile produced by the clones was compared to the actual T-RFLP profiles. The Sarensen similarity index was calculated for hypothetical and real T-RFLP profiles from the same sediment depths. Cloning of nirS and 16S rRNA gene sequences Clone libraries of nirS and eubacterial 16S rRNA gene sequences were constructed from Shallow Budd Inlet, Washington Margin, West of Juan de Fuca, Barrow Canyon Shallow, and East Hanna Shoal Shallow sediments from one depth at each site.(Tables 3.2 and 4.2). The depth chosen in each case was related to a high richness in nirS sequences based on T-RF LP. The nirS clone libraries were constructed as described in Chapter 3. The eubacterial 16S rRNA clone libraries were constructed after PCR amplification of environmental DNA extracted from the sediments. Primers 8-27F (5’- AGAGTTTGATCMTGGCTCAG-3’ with M for A or C) and 1392-1407R (5’- ACGGGCGGTGTGTACA-3’) (Braker er al., 2001) were used in three replicate PCR reactions, containing 100 ng environmental DNA, 0.2 M of each primer, 200 M of each deoxyribonucleoside triphosphate, 3 mM MgC12, 0.2 pg pl"l bovine serum albumin (Roche Molecular Biochemicals, Indianapolis, IN), 2.5 U T aq polymerase (Promega, Madison, WI), and 1X buffer provided with the enzyme. After an initial denaturation step of 3 min at 95°C, 30 cycles of 45 sec at 94°C, 1 min at 57°C, and 2 min at 72°C were carried out, followed by a final extension of 7 min at 72°C. The replicate PCR reactions were pooled, concentrated to approximately 70 pl with a Speedvac (Savant, Holbrook, NY) and the amplified 1400 bp 16S rRNA fragments were purified from a 1.2% preparatory agarose gel with the QIAquick Gel Extraction kit (QIAGEN, Valencia, CA) following the manufacturer’s instructions. The purified fragments were cloned using the 122 TA cloning kit (Invitrogen, Carlsbad, CA) and 188 white insert-bearing clones were randomly selected for sequencing. These clones were grown on LB freezing buffer (Sambrook and Russell, 2001) with 50 pg ml'l kanamycin and inserts were sequenced with vector primer M13F using an ABI 1730 Genetic Analyzer or an ABI Prism 3700 DNA Analyzer (Applied Biosystems, Foster City, CA). Phylogenetic analysis of cloned nirS and 168 rRNA gene sequences The nirS and 168 rRNA sequences included in the clone libraries constructed from each site were compared to the GenBank database by using BLAST to remove sequences with no homology and select sequences from cultured strains and closely related environmental clones from other studies. Alignment and translation of the nirS sequences, followed by the construction of a neighbour-joining tree (Saitou and Nei, 1987) with Mega 2.1 (Kumar et al., 2001) using Poisson correction (Nei and Kumar, 2000) and 100 bootstraps was performed as described in Chapter 3. The 16S rRNA sequences retrieved in this study were checked for chimeras with the Chimera_Check program from the Ribosomal Database Project II (http://rdp.cme.msu.edu). The cloned sequences, in addition to the sequences selected from the GenBank database, were then aligned against the RDP public alignment using RNACAD in myRDP (Cole er al., 2006). The aligmnent was manually corrected in BioEdit (Hall, 1999) and phylogenetic trees were inferred by the neighbour-joining method (Saitou and Nei, 1987) with Mega 2.1 (Kumar et al., 2001) using Jukes-Cantor correction (Jukes and Cantor, 1969) and 100 bootstraps. The 16S rRNA sequences from individual libraries were assigned to taxonomic groups by placing them into the RDP Hierarchy with the RDP classifier (Cole er al., 123 2005). Significant differences in the taxonomic composition between the different libraries was further determined with the RDP Library Compare tool (Cole et al., 2006). Community description and comparison based on 16S rRNA and nirS clone libraries Sequences included in the 16S rRNA and nirS clone libraries from each site were assigned to individual operational taxonomic units (OTUS) with the purpose of community analysis. A Jukes-Cantor corrected distance matrix (Jukes and Cantor, 1969) generated with the DNADIST program from PHYLIP (http://evolution.genetics.washington.edu/phylip.html) based on the DNA alignments was used for the OTU assignment by DOTUR (Schloss er al., 2005). The furthest neighbor sequence assignment method was applied, with a 3% and 5% distance threshold for assignment of 16S rRNA and nirS sequences to the same OTU, respectively. The grouping of the sequences into OTUS allowed the evaluation of the sampling effort by the construction of rarefaction curves, the determination of diversity with the Shannon Index, and the estimation of richness with the bias-corrected Chaol nonparametric richness estimator, also using DOTUR. The statistical significance of the differences between libraries constructed from sediments retrieved at different sites was determined with l-LIBSHUFF (Schloss et al., 2004), which uses the integral form of the Cramér-von Mises statistic. This statistic compares the homologous and heterologous coverage curves between two libraries. A Jukes-Cantor corrected distance matrix constructed as above, but including the sequences retrieved from all sites was used for the calculations. To determine the probability that the differences observed between the original libraries were due to chance, a Monte Carlo 124 procedure was further applied, by the construction of randomized libraries (10,000) and comparing the value of the Cramér-von Mises statistic obtained in these random‘izations to the original value. Significant P values (P < 0.0026; 0 = 0.05) were determined after applying the correction for multiple comparisons, considering that five libraries were compared in 20 tests. To better describe the similarity between the communities at different sites, the richness of shared OTUS between each pair of communities as well as the community overlap between different sites was determined with a nonparametric richness estimator of shared OTUS analogous to Chaol and with the abundance-based Sorenson similarity index (Labmd), respectively, using SONS (Schloss and Handelsman, 2006). The preliminary assignment of sequences to OTUs was performed applying DOTUR on a Jukes-Cantor corrected distance matrix constructed as above, including the sequences retrieved from all sites. Also as above, sequences were assigned to OTUs with the furthest neighbor sequence assignment method, with a 3% and 5% distance threshold for 16S rRNA and nirS sequences, respectively. Quantification of nirS from specific clusters and strains by real-time PCR Five clusters of highly related nirS sequences identified previously in the Shallow Budd inlet clone library (Chapter 3), as well as the nirS gene from P. stutzeri were quantified in the studied sediments by real-time PCR. One depth from each Arctic site and from West of Juan de Fuca (Table 4.2) was studied as described previously (Chapter 3). In addition, the results obtained at four sediment depths (1-1.5 cm, 1.5-2 cm, 2-2.5 cm, and 10-10.5 cm) in Puget Sound and Washington Margin (Chapter 3) were considered biological replicates for the determination of statistically significant 125 differences of cluster and P. stutzeri abundances between locations or vice versa by one- factor analysis of variance (ANOVA) with Tukey’s HSD procedure as past hac test. No two-way ANOVA was performed, as a different number of groups were included in each comparison, as several samples had values below detection limit. Results Biogeochemical characteristics of the sediments Puget Sound and Washington Margin sediment characteristics have been described previously (Chapter 3). All Arctic sites showed high oxygen concentrations, reaching 320 M at Barrow Canyon Shallow. This highly productive site in the Arctic (Bates et al., 2005; Hill and Cota, 2005) presented the same phenomenon observed in Puget Sound and Washington Margin, namely oxygen and nitrate exhibiting narrow penetration and being depleted at similar depths (Figure 4.2), as generally observed in highly respiring near shore and continental shelf and slope sediments (Brandes and Devol, 1995). Barrow Canyon Deep, as expected for deeper sediments, exhibited a slightly deeper oxygen penetration of 8.5 mm. Nitrate was detected throughout the sediment core, although its concentration was < 1 nM below 7.5 cm depth. The ammonium profile at this site (Figure 4.3) indicates a sink at about 30 cm depth, maybe indicative of anammox occurring within these sediments. East Hanna Shoal sites present a higher oxygen penetration depth compared to Barrow Canyon of 1.5 and 3.8 cm for the Shallow and Deep sites, respectively. This is consistent with the lower productivity determined in this area in comparison to Barrow Canyon (Bates et al., 2005; Hill and Cota, 2005). Nitrate presented a similar trend to the observed in Barrow Canyon Deep, namely being present throughout 126 the sediment core, but its concentration dropping significantly below 1.25 and 3.5 cm depth for East Hanna Shoal Shallow and Deep, respectively. The West of Juan de Fuca site presented typical profiles for abyssal sea floor samples with very low organic carbon input (Table 4.1). At this site oxygen penetrated down to 5.1 cm into the sediments, while nitrate reached its maximum concentration at 1.2 cm depth (60.17 uM) and slowly declined to a concentration of 15.8 M at the deepest analyzed sediment sample (30 cm). Denitrifier community structure by T-RFLP of nirS genes The denitrifier community structure at various sediment depths at each site was studied by analyzing T-RFs generated after restriction of heme cd) nitrite reductase genes (nirS). An extremely low number of T-RFs were detected after restriction of West of Juan de Fuca samples with Mspl, therefore, Hhal was chosen for the subsequent data presentation. However, analysis performed with Mspl led to similar results and further supported the conclusions. As previously observed in Puget Sound and Washington Margin (Chapter 3), T- RFLP patterns in West of Juan de Fuca and Arctic sediments were highly' similar throughout the studied sediment depths, with a notable shift in community structure in the deep West of Juan de Fuca sample (20-21 cm) and a minor change detected in sediments below 5 cm depthiat the Barrow Canyon Shallow and East Hanna Shoal Deep sites (Figure 4.5). No correlation was observed between community structure and redox gradients. From 130 identified T-RFs, only seven (64, 108, 232, 236, 348, 388, and 436 bp), were common to all sites, although none was detected at all depths. These were generally present in significant relative abundance, reaching values as high as 64.4% for T-RF 64 bp in Carr Inlet sediments. The majority of T-RFs were absent in at least one 127 site, with 31% of the total number of T-RFs being characteristic for one specific sampling location. The highest richness, as determined by total number of individual T-RFs, was detected in Barrow Canyon Shallow sediments (59 T-RFs) (Figure 4.6A). The other Arctic sites, as well as Washington Margin sediments, also contain a rich denitrifier community, while the lowest richness was detected in Puget Sound sediments, with only one T-RF identified in the 69 cm deep sample at Turning Basin. The West of Juan de Fuca sample was the most diverse, as determined by the Shannon diversity index (H ’ = 3.12 at 4-5 cm depth) (Figure 4.6B), reflecting the high evenness in T-RF distribution detected at this site. On the other hand, the dominance of T-RF 64 bp in Carr Inlet sediments combined with a low richness, led to a low Shannon diversity index throughout the sediments at this site (1.44 < H’ < 2.07). Arctic and Washington Margin sediments presented a relatively high diversity, coincident with the high richness detected in these areas. Cluster analysis of the T-RFLP profiles (Figure 4.7) grouped the samples according to geographic location, with the exception of deep Turning Basin, Carr Inlet, and West of Juan de F uca samples, which formed a separate cluster, coincident with previous observations of clustering of Puget Sound samples (Chapter 3). In addition, Carr Inlet top sample (0-0.5 cm) clustered with Turning Basin samples. Besides these exceptions, clusters specific to each site were observed. Furthermore, clusters from close geographic locations were combined into larger clusters. So, Barrow Canyon Shallow and Deep, as well as East Hanna Shoal Shallow and Deep clusters, were combined into larger clusters from each transect, which, in turn, were combined into an Arctic cluster, clearly 128 separated from the Pacific Northwest cluster. The latter was, in turn, constituted of a Puget Sound cluster, combination of the three site-specific clusters from this area, and another one combining West of Juan de Fuca, Washington Margin and deep Puget Sound samples. The same pattern detected with cluster analysis, was also observed by principal component analysis (PCA) (Figure 4.8), which explained 27% of the total variability in the first two dimensions. T-RFLP patterns were grouped according to their sampling location, with samples from close geographic sites, being in turn placed adjacent to each other. So, Barrow Canyon Shallow and Deep samples are placed closely together, as well as East Hanna Shoal Shallow and Deep samples. F urtherrnore, Puget Sound samples form a compact group adjacent to the Washington Margin samples. T-RFs present only or at a higher abundance at one site or a group of sites were mainly responsible for the separation of these samples from the rest (data not shown). So, T-RFs 275 and 391 bp, present only in Arctic and West of Juan de Fuca samples, were mainly responsible for the separation of these samples from Puget Sound and Washington Margin samples. T-RFs 299 and 183 bp detected only in Arctic samples, although at low relative abundance, were mainly responsible for the separation of these samples along the first dimension. Barrow Canyon samples were mainly affected by T-RF 129 bp, in addition to several T-RFs of low relative abundance, all present exclusively in this area. T-RF 230 bp, present in East Hanna Shoal and West of Juan de Fuca samples had a main effect in their separation from the rest along the second dimension. The separation of West of Juan de Fuca samples was mainly affected by T-RFs detected only at this site, e.g. 198 and 328 bp. Samples from Puget Sound and Washington Margin were ordinated closely together, coincident with 129 cluster analysis. Although no exclusive T-RF for all these sites was identified, two T-RFs of higher relative abundance at these locations (104 and 272 bp) were mainly responsible for their ordination. . Canonical correspondence analysis was applied on one T-RFLP profile fiom each site, to study the influence of biogeochemical parameters in the differentiation of the denitrifier communities (Figure 4.9). The first two dimensions explained 67% of the total variability (43% and 24% in dimensions 1 and 2, respectively). Similar to the ordination achieved by PCA, Arctic samples were grouped together and the West of Juan de Fuca sample was separated from the other sites. In addition, Puget Sound and Washington Margin samples were ordinated closely together, with the exception of Turning Basin. Higher latitude and oxygen concentrations were correlated to the separation of the Arctic samples from the rest. The separation of the West of Juan de Fuca sample was mainly influenced by the higher water depth and lower organic carbon at this site. In addition, the higher nitrate concentration also influenced the separation of this sample, as well as Washington Margin along the first dimension. Higher, i.e. less negative, longitude, ammonium concentration and temperature were correlated to the separation of Puget Sound and Washington Margin samples from the remaining sites. Phylogenetic analysis of nirS gene clone sequences A total of 649 cloned sequences from five locations (153, 124, 137, 161, and 74 clones from Barrow Canyon Shallow, East Hanna Shoal Shallow, Shallow Budd Inlet, Washington Margin, and West of Juan dc Fuca sites, respectively) (Table 4.7) were identified as heme cd) nitrite reductase (nirS) sequences longer than 600 bp. These were translated and used to construct a neighbor-joining tree including closely related and 130 reference database sequences (Figure 4.11). For graphical purposes, subtrees were identified as depicted in Figure 4.10 and represented in detail in Figures 4.12 to 4.16. The divergence range in the cloned sequences was very large, with clones presenting similarity values between 5.3 to 100% (i.e. 38.8 to 100% amino acid identity). Six previously identified clusters (Clusters 1 to 6) of highly similar Shallow Budd Inlet clone sequences (Chapter 3) were again detected, in spite of the addition of nirS sequences from additional sites. These clusters included the same nirS sequences as described in Chapter 3, except for Cluster 2’, which differs from Cluster 2 due to the inclusion of clones NIRS4A H01 and NIRS4A D03 and the exclusion of clone NIRS4B A10, all from Shallow Budd Inlet, therefore no inclusion of sequences from other areas was observed. In addition, 26 other clusters of highly similar sequences were identified (Clusters A through Z). These generally included nirS sequences retrieved from one location only, although several clusters were constituted of sequences from two (Barrow Canyon Shallow and East Hanna Shoal Shallow) or three (Barrow Canyon Shallow, East Hanna Shoal Shallow, and Washington Margin) locations. Only 9.9% of the clones were at least 80% similar to cultured denitrifiers, with the highest similarity (86.9%) detected between clone NIRSlB C12, retrieved from Washington Margin, and Silicibacter pamerayi DSS- 3, a member of the marine Raseabacter clade. On the other hand, all clones were at least 49.9% and up to 100% similar (60.6% and 100% identical, respectively) to environmental clones retrieved from marine sediments (Braker er al., 2000; Santoro er al., 2006; Nogales et al., 2002) or water column (Hannig et al., 2006; Jayakumar er al., 2004; Castro-Gonzalez et al., 2005; Tuomainen et al., 2003) in other studies, with most clones (95.1%) presenting a similarity higher than 80% to some previously described 131 environmental clones. Interestingly, four Washington Margin clones and one from East Hanna Shoal Shallow (NIRSlA GlO, NIRSlA H11, NIRSlB A12, NIRSlA C11, and EHNIRS2H07) were 100% similar to nirS clones (pB49, pB6, pB46, pA5, and pB66, respectively) retrieved previously from Puget Sound sediments (Braker er al., 2000). The in silica digestion of the nirS clones from East Hanna Shoal Shallow, Barrow Canyon Shallow, West of Juan de Fuca, Washington Margin, and Shallow Budd Inlet generated 17, 17, 11, 17, and 9 hypothetical T-RFs, respectively (Figure 4.17), almost all correlated to T-RFs detected in those same sediments, after adjusting for a shift between real and hypothetical T-RF sizes described earlier (Chapter 3). Similar patterns between hypothetical and real T-RFs from the same sediment depths were therefore identified, with Sorensen similarity indeces of 0.61, 0.59, 0.50, 0.43, and 0.36 for Shallow Budd Inlet, East Hanna Shoal Shallow, Barrow Canyon Shallow, West of Juan de Fuca, and Washington Margin, respectively. Furthermore, dominant measured T-RFs generally corresponded to hypothetical T-RFs generated by a large number of clones, indicating that some abundant T-RFs are the result of coincidence in restriction site between different clones rather than by representation of a dominant nirS sequence. Quantification of clone clusters by real-time PCR The nirS clone clusters previously identified in the Shallow Budd Inlet clone library and quantified in Puget Sound and Washington Margin sediments by real-time PCR (Chapter 3) were additionally quantified in Arctic and West of Juan de Fuca sediments with the same primer and primer-probe sets used previously (Chapter 3; Griintzig er al., 2001). Clusters 2 and 3, the most abundant clusters in Shallow Budd Inlet, were below detection limit in all Arctic and West of Juan de F uca sediments (Table 132 4.3). Cluster 1 also was below detection limit at these areas, except for East Hanna Shoal Shallow, which presented a similar abundance as Shallow Budd Inlet sediments (1.6 x 105 nirS copies per pg community DNA). Furthermore, cluster 6 was detected only in Barrow Canyon Deep and East Hanna Shoal Deep, but at a low abundance three orders of magnitude smaller than in Shallow Budd Inlet sediments. Interestingly, cluster 5 was detected in West of Juan de Fuca and all Arctic sediments, although its abundance was an order of magnitude lower as in Shallow Budd Inlet. This cluster had been also detected in all Puget Sound samples, being absent or below detection limit in the Washington Margin samples (Chapter 3). This suggests that this cluster, though not abundant, might represent a fairly cosmopolitan group of denitrifiers. P. stutzeri nirS was detected in West of Juan de F uca and all Arctic samples, presenting an abundance similar to Puget Sound sediments and between 1 and 2 orders of magnitude lower than in the Washington Margin site studied earlier (Chapter 3), reaching a value of 9.0 X 104 nirS copies per pg community DNA in East Hanna Shoal Shallow sediments, equivalent to 0.04% P. stutzeri DNA in total community DNA. In order to identify significant differences in cluster-specific and P.stutzeri nirS gene abundance in Puget Sound and Washington Margin sediments, the previously determined abundances from different depth intervals between 1 and 10.5 cm sediment depth (Chapter 3) were considered biological replicates for ANOVA (Table 4.4). Shallow Budd Inlet had a significantly higher abundance of cluster 2 compared to the other clusters. This was also true for Turning Basin sediments, with the exception that it was not significantly higher as the abundance of cluster 3. In Carr Inlet, on the other hand, the abundance of P. stutzeri nirS was significantly higher than the abundance of the 133 quantified clusters. At this site cluster 1 and 3 abundances were significantly lower than in Turning Basin and Shallow Budd Inlet, respectively. Cluster 6 abundance was significantly different in all Puget Sound sites, while P. stutzeri nirS abundance was not significantly different between these locations, but significantly higher in Washington Margin sediments. Phylogenetic analysis of 16S rRNA clone sequences In order to study the general bacterial community, 16S rRNA clone libraries were constructed from the same sediments used for nirS clone library construction (Table 4.2). A total of 399 non-chimeric 16S rRNA clone sequences longer than 300 bp were obtained from the studied areas (81, 83, 80, 80, and 75 from Barrow Canyon Shallow, East Hanna Shoal Shallow, Shallow Budd Inlet, Washington Margin, and Westiof Juan de Fuca, respectively) (Table 4.7). Assignment of the sequences into taxonomic groups with the RDP Classifier allowed the placement of 61% of the 16S rRNA clones into a specific taxon with a bootstrap confidence estimate higher than 80% (Table 4.5). Proteabacteria were dominant at all stations, constituting between 33 and 51% of the bacterial community (i.e. 56 to 72% of the clones assigned to a taxon). The most abundant classes within this group were Gamma and Delta Proteabacteria, with similar number of sequences assigned to each of them in all cases, with the exception of West of Juan de F uca, which presented a lower abundance of Delta than Gamma Pratealiacteria, significantly lower (P S 0.05) than the number detected in East Hanna Shoal Shallow and Washington Margin sediments (Table 4.6). Although at lower abundances, Alpha Proteabacteria were identified at all sites, while Epsilon Proteabacteria were retrieved from Barrow Canyon Shallow and Shallow Budd Inlet only, being its abundance 134 significantly higher at this latter site than at East Hanna Shoal Shallow, Washington Margin and West of Juan de Fuca (Table 4.6). A single clone, retrieved from West of Juan de Fuca sediments, was assigned to the Beta Proteabacteria. The second most abundant phylum identified in Barrow Canyon Shallow and Shallow Budd Inlet sediments was Bacteraidetes, presenting a significantly higher abundance (P _<_ 0.05) than in the other areas, where they were detected at a similar abundance as Planctamycetes. This latter group was detected at all sites except at Shallow Budd Inlet, leading to a significant difference in abundance of this phylum between this site and West of "Juan de Fuca. F irmicutes, Cyanabacteria, Verrucamicrabia, Nitraspira and Actinabacteria, all had representatives in at least two of the studied areas, although each constituted less than 5% of the bacterial community at a site. In addition, a few clones were assigned to candidate divisions WS3 and CPU established in previous studies (Hugenholtz et al., 1998; Dojka et al., 1998). Between 30 and 51% of the clones could not be assigned to a specific taxonomic group with the chosen confidence estimate threshold, these could therefore, if assigned, increase the abundances of the detected taxons or belong to different phyla not included in Table 4.5. Individual phylogenetic trees for different taxonomic groups were built by neighbor-joining analysis including the 16S rRNA clones, closely related sequences from public databases and reference sequences (Figures 4.18 to 4.25). The assignment of clones to individual trees was based on the clustering observed in a general 16S rRNA tree including all sequences. The similarity between the retrieved clones reached from 41.4 to 100% (59.3 to 100% nucleotide identity), indicating a wide range in similarities. Clusters of highly similar clone sequences (similarity > 98%) were generally from the 135 same site, however no specific clustering pattern was observed for less related clones, with broader clusters including sequences from various different environments. Clusters of Gamma, Delta, and Alpha Proteabacteria, Bacteraidetes, Chlarabi, Acidabacteria, Chlaraflexi, F irmicutes, Planctomycetes, and Verrucamicrabia included clones from all studied sites. In addition, clones from at least one, but not all sites clustered with members of the Beta and Epsilon Proteabacteria, Deferribacteres, Spirachaetes, Actinabacteria, Nitraspira, F usabacteria, F ibrabacteres, candidate divisions ODl, OP11, and WS3, and Cyanabacteria. The latter were only detected in the shallower studied sites, namely Shallow Budd Inlet, East Hanna Shoal Shallow, and Barrow Canyon Shallow, being detected in the sediments probably due to deposition from the overlying water column. Several clones were closely related to environmental 16S rRNA sequences retrieved in other studies mostly from marine environments and also to cultured strains, though generally more distantly. The closest similarity with a cultured strain was detected in some Shallow Budd Inlet clones, which were highly similar (up to 99.6%) to Laktanella rasea, an Alpha Proteabacteria recently isolated from sediments of the north-west Pacific Ocean (Ivanova et al., 2005). The psychrophilic bacteria F lavabacterium frigaris and Psychramanas antarctica were closely related to clones from the Arctic sites Barrow Canyon Shallow (97.8% similarity) and East Hanna Shoal Shallow (94.1% similarity), respectively. Although Beta Proteabacteria and Nitraspira clusters included only one clone from Washington Margin and one from West of Juan de F uca sediments, the similarity between cultured strains in these phyla and West of Juan de Fuca clones was high. The ammonia oxidizer Nitrasaspira multifarmis from the Beta Proteabacteria and the nitrite oxidizer Nitraspira marina from the Nitraspira phylum 136 were 96.5 and 94.5% similar, respectively, to West of Juan de Fuca clones. Ammonia monooxygenase (amoA) gene sequences related to Nitraspira had also been detected previously in Pacific Northwest sediments, although at a different site (Nold et al., 2000). Although some individual or small clusters of clone sequences retrieved in this study could not be related to any cultured strain or environmental clone from other studies, most clone clusters included at least one 16S rRNA sequence from other environmental clone libraries. These related clones from other studies were often from close environments to our studied sites, as sediment clones from the Arctic Ocean off the coast of Spitsbergen island (Norway) (Ravenschlag et al., 1999), which clustered closely to several Barrow Canyon Shallow and East Hanna Shoal Shallow clones, or seafloor basalt clones from the Juan de Fuca Ridge clustering closely to clones retrieved from the same site in this study. However, most of the retrieved clones clustered with environmental clones from distant geographic locations and often dissimilar biogeochemical characteristics, including, for example, Mid-Atlantic Ridge (Lopez- Garcia et al., 2003; Nercessian et al., 2005) and Guaymas Basin hydrothermal vent sediments (Teske et al., 2002; Dhillon et al., 2003), Japan Trench cold-seep sediments (Li et al., 1999), Wadden Sea sediments (Mussmann et al., 2005) and water column (Stevens et al., 2005), and Mediterranean Sea sediments (Polymenakou et al., 2005; Heijs et al., 2005). Community description and comparison based on nirS and 16S rRNA clone libraries The nirS and 16S rRNA clone sequences were grouped into operational taxonomic units (OTUS) (Table 4.7) using a 5% and 3% similarity threshold, respectively, to allow 137 the description of the communities at the studied areas in terms of richness and diversity of phylotypes and to estimate the similarity between communities based on their shared gene sequences. Most OTUs were represented by only one gene sequence (Figure 4.26) in 16S rRNA as well as nirS clone libraries. However, these singletons constituted a larger fraction of the 16S rRNA OTUS (81 to 91%) than of the nirS OTUS (43 to 79%) from the different sites. On the other hand, the most abundant nirS OTU included three times more sequences (27 sequences) than the most abundant 16S rRNA OTU (9 sequences), both from Barrow Canyon Shallow sediments. The large fiaction of singletons in 16S rRNA clone libraries led to steep rarefaction curves far from reaching a plateau at all sites (Figure 4.27), indicating that the sampling effort was insufficient to cover the large diversity present. Rarefaction curves for nirS clone libraries, on the other hand, indicated that a larger fraction of the nirS community diversity has been detected, as a tendency to reach a plateau could be observed. Rarefaction curves also suggest a significantly higher diversity of nirS sequences in Washington Margin sediments than at the Arctic sites, which in turn present a higher diversity than Shallow Budd Inlet and West of Juan de F uca sediments. This is consistent with the Shannon diversity index and the Chaol nonparametric richness estimator results (Table 4.7), which indicated-that the Washington Margin sediments contained the most diverse (H ’=3.94) and richest (Chaol=348) nirS-containing community, followed by the Arctic sediment samples. West of Juan de Fuca sediments presented the lowest nirS richness and diversity (Chaol=41; H ’=3.07). Although Shallow Budd Inlet’s richness estimate was similar to East Hanna Shoal Shallow and more than twice as high as for West of Juan de Fuca, its diversity was as low as at the latter site. This is probably an indication of a less even 138 distribution of nirS sequences at Shallow Budd Inlet, which lowers its diversity. At all sites, richness and diversity was higher for 16S rRNA than for nirS gene sequences, with Washington Margin sediments presenting again the highest values (Chaol=574; H’=4.18), indicating that these sediments contained the richest and most diverse bacterial community from all studied areas. Although all the remaining sites presented similar diversity values, Barrow Canyon Shallow sediments contained a richer bacterial community, though less evenly distributed. The 16S rRNA as well as the nirS gene libraries from all sites were significantly different from each other (P<0.0026; a=0.05) as determined by the comparison of the difference between the homologous and heterologous coverage curves for each pair of libraries with the difference obtained after multiple randomizations of the sequences in the libraries (Table 4.8). This suggests that samples from sediments used to construct the different libraries may harbor distinct general bacterial as well as nirS-containing denitrifier populations. The significant differences between communities from different areas are further supported by the generally low similarity between libraries, as determined by the abundance-based Sarenson similarity index (Labund), as well as the low richness estimation of shared OTUS between sites, determined with a nonparametric richness estimator analogous to Chaol (SM; Chao) (Table 4.9). West of Juan de Fuca as well as Shallow Budd Inlet libraries (16S rRNA and nirS) presented extremely low similarities to all other libraries, with similarity values often equal to zero. This was correlated to the absence or the presence of only few estimated shared OTUs between each of these sites and the others. On the other hand, East Hanna Shoal Shallow, Barrow Canyon Shallow 139 and Washington Margin were more similar to each other and presented a higher number of estimated shared OTUS. The highest similarity (Labunf0.676) and estimated richness of shared OTUS (S A.B Chao=47) was observed between the two Arctic sites for the nirS gene, consistent with the observation of several clusters of high similarity including sequences from these two areas on the phylogenetic tree (Figures 4.10 to 4.16). The estimated number of shared OTUs represent, however, only a 42 and 48% of the estimated richness of nirS genes at Barrow Canyon Shallow and East Hanna Shoal Shallow, respectively, consistent with them being significantly different as determined before (Table 4.8). For the 16S rRNA libraries, the highest similarity (LabumFO-239) and estimated richness of shared OTUS (8A3 Chao=45) was observed between East Hanna Shoal Shallow and Washington Margin sediments. However, the number of shared OTUS represents only 21 and 8% of the estimated richness in the former and latter sediments, respectively. The two Arctic sites presented the next highest similarity and richness of shared OTUs estimation. 140 470' 465' 460' 45' sifl' 445' 440' 43' I East Hanna Shoal Deep Barrow Canyon Deep 72' I 72' East Hanna Shoal Shallow . | . 't-....esv--iy Puget Sound (Shallow Budd Inlet, Turning Basin and Carr Inlet) . West ofJnan de Fuca . 9 la. .27 -. Juan de Fuca ridge Vin: ‘_ . 440' ‘ . S Washington Margin _1‘ km Figure 4.1. Locations of sampling stations. For detailed view of Puget Sound stations refer to Chapter 3 (Figure 3.1). Table 4.1. Locations and characteristics of sampling stations. Water Or anic Sulfate . . g Temp. Salinity reduction rate Station Location depth carbon 0 .2 (m) (°/) ( C) (PSU) (mmoles m 0 day")' Shallow 47°04.99'N Budd Inlet 122054.1o'w 3 2'5 19'” 26‘2 12 Turning 47°03.13'N Basin 1220549995,“, 12 3.2 14.80 21.2 25- 47°17.21'N Carr Inlet 12294295,“, 84 1.8 16.80 29.6 7.5 Washington 46°25.57’N Margin 12404150,“, 1138 3.1 3.65 34.3 86 0.04 West of Juan 46°47'N de Fuca 133040,“, 3869 <0.5 1.56 34.661 ND Barrow 71°36.24’N Canyon 156°12.49W’ 186 1.6 -0.87 33.854 1.19 Shallow Barrow 72°12.33’N Canyon 154°5.56’W 2000 1.7 -0.402 34.942 0.3 Deep East Hanna 72°39.9’N ' Shoal 158°44.7’W 160 1.9 -1.32 33.407 1.5 Shallow East Hanna 72°53.64’N Shoal Deg) 158°15.78’W 1450 1.2 -l .66 32.601 0.05 3ND, not determined. 142 Table 4.2. Sediment samples analyzed from different sites. Sediment depth analyzed (cm) Analysis Barrow Barrow East Hanna West of East Hanna method Canyon Canyon Shoal Juan de Shoal Deep Shallow Deep Shallow Fuca T-RFLP 0-2 0-2 0-2 005 0-1 2-4 2-4 2-4 1-1 .5 _ 1-2 8-10 8-10 4-6 2-3 2-3 20 20 6-8 5-6 3-4 8-10 9-10 4-5 10-12 7.5-8.5 12-14 20-21 14-16 16-18 18-20 Cloninga 2-4 4-6 2-3 Real-time 2-4 2-4 4-6 2-3 2-3 PCR aAlso sediments fi'om Washington Margin (2-2.5 cm) were used for cloning in addition to analyses described in Table 3.2. 143 Figure 4.2. Pore water profiles for O; and N03” in Arctic (Barrow Canyon Shallow, Barrow Canyon Deep, East Hanna Shoal Shallow, and East Hanna Shoal Deep) sediments. 144 warm? 29 1° 9° .0 0219M) 0 50 100 150 200 250 300 350 o .. l A I .5. 1 ~ 5 § ., 2 4 Barrow Canyon Shallow 3 l A l a 4 N0,‘(uM) 9 20 40 so so 02(pM) 0 50 100 150 200 250 300 350 0; =3 #‘ = 5.; 10 <1 .1 I 20‘ l Depth (cm) 25 «, Barrow Canyon Deep 30 .l NO,'(pM) 0 20 40 60 80 OzluM) o 50 100 150 200 250 300 350 Depth (cm) East Hanna Shoal Shallow I; 1 4L % N0,'(uivi) o 20 40 60 80 4L 0201”) 0 50 100 150 200 250 300 350 0‘ 5! 101 I 15 1' Depth (cm) 20‘ 25 1 30 , East Hanna Shoal Deep L 145 Figure 4.3. Pore water profiles for F e(II), Mn(II), and NH4+ in Arctic (Barrow Canyon Shallow, Barrow Canyon Deep, East Hanna Shoal Shallow, and East Hanna Shoal Deep) sediments. 146 F0000"). M00001") 0 29° 4° 9° 9° 19° NH;(uM) 0 100 200 300 400 500 600 700 ° 1 A 5: E 2 i 10 . 8 15 1 Barrow Canyon Shallow 20 Fe("KIJ-Mli Maintain)? 29° 49° 59° 89° 109° NHfhtM) o 20 40 so so 0 L l . . c 10 - ’e‘ 2 i 20 8 30 . Barrow Canyon Deep 40 . . Fe("XuMI. M00001") 0 59 19° 19° 29° 2"” 39° NHJpM) 0 20 40 60 80 E 2 8 East Hanna Shoal Shallow 25 r 1 n Fe(ll)(pM), Mn(l|)(pM) o 50 100 150 200 250 300 NHflpM) 0 20 40 60 80 E 2. 5 D. 8 + NH: 20 « ' —-— Fe(Il) East Hanna Shoal Deep —o-—- Mn(Il) 25 ‘ 147 N0,'(plvl), NH,*(uM) o 20 4o 60 so 0201'“) o 20 40 60 80 100 120 140 160 Depth (cm) a 20 i + 02 25 %, —-— No,‘ 30 9 West of Juan de Fuca ° ””4 Figure 4.4. Pore water profiles for 02, N03" and NHI' in sediments from the abyssal sea floor west of the Juan de Fuca ridge. 148 EVI/I/I/I/II-‘ ‘ » \\\\\\\\\\\‘s§ 444'; .\ _E'l/l/l/I/l/A I4 -.1 “\\\\\\\\L+f ""WTW sg . I ~— =7/l/IIIII- u. Iii! p» " a 58 as g e-a:-_IIIIIIII;1II1 :1m\\\ M . aw _W A 1 m o = ”ll/Il/IIIIA- L» \\ _ . s» 1 I 72 a 76 :l/l/l/l/l/I/l/I/Ifl' . :\ t- \\ :6, 1 p . _\ ‘ a 83 n 86 _ :I/I/II/II/III/II/III ' s\\\\‘v+7__=_ '93» .1 m 91 a 97 =II/IIIIIIIIIIIIIII-1-J . , l a \\\\\\\\\\\\\\‘ am we: : .102 E104 a, :-\\\\\\\\\\\\\\\\ aw» ' ‘D 108 Um I \\\\\ "4 g,“\“\{*?_"_’"' [3113 I116 12129 I137 33 I 141 I 158i war—I; , z ,_ 1 1m161l173 A Q III/I In; :Illlll ' _ 1 E o )m M01 1: 176 I 194 2. m IIIII. 3198 I 2041 5 1 [:1 208 a 21 o i E LL. ‘ b g a 230 a 232 3 1:1 234 236 E I 238 I 257 .. rIIIIIIIIII;-: am _ ' , 1 g E III/III/I/IIZ" -:--III 1 Kiss: 0265 C1270 a 3 'l/I/I/I/I uIII III ‘ ' v ' IIIIIIIIIIIIIII.. .II-I ‘ a 272 a 275 'l/I/I/I/I/I/I/I/I ‘ : 1 III/IIIIIII :1 -:-::n l- 1 [:1 281 n 295 - III/IIIIIIIIIIIIIIIIIIIIIIIIIII IIIII ' ' : _- 1:3 297 I 317 o =II/I/III/IIII/IIIII/IIIIIIII .- . V 1 -_IIIIII/IIIIIIIIIIIIIIIIIIIII l a 328 I 332 Vl/I/I/I/I/ll/I/I/I/I/I/I/I/II- . V/I/I/I/I/I/I/I/Il/I/I/l/l/l/l/I/I/I/IIlEl i I 335 343 III/IIIII —mm: VI/Illllli-l—Ilflllllll mm 7 7 j i a 353 " 356 m III/IIIIIIIIIIrel—mt _— w .- g 15380 E11383 w x ‘ 1.,,”"1 taxman» 1:1 433 E 436 an» ‘ ,9 ”EH m ‘I 462 540 :23 546 I 589 Relative abundance of T-RFs (%) Figure 4.5. T-RFLP profiles obtained by amplification of nirS genes from Arctic (EHS, East Hanna Shoal Shallow; EHSD, East Hanna Shoal Deep; BC, Barrow Canyon Shallow; BCD Barrow Canyon Deep) and Pacific Northwest (WJF, West of Juan de Fuca; WM, Washington Margin; CI, Carr Inlet; SB, Shallow Budd Inlet; TB, Turning Basin) marine sediments from various depths. Presented T-RFs were generated after restriction with HhaI. Numbers in legend indicate T-RF length in base pairs for fragments representing more than 2% relative abundance. 149 Richness (number of T-RFs) 0 10 20 30 40 50 60 0 ‘ , -, a L , g 1 ‘r - ?" 20 q. . _ 1 ,_ _ , 4o _l,~._~, _ _____.__ - , _ _‘—_s_., , fl_ 1 +TB +83 E 60 - , ~ 7 k an +CI 3 / —x—WM .: +WJF H 0. +30 8 80 4 *fi * “‘ ‘ +300 +EHs +EHSD 100 + - . _.___.___ sewn 120 -- ~ —vw #77 , iffimgW 140 Figure 4.6A. Richness of the nirS-containing community in Pacific Northwest (TB, Turning Basin; SB, Shallow Budd Inlet; CI, Carr Inlet, WM, Washington Margin; WJF, West of Juan de Fuca) and Arctic (BC, Barrow Canyon Shallow; BCD, Barrow Canyon Deep; EHS, East Hanna Shoal Shallow; EHSD, East Hanna Shoal Deep) sediments from different depths, based on number of individual T-RFs detected after restriction with Hhal. 150 Shannon diversity index 0 0.5 1 1.5 2 2.5 3 3.5 +TB +88 +CI -x-WM +WJF —o—BC 80 *- ___,,,, * " *—--' r _ —e—BCD +EHS —e—EHSD Depth (cm) 100; ~ ~~~ -~~ ~ —- 120~——-— ——~**~ , —*A~~~~~———~——~*4 140 Figure 4.6B. Diversity of the nirS-containing community in Pacific Northwest (TB, Turning Basin; SB, Shallow Budd Inlet; CI, Carr Inlet, WM, Washington Margin; WJF, West of Juan de Fuca) and Arctic (BC, Barrow Canyon Shallow; BCD, Barrow Canyon Deep; EHS, East Hanna Shoal Shallow; EHSD, East Hanna Shoal Deep) sediments from different depths, as determined by the Shannon diversity index based on the T-RFLP results. The T-RFLP analysis was performed with restriction enzyme HhaI. 151 54 405 2] 135 0 MN MN ?@OAN ‘wdmwmeQ oawAummm QM m wwmwmg 95.4%: 010101 ? o . b 0001 u -‘¢?&#?o Migomm m AN-‘NDOJ N (.71 saw '01 ———_CDCDCDCDCDCD _DN 01 dN-‘(AJ :w9m wwfifinnmmmmmmnqqqqq ommm. D ”#40 . o pewmd mm D EHS 0" #00000" m I .Qodmmm N I O I O o m I U) _L U) _s m EHS1448 EH31244 EHSlOJZ EHSBJD EHSSG EHS4£ EH824 EH502 Figure 4.7 . Dendrogram obtained by cluster analysis of T-RFLP profiles. Hierarchical clustering analysis was performed applying Ward’s method on the Serensen’s distances between T-RFLP profiles from Pacific Northwest (TB, Turning Basin; SB, Shallow Budd Inlet; CI, Carr Inlet, WM, Washington Margin; WJF, West of Juan de F uca) and Arctic (BC, Barrow Canyon Shallow; BCD, Barrow Canyon Deep; EHS, East Hanna Shoal Shallow; EHSD, East Hanna Shoal Deep) sediments from various depths. Numbers next to sample names indicate sediment depth in cm. 152 x 6 , %X 0‘ XWM xXO 4 4 mo . OTB 3 ° 0 e 088 E: ‘2‘ 0‘ ACI ;; I A 0 TO , AWJF E -10 -5 2 o 5 10 1F 0 BCD 5 Eng? . 0 BC A AA .4 - i D EHSD am 'I I EHS 5; I I #3 Dim 1 (15%) Figure 4.8. PCA ordination of T-RFLPs of nirS genes from Pacific Northwest (WM, Washington Margin; TB, Turning Basin; SB, Shallow Budd Inlet; CI, Carr Inlet; WJF, West of Juan de Fuca) and Arctic (BCD, Barrow Canyon Deep; BC, Barrow Canyon Shallow; EHSD, East Hanna Shoal Deep; EHS, East Hanna Shoal Shallow) sediments. Percentage of variance explained by each dimension is presented in brackets. Dim, dimension. 153 WM 1 ‘ Organic Carbon f3 ° Temperature Ammonium Longnude N' Salinity itrate _05 1 Latitude BC.D '3‘: Dim 2 (24%) O O a ' Water Depth 4 ~ EHSD -2 - WJF Dim 1 (43%) Figure 4.9. Canonical correspondence analysis (CCA) based on T-RFLPs of nirS genes and environmental variables from Pacific Northwest (TB, Turning Basin; SB, Shallow Budd Inlet; CI, Carr Inlet, WM, Washington Margin; WJF, West of Juan de Fuca) and Arctic (BC, Barrow Canyon Shallow; BCD, Barrow Canyon Deep; EHS, East Hanna Shoal Shallow; EHSD, East Hanna Shoal Deep) sediments. Dots and vectors represent the sampling stations and environmental variables, respectively. Percentage of variance explained by each dimension is presented in brackets. Dim, dimension. 154 Subtree E Subtree D f —: H: Subtree C E P E.— (_— Subtrge B Li Subtree A > L— Azospirillum brasilense Sp7 (AJ224912) l——| 0.1 Figure 4.10. Schematic representation of the nirS phylogenetic tree shown and described in further detail in Figure 4.11. For graphical purposes, detail of subtrees A, B, C, D, and E is presented in Figures 4.12, 4.13, 4.14, 4.15, and 4.16, respectively. 155 Figure 4.11. Phylogenetic tree showing the affiliation of nirS clone sequences retrieved from Shallow Budd Inlet (SB), Washington Margin (WM), West of Juan de Fuca (WJF), Barrow Canyon Shallow (BC), and East Hanna Shoal Shallow (EHS) sediments to selected reference sequences. The tree was generated by neighbor-joining analysis of approximately 215 aminoacids with Azaspirillum brasilense Sp7 as the outgroup. Values at branch points indicate the percentage of 100 replicate trees supporting that branch. Bootstrap values below 50% were omitted. Names of clones retrieved in this study begin with NIRSl (Washington Margin), NIRSZ (Barrow Canyon Shallow), NIRS3 (West of Juan de Fuca), NIRS4 (Shallow Budd Inlet), and EHNIRS (East Hanna Shoal Shallow). Clusters of highly similar m'rS clone sequences are identified by numbers or letters, with quantity of clones in the cluster and retrieval site indicated in brackets. Clusters identified by numbers were previously identified (Chapter 3). Cluster 2’ is equivalent to Cluster 2 (Chapter 3), except for the inclusion of clones NIRS4A H01 and NIRS4A D03 and the exclusion of clone NIRS4B A10. The scale bar represents 0.1 substitutions per sequence position. sed., sediment; cl., clone; Wash, Washington; col., column. . 156 I so 1;“: Cuprgvli’dli’s nemlAr r)(X91394 l Ralstonia metallidurans CH34 (NZ AAAI03000011) 1 9 L Clusterz 2(3 WM clones fl NIRS1B 001 L ” ’ Huntington Beach sed. cl. th SD (DQ159648) l -— A” NIRS1A 003 ‘51 r EHNIRSZA03 1 Huntington Beach sed cl hbD 5A (DQ159611) l 1;: NlRS1A 11 100 NIRSlBC H06 1L_,vJ NIRSSA cos ,‘ 11m ‘NIRSSB (307 1_ 93 1LNIR51ACOZ 1 .1 7‘7—1 NIRS1B D08 1 H L * NIRS1BG03 ; _ NIRS1A B12 LH ‘ NIRS1A 303 1011 5* l} 2% EHNIRS1A02 7 a 99 NlRSZB H02 1 1“: EHNIRSZEO7 1, 1 ‘7 NIRSZAH10 9. 11“ Huntington Beach sed. c1 mm 40 (DQ159496) 7‘ j (NIRS1B ;: EHNIRS1H01 r NIRS1AA11 , 'u NlRS1A DOB 90 r NIRS4A E11 ’ *i A Hydrogenobacterthermophilus(A8210046) E RS1GO1 53 ‘L NIRS1A DOZ NIRS1BCOQ 97 NIRSZA G11 ,3 Lr NIRS1A E06 v L—“NIRSZA F05 ,, _ a, 1 1 F EHNIRS1E09 so - a ,, l___J NiRs1eBoa 100 ‘NiRs1B F05 ’ Wash. margin sed. d. wA20 (AJ248428) _r NIRS1B E08 ‘ 11. * Puget Sound sed. cl. p816 (AJ248420) ' ~ ** , ¥ *NIRS1B 1305 l __ ——* #’ "# Arabian Sea water col. cl. 6840-5A (AY336820) L»— , ,- Too < 31 _791 :NIRS1ADOS F 7—1 ‘~ NIRS1B B10 1. l7 NIRS1BH10 1* LB, ¥ NIRS1A H03 1 u Puget Sound sed. d. p36 (AJ248419) 1:H Puget Sound sed. cl pA17 (AJ248407) 1 NIRS1B ‘ ‘ NIRS1A H11 1 L NIRS1A £02 NIRSZB A05 1 ILNIRS1B F06 NIRS1B A08 ll NIRSZA GOQ 1 1 NIRSZB G09 1 NIRSZAA11 ELM NIRSZB CO1 5 1,1 Baltic Sea water col. d. M150-85 (DQO72183) 2 ‘ Baltic Sea water ool. cl. MSG-54 (00072203) 54— 06‘5"— Thauera selenatis AX (AY078264) l ’ _Thauera aromatica K172 (AY078256) 10°—Pseudomonas stutzeri ZoBell (X56813) ,7 L.-. Marine denitrifying isolate E4- 2 (AJ248398) 100 Alcaligenes faeaaI/sA15 (AJ224913) —Az “ll bra '/ nse 7 N224912 ospmfum 816 a) ( ) 0.1 157 86 0.05 55 ,1 ’ Subtree B l 1EHN|RSZE01 as “' 1%L Cluster U (2 WM, 11 BC and 6 EHS clones) ' Cluster V (1 BC and 2 EHS clones) f ‘ " ' “ ’ * Huntington Beach sed cl. th 4G (DQ159571) Thiobacillus denitrificans (NC 007404) 100 1 H Pseudomonas aeruginosa PAO1 (AEOO4488) L‘ ‘“ ’ “w Pseudomonas fluorescens (AF 197466) 1—* —— “m ‘— Cluster W (4 BC clones) " Marine denitrifying isolate B9-12 (AJ248393) "‘——“‘" ‘ Marine de nitrifying isolate 04-14 (N248395) ,0 ____#' -___5. Marinobacter aquaeoleiVTB (NZ AALGO1000002) 99 L——-~~ Marine denitrifying isolate 010-1 (AJ2483Q4) l I ‘ ‘ —_—Arabian Sea water col. cl. GB40-4F (AY336818) l 1 1 EHNIR82F04 5555 ‘~°° L NIRSZA 004 1 —- EHNIRS1E11 ,, L —-- , NIRS1B 003 1 92 1 ~_ NIRSZAA10 68 ; ,, , --55__ NIRSZA F04 91 1, ,, __ EHNIRSZBOS 100 H NIRSZA H01 LNIRSZB H01 ” ' Cluster 4 (4 SB clonesI‘ I 82 ”fl Huntington Beach sed cl DB 26(00159527) H ’" fl” *‘ ,,___ Roseobacter denitrificans (AJ224911) ‘1—41 EHNIRS1FO9 1 100 EHNIR82H08 11 '— Silicibacter pomeroyi USS-3 (CP000032) ’1 : --- NIRS1B H05 .1 100 _ ‘ Huntington Beach sed. cl. hbD 2F (DQ159599) 1 J5 A ~ - NIRS1A G12 ‘____ ‘ L 'H" _,“_ NIRS48 A12 34 1 Paracoccus denitrificans PD1222 (005002) 100 fParacoccus pantotrophus (AJ401462) I EHNIRS1GOI3 f EHNIRS1AO4 1*, EHNIRS1D11 5, 15 EHNIRS1AO7 " Cluster X (1 WM, 2 BC and 3 EHS clones) __512 ClusterY (5 EHS clones) I -I 5- NIRSZB E08 1 EHNIRSZFO1 1 I ~* NIRssA E01 r" ‘ NIRS1B C12 r ‘ ’4 Huntington Beach sed. cl. hbD BD (DQ159624) 151" ~~— NIRS1B H01 1 1 5— NIRSBA E02 151 I H NIRSSB GO8 NIRSBAA12 I' — NIRS1B F04 NIRS3AA1O 118—IL NIRSSA F06 “I 1— _ NIRS1AE10 1 __1 NIRSSA F10 11 L -5 NIRSSA H06 11* NIRSBB A06 ‘i1’ NIRS3A 604 r NIRS1B A04 11 .555 NIRSSB G11 L1; * NIRS3A 804 ' NIRS3B A11 531 159 Figure 4.12. Detail of Subtree A as described in Figures 4.10. and 4.11. The scale bar represents 0.05 substitutions per sequence position. sed, sediment; 0]., clone; col., column. 158 Figure 4.13. Detail of Subtree B as described in Figures 4.10. and 4.11. The scale bar represents 0.05 substitutions per sequence position. sed, sediment; cl., clone; col., column; E.S. Pacif. OMZ, Eastern South Pacific Oxygen Minimum Zone. 160 LA} ‘ ‘ “-9;— NIRS4A F0§Ubtree C 1 I‘— Cluster P (4 WJF clones) “ “#3; I EHNIRS1BOG . IEHNIRSZAOZ 93 I EHNIRS1E06 11 N1RS18 010 —Huntington Beach sed cl th 66 (00159579) 1 NIRS1B 007 99.5 NIRS1A E03 A L NIRS1B 006 1 — 7*1 . NIRS1A 808 1 ‘--—- N1R838 E12 5. _AI 1—-—- -—- 7- ,_ Huntington Beach sed. cl. hbD 4C (00159507) 113 i - 90 NIRS3A 805 L115 7 5&0 IL NIRS3B H05 1 I NIRS3B 809 .9951w NIRS1A H07 ,5 1 9‘ II '- NIRS1B H03 1— NIRS3A 808 ‘2 DI 'NIRSSB H01 76 Cluster Q (1 5 WJF clones) , 7 94,1 "‘ ' " NIRS1A F04 I L——— Huntington Beach sed. cl. th TG (DQ159582) 14‘ -- —*~ -- - 'NIRS1A 012 _ _ 75 _r:—"“——NIRS4A (304 1 [~71 1 Huntington Beachsed. cl th BG(DQ159557) ‘ 1 1 ;_ NIRS1A 806 I I l 64 .1" "~ EHNIR82F12 . rI L NIRSZB 002 5__ 5. - * NIRS1A E04 “7 L— NIRSiB A03 . . . -- - --‘— EHNIRSiD10 ' "7* Azoarws talulyticus 2FBG (AY078272) "" Cluster R _(1538 WM clones) SPacif OMZwaterooI d AC100—135 (AJ811517) I if_‘ EHNIRS1GO7E . ___ _JI NIRSZA 806 I1 100 NIRSZA F03 . ' " " Arabian Sea water col. cl. V483 -3EE (AY336925) ”fl III—1”? NIRS1A DO1 1 I I r4 I 1 Arabian Sea water col cl (5857-781AY336858) 1 HI LI; 1" EHNiRSS16804 EHNIR 1 10 I 10° LL“ NIRS1AA06 "‘ “m ‘ ‘ —‘TAtabian Sea water col. cl. V483-8H (AY336912) I 611,71 EHNlRS1807 1 EHNIRS1GOG M ‘ ‘ NIRS2B GOG 1’" I EHNIRszH11 1 NIRS1B E02 I 5’ As 1- NIRS1A BO4 1 I L—“‘NIRS1A 805 z 1 #555 Huntington Beach sed. cl. hbD 26 (00159600) _I EHNIRS2F06 1 III rrrrr -- NIRS1BCOZ '1 EHNIRS1BO3 I EHN1R32812 1. 5, N1RS1BA06 : I “‘7'”— NIRS1B 005 . I EHNIRS1E03 EHNIRS1007 * , ’NIRS1A H08 EHNIRSZGO1 .__.ea_; - —~N1RS1A cos — NIRSiA (302 _I EHNIRS1E04 EHNIRS2F08 I EHNiRsone IEHNIRS1A06 1;“ IEHNIR81C08 I” 8 11 'EHNIRS1A05 EHNIRS1B11 III EHNiRsona EHNIRS2D10 :— EHNIRS1012 I1 EHNiRsons I IIEHNIRS1DO4 I . ., EHNiRszoos I I ‘1 NIRS1B A10 '1 ‘NIRS1B H07 6‘ 1. EHNIRSZG10 7‘ -_ NIRS1B F09 MIII NIRSZB A03 NIRSZB c02 7' Cluster T (17 WM clones) 161 fl“— ree E 77 Su btreells NIRS1AA04 NIRSZA 807 109 'NIRSZA F01 62 I II NIRSZB E07 1 I 1 83 LNIRSZB H11 . 71 —7 EHNIRS1F12 77 I LIV—I— NIRSZB F09 7 - ___I 7—7NIRSZAA03 I 55 17 7 _ Baltic Sea water col. cl. M60-145 (DQO72215) -— EHNIRS1CO4 1 - - - - - - -—— NIRS1A CO7 I ~ ~-—~- --r- -~ NIRSSB H10 ’ '53' EHNIRS1G08 JP NIRSZB 607 1 96 L EHNIRSZEOS ”'7 7 777* 7' 7 Cluster M (7 BC and 2 EHS clones) NrRszAA07 Puget Sound sed. cl. p820 (AJ248421) I' NIRSZB D10 99 NIRSZB F11 71 1" I NIRSZB F08 .- EHNIRS1E07 L - -— NIRS2B 007 We I— EHNIRSZDO4 i‘II mmam1 ‘ NrRssA (308 59 L__ 1 as LI NIRSSA 806 1 1 I 91 't——- NIRSSB F04 I NIRS3A B10 51 , NIRSZA (:05 ' . ' NIRS3B F02 ' 53 r— NrRsaAAoe "—TLI NIRS3AA05 12'--— NIRS3B B10 117-77 Cluster N (9 WM clones) ' .‘EHNrRszcrz f__1 EHNIRSZA07 2 L~- NIRSZA E08 —— NIRS1A H10 'NIRS1B B11 77 r—— EHNIRszoos 1 L—— NIRS3A C11 Arabian Sea water col. cl. GB40-6A(AY336822) I as I EHNIRS1B10 ' 55 JL NIRSZB 005 1. I ’ EHNIRSZB10 8° 6‘iIII_N18328 803 NIRS1B E04 , EHNIRSZDOZ . 01 ~-w NIRS1AA05 - as {~— NIRS1A E12 fig ___“ NIRS1AA08 I L 61 L—~ NIRSZA (304 I ‘ .7 Arabian Sea water col cl V483-4E (AY336904) 7 '93 7 Cluster 0 (5 BC and 2 EHS clones) I52__ rt -~1 0.05 Figure 4.14. Detail of Subtree C as described in Figures 4.10. and 4.11. The scale bar represents 0.05 substitutions per sequence position. sed, sediment; 0]., clone; col., column. 162 Figure 4.15. Detail of Subtree D as described in Figures 4.10. and 4.11. The scale bar represents 0.05 substitutions per sequence position. sed, sediment; 0]., clone; col., column; Wash., Washington. 163 so 1'{ -~ EHNIRS2E05 I" NIRSZA H02 . ._ Wash. margin sed. cl. wA15 (AJ248427) Lu NIRS1AA1O 56 :1 95 L NIRS1A 006 f1. ; Cluster l (1 BC and 3 EHS clones) 98 ; {EHNIRS2801 ' {*NlefiincgaiAos L -—1 .r-- NIRSZA £11 6' = 66 EHNIRSZCOS ! Ln NIRSZA 001 , I :— NIRSZB 805 1 99 ‘ NIRSZB H03 1 . ,100 1—*"'N|RS4A 810 1 t ‘ ~ Huntin ton Beach sed. cl. th 6F (DQ159546) | a too 7 IRS4A 806 , 9' 99 8.2 T: NIRS4AC11 { 62 l 4. NIRS4A 806 E t’ —— Huntington Beach sed. cl. th 9F (DQ159587) ; 72 l 1 -fi~ NIRS1A C10 ' 1 “snares ' 9° “Ct—EHNIRSZGO6 ,0 or NIRS4B F09 I 93 L Puget Sound sed. cl. pA33 (AJ248412) _29_J NIRSZA H05 .g t t— Baltic Sea water col. cl. M60-156(DQO72220) 9, {___ _1 EHNIRStDOQ 96 LJ— EHNIRS2006 93 77 1— EHNIRSZDOB fiJ [ , 92 ; ~- Cluster J (4 BC clones) c, 1 ‘ NIRSZA H03 1 La”, f_) NIRSSAD12 , mo 1_r~w~— NIRSBB 510 ,. 89 L NIRS3B H09 ( ,, M NIRSZA 001 1 , J , 190 tNlRS3B 801 L L_ 1 'NIRSSB 007 t 9917—— NIRSSB B12 ; 1 5|— NIRSSB C06 ‘ 55100 — NIRS3B F03 ) 84 ; Cluster K (4 WJF clones) 3 viii L—a Wash. margin sed. cl. wF16 (AJ248437) L ' L_; NIRSSB A05 100 ‘ NIRSBB E09 i __ t_‘ _ t . ~ _. NIRSBA H02 i “411 » ~ NIRS1ADO3 ; 122‘ — NIRStB D12 7‘ 89 r NIRS1A GOG 1 99 9‘ 1' 'NIRS1B c302 ‘ ‘ 1 NIRS1B 804 t NIRS1B G11 1 L NIRS1A F12 y—“h - Huntington Beach sed. cl. th 4C (00159534) 1 L824 NIRSZB 009 W _J ‘NIRSZB H09 u.__1 NIRS3A 1503 100 L4 NIRSSB 007 76 ) NIRSBA 004 [a 19’1" NIRS3BE11 ‘2 76,1 NIRS3B 004 1 Lar— EHNIRS1H05 1 1 75 L130“— NIRSZB G10 " 54 ‘r— Cluster 5 (4 SB clones; ‘ 7 _ _ 1 Huntington Beach sed. cl. hb 7C (DQ159549) ;__; 100 L Huntington Beach sed. cl. th 1E (DQ159522) 86 1 ;' ‘—_ ’ NIRS4A H05 ~ : ea NIRS4A 001 82 .., 1NIRS4A E12 72 ‘85“ ‘ Cluster1 (21 SB clones) 1————~—+ 0.05 164 Figure 4.16. Detail of Subtree E as described in Figures 4.10. and 4.11. The scale bar represents 0.05 substitutions per sequence position. sed, sediment; 0]., clone; 001., column; cyanob., cyanobacterial. 165 - Cluster 2' (29 SB clones) NIRS4A H09 NIRS4B 611 ——— NIRS4A Ho7 NIRS4A 802 NIRS4A 603 NIRS4B 011 9r __NIRS4A 012 L‘_NIRS4A Hoz w NIRS4A 01o Loo NIRS4A 11'_N Cluster BE(11013 SB clones) 1‘1‘BRiter Cohe estuary sed. cl. ANIS—54 (N440470) Puget Sound 2sed. cl. pA12 (AJ248405) F ' NIRS4A F1: NIRS4A F11 NIRS48 602 _ 69:13:12“ Cluster A (14 88 clones) 1 —1 NIRSS4A 605 1 1 98 NIRS4A 608 1 ! 97 LNIRS4B 008 "—1N_1F}S1A F09 1 NIRSS4A 809 L NIR 4AF02 __63 82 L1 NIRS4A 804 551—" NIRS4B (:09 _I_" NIRS4B C11 L‘fi Baltic Sea cyanob. aggregate cl. BS1270 (AJ457196) 1 ' NIRS4A H10 1 1 1 61_1--N1RS48 808 1 __ 77 J _fl NIRS4B F04 NIRS4A F10 1 1 l 1 J ——- -' NIRS48A10 1 *~N1RS48 010 11 1—— NIRS4B (:02 ---- NIR '- Cluster BE (6 SB clones) EHNIRS2807 : V 2_,, NIRS48 (:10 4-- J EHNtRS1605 571 93- N1RS18 A12 1 ,3 '— Puget Sound sed. ct. p846 (AJ248422) 11 ‘ -2 EHNIRS1GOZ ~— NIRS4A 612 EHNrRszHo7 1 88 N1RS1A 610 1 Puget Sound sed. cl. p849 (AJ248423) _59- 1 - NIRSZA Hoe 11 Puget Sound sed. cl. p866 (AJ248424) t 11—- N1R618 EDS 1 L N1RStB E12 Cluster C (5 BC and 5 EHS clones) BalticSS1eam1 water col (:1. NBO- 76 (00072204) --~~~—— NlR 11f N IR Cluster D (1 BC and 5 EHS clones) 99 $111 : C'UStgbe E (8 BC clones) anglers/1F (80 BC and 2 EH8 clones) —“-*—- NIRS1B (:05 52 N1RS18 E03 96 NIRS18 E10 '7 Cluster 6 (3 SB clones) NIRS1A C11 001J1Puget Sound sed. cl. pA5 (AJ248403) , N1RS18 608 NIRS1A Hoz ‘NIRS1A HOG NtRS1A F01 55 NIRS18 511 6 F NIRStA 004 1 N1RS18 F06 £911N1RS18 604 ! L N1RS18 CO1 LN1R518A02 _g' 58 NtRS18 612 M 51;; EHNIR81F10 . 511 '- "L’NIRSZADOZ 1 1 ._.r—N1RS18609 1 1‘ .1‘— —'““‘ NIRS1A (307 1 ‘ - N1R838 602 7" ~~ Battic Sea water col. (:1 MSG-165 (00072223) r NIRSZA C11 92 100 NIRSZB CO71 1 EHNIRszcto 11'th, NIRS1AA12 11605 L Cluster 61(36 BC and 10 EHS clones) 1—: EHN1R32811 - v» NIRS3AA06 96 ‘ "~— NIRS3AE12 51871— EHNIRS1DOB 5‘1 EHNIRSZEOZ ~ "‘NIRS1AAO3 — EHN1R51601 L1____1 :‘ ~Hunungton Beach sed ct th 8F (00159664) 95 ——--- -—~-- ~ Cluster H (5 WM clones) 1_—._4 0.05 166 8-10 I 10-1 2 ul 12-14 14-16 16-18 18-20 EHS clones 0-2 0 2-4 0 8-10 20 BC CIOHOS WJF clones Depth in sediments (cm) 04).!) s :1: g - 1040.5 3647 62 SB clones (1 ( \ I < . I-l'l-l-III 7 .7'.\f.' 3.3 -'.'. .'.'.'. “(41111“ L‘Ifll‘l‘tl ttttttttttt 1).), ........ 20% 40% 60% 80% Relative abundance of T-RFs 111. ‘0 511(505) 540 | P '87 58 0 64(70) (:1 83 a 86 m 91 n 97 ‘0102(108)a104(109) D108(112)C1111 :3 129(133) I 141 I 158 121 161 I 173 198 121208 n210(212) (a 230 a 232(234) a 234(237) n 236(239) a: 238(241) I 257 :1 270 1:1 272(274) n 275(277)13 281 1a 295 :a 297(299) 1I 299(302) 303(304) I 317 (a 328 I 332 335(337), I 348(350) a 353(356) a 356 n 359(362) a 380(382) m 383(385) 131 388(390) a 391(393)‘ n 433(435) 1:2 436(437) E] 546 I 589(587) 1:1 595(590) 100% Figure 4.17. Comparison of T-RFLP profiles obtained by amplification of nirS genes from Arctic (EHS, East Hanna Shoal Shallow; BC, Barrow Canyon Shallow) and Pacific Northwest (WJF, West of Juan de Fuca; WM, Washington Margin; SB, Shallow Budd Inlet) sediments from various depths to distribution of T-RF sizes obtained by in silico digestion of nirS clones retrieved from those same sites. Presented T-RFs (real and hypothetical) were generated by restriction with HhaI. Numbers in legend indicate real T-RF lengths in base pairs for fragments representing more than 2% relative abundance or correlated to a nirS clone. Values in brackets indicate corresponding T-RF size generated by in silico digestion of clones. 167 Table 4.3. Cluster specific and P. stutzeri nirS gene abundance in Pacific Northwest and Arctic sediments as measured by real-time PCR. Sitea Average nirS gene copy number per 11g community DNAB Cluster 1 Cluster 2 Cluster 3 Cluster 5 Cluster 6 P. stutzeri WJF BDL BDL BDL 1.3x 103 BDL 6.9x 104 (6.2><103) BC BDL BDL BDL 4.5><103 BDL 2.7x 104 (2.7x103) (2.1x103) BCD BDL BDL BDL 8.7x103 4.1><102 7.3><104 (5.1x102) (1.7) (1.7x10“) EHS 1.6><105 BDL BDL 8.0x103 80L 9.0x10“ (2.0x10“) (3.3x10‘) (3.1x104) EHS BDL BDL BDL 2.1><103 7.0><102 2.3x10“ D (2.0x103) (7.9x103) aWJ F, West of Juan de Fuca; BC, Barrow Canyon Shallow; BCD, Barrow Canyon Deep; EHS, East Hanna Shoal Shallow; EHSD, East Hanna Shoal Deep. bValues in parenthesis are standard deviations of duplicate assays. Single assays were performed when no standard deviation is indicated. BDL, below detection limit. 168 Table 4.4. Cluster specific and P. stutzeri nirS gene abundance in Pacific Northwest sediments as measured by real-time PCR and determination of significant differences by ANOVA Sitea Average nirS gene copy number per 1.1g community DNAb Cluster 1 Cluster 2 Cluster 3 Cluster 5 Cluster 6 . P. stutzeri SB 1.4x105 2.0x10" ”"3 5.3><10SM 5.2x104M 1.1><10SM 8.4x10T “1"," (1.2x10") (3.3x105) (1.1x10‘) (3.1x10“) 9" (3.9x10“) (4.4x103) TB 1.6x105 81" 1.1><10""LB 4.3><105 “4",” 3.5><10‘”‘*A 5.3><104M 5.2><104 (9.8x 104) (7.7x105) (3.1x105) (4.5x104) (2.8x10“) “A (1.4x104) CI 2.7><10""’A NI 1.4><10‘””A 3.4x103M 2.1><103°’A 7.8x104 (1.4x104) (4.5x103) (1.6x103) (4.5x102) “’3 (4.7x10‘1 WM N1 N1 N1 N1 N1 1.1><106 (5.3x105) aSB, Shallow Budd Inlet; TB, Turning Basin; CI, Carr Inlet; WM, Washington Margin. bQuadruplicate samples were analyzed. Different lower case superscript letters (a,b,c) indicate significant differences between sampling locations as evaluated by one-way ANOVA (Tukey's test) at the 0.05 level. Different capital superscript letters (A,B) indicate significant differences between cluster abundances as evaluated by one-way ANOVA (Tukey's test) at the 0.05 level; Values in parenthesis are standard deviations. NI, not included in analysis. 169 Figure 4.18. Phylogenetic tree showing the affiliation of 168 rRN A clone sequences from the Gamma Proteabacteria retrieved from Shallow Budd Inlet (SB), _ Washington Margin (WM), West of Juan de Fuca (WJF), Barrow Canyon Shallow (BC), and East Hanna Shoal Shallow (EHS) sediments to selected reference sequences. The tree was generated by neighbor-joining analysis of approximately 550 bp with Thermus thermophilus and Thermus aquaticus as outgroup. Values at branch points indicate the percentage of 100 replicate trees supporting that branch. Bootstrap values below 50% were omitted. Names of clones retrieved in this study begin with 168 1 (Washington Margin), 16$ 2 (Barrow Canyon Shallow), 16$ 3 (West of Juan de Fuca), 168 4 (Shallow Budd Inlet), and EH 168 (East Hanna Shoal Shallow). The scale bar represents 0.05 substitutions per sequence position. sed., sediment; cl., clone; Pac., Pacific; Atl., Atlantic; hydroth., hydrothermal. 170 L EH1681 D08 5411 East Pac. Rise basalt glass cl. 9NBGBact 23 (00070798) _ .. 168 38 E02 1—1 166 3AH09 1 66 3A E09 154. EH1682 A02 1 ~ Arctic sed cl SvaO115(AJ240974) 1 . 1661BF04 53— 166 418E 1 1, 1 EH16S1A05 ‘oafLLL EH 1681 A07 EH16 160 S1 COS 11 __..________ 16S1ACOS . ,:LLLLLL— 168 4C A10 LLJa taGnSTian%h cold- -seep sed cl. JT8255 (ABO15254) 1 EH16S1 E04 .4 Japan Trench deep~sea sed. cl NB1-h (ABO13829) 1! 16$ 3AE11 - _r-L L EH1681 H06 _1 1 ‘LL-- . LLLLLLLLL 16848811 1“ 1 LLLL16S4B H08 , _4- L L ___ EH1681 807 L—L— 16S 18 H08 i 168 38 805 ‘ L __ 168 38 003 ,. _.2..-V 991L168 38H10 V ‘9: Mid-115689., Ridge2 hydroth vent sed. cl. AT-62- 59 (AY225636) ._1 1 __ i ’168 3A007 1 78I —1 168 3A E 1 254.... ‘ "”128 noiSS 3”” i 1 L-LLLL- L-L . 16838001 L1—1 16$ 2AC11 1 L—LLL 168 38 E04 1 971*“ 99F EH1GS1 E09 L1 ,. Pac. whale carcass microbialmatcl131760(AY922232) _ 1 __ EH1681 003 811 1"“LL‘ EH1682 F12 89L - -. 1EH16$1 (306 - 65 LLLLL1682A 009 ! 1 EH16SZ 606 T L LP ; LLLLL1681AG12 1 "" ‘ "___1' L 168 40 C07 . 95 6,12, 166 48 002 1 ‘. __ Wadden Sea bulk water cl GWS- K5-2 (AY515458) 1 1 ~— -— —- .._ Beggiahtoa alba (L409 94) 1 1 L hioploca ingrica (L40998 . _ ._ ,, -. “.99 LLL— Pseudomonas stutzeri (U26 62g 1 L LPseudomonas aeruginosa(XO 684) | .__ -2..,W_EH1631302 LL—, 1 Pseudoalteromonas denitrificans X82138) ’ ”“L ““ ‘EL_____1LLL LL Psychromonas profunda (AJ4167 6) 981._____r—LL _E1H16S F10 L _ Psychromonas antarctica (Y14697) __- i Leucothrix muoor (X87277) A 1 “L L LL __ __ __ LOceanospi'n'I/um mans (A8006771) LL LLL _ _ LL _ LL“ Alcanivorax borkumensrs(Y1257 9) F1 .1—~~ _ ~77 - — EH1662 011 ' Macrobu/bifer maritimus (AY 377986) 1 63.1 168 488 157611 LL168483F02 991 1 168 4CF01 , L Gua mas;3 Basin hydroth. vent sed. cl. C1 8038 (AF420367) .~ 571 168 48360 i. ___- _2 V0 LL— 16S1AH10 r P» 168 4 F0 08 1 1___.__, - ”tangy/068mm? fibrata (AF177296) 1 égh 1 .1 L LL16$1B G02 1 777 ~64 163 28 008 1 2.. L VjééVfi- L 168 28 E02 1 68 _--L L 16828010 1 166 48 F09 1 168 38 A12 : 99L LLLL LLL1683BG1O m: - 781L—LLLL—LL 16S 18 806 1 __1 'LLLL'LL LLLLL 16 G06 1‘_—1-_.__._-_ LLL LL LLL LL “L “168 2AE07 1 EH16S1 602 _ 9V ILArctic sed d. Sva0091 (AJ240987) 66 28 A08 1— EH11631 611 86 EH16S1 CO1 9’26 591 marine h Srocarbon seep sed cl SBseep4(AY456982) 1A 608 3° 791———1816618A02 _J| 1 LL LThermus thermophilus (1X07998) 99L L LThermus aquaticus( 09663 171 Figure 4.19. Phylogenetic tree showing the affiliation of 168 rRNA clone sequences from the Delta Proteabacteria retrieved from Shallow Budd Inlet (SB), Washington Margin (WM), West of Juan de F uca (WJF), Barrow Canyon Shallow (BC), and East Hanna Shoal Shallow (EHS) sediments to selected reference sequences. The tree was generated by neighbor-joining analysis of approximately 550 bp with Thermus thermophilus and T hermus aquaticus as outgroup. Values at branch points indicate the percentage of 100 replicate trees supporting that branch. Bootstrap values below 50% were omitted. Names of clones retrieved in this study begin with 168 1 (Washington Margin), 168 2 (Barrow Canyon Shallow), 168 3 (West of Juan de Fuca), 168 4 (Shallow Budd Inlet), and EH 168 (East Hanna Shoal Shallow). The scale bar represents 0.05 substitutions per sequence position. sed., sediment; cl., clone. 172 L 99 ”—641 1663A 805 9994f 16S 48 F10 ‘Wadden Sea bulk water Cl GWS- Kdna24 (AY515480) 1 168 4861 10 | 1_ _ 1168 4C E05 1 98" 1L 168 4C 803 l 54 163 4C E11 1 , "v168 48 608 1 LLL ' " L 168 28 C03 581 89 _1———16328G06 "" i 1—L— EH1682 E12 1 95 7., Arctic sedcl dSva1036 6110240990) 67 L LLDesquorhopalus vacuolatus L42 1 581________Aérc1166§eé<188156va10 13 (AJ 40982) : 99 L- 67: 96'th EH1662 F09 , 1 L—L—LL Desulfota/eagsychroghlla (AF099062) F" “LLLL S 2AE 2 99 ,1 ' -~L EH1BS1F01 74 1 9:3,1EH1681 C04 99‘ , L EH1682 C10 4 Arcticsed d Sva1041 (AJ240984 1Wadden Sea sed d SK4 (AY771 54) 9794 EH1682 E02 EH1682 608 991—7 168 1B 807 7 ~591 Caswdia Margin hydrate ridge sed. cl. H d89—23 (AJ535245) __ 1 Desulfacinum hydrotherma/e (AF170 17) " 1LL L LL LLLLLLL 16S 18 E08 1% ¥ 1 1m _ 64 168L18086ulfonema magnum (U45989) 99 IL— _ : Wadden Sea sed cl SK6(AY771941) 9°! 3111682 603 797 93 1 i 51 1 1 168 48 E06 1 6891011 EH1662 801 EH1682 C08 LL L LLLL LLL LLLL L Desulfonatronovibno hydrogenovorans (X99234) ‘ '1 1‘ " 'LLLLLLL Malonomonas mbra (Y1 17 2 'LL—LLLLL Desu/furomongsgoa/mltatls (U28172) 1 991 --.-—r l 816328 [ 991 - WgSsedB d ESvaO556 (AJ241010) 1 1 1 l 1 L 168 3A C05 991 L. LL16S 3A H05 fivc—W—u _. v~1__ L LL Juan de Fuca Rid e basalt lass cl JdFBGBact 10 (00070818) LL Nitrosprna 6gracrH rs O(2L355) _._fi _,. .1 a, __ LLLL 1633A 809 1 1. 1 99 LJamgikergggcoid- seep sed cl JT836(A8015242) ,,,_ 95 rL—L 16 ‘1 99: EH1JESPan05 Trench deep-sea sed d N81 -1 (A8013831) L L L .- ‘ L _ LL L‘ __ “ LLL ”L 168 AOE3 1 E2601 Pelyangium cellulosum (00256395) 1 1 ‘LL L L L LL L' fiannxyais exegearbs (AJ233946) -- - 51:1— 16S1AG11 1 7 1662Ao11 1 ,_ , , , L__EH1166§2AH 00% ,, c. -, 1 M 1—A1rct1csec1 c1 Sva1009 (AJ297456) 1 1681 BG05 1 _EH1661 E151 ‘16 9rEH 16 682 A07 LArctic sed. cl. Sva0537 (AJ297473) _._ L LL LL 16S 3A6F05 - H g LArctic sed1mentc|oneSva0506 (AJ241006) 9" 1— ___ L - EH1662 010 Thermus thermophilus (X07998 99 LL— Thermus aquaticus (L0966 ) 005 173 Figure 4.20. Phylogenetic tree showing the affiliation of 168 rRNA clone sequences - from the Alpha, Beta, and Epsilon Proteabacteria retrieved from Shallow Budd Inlet (SB), Washington Margin (WM), West of Juan de F uca (WJF), Barrow Canyon Shallow (BC), and East Hanna Shoal Shallow (EHS) sediments to selected reference sequences. The tree was generated by neighbor-joining analysis of approximately 550 bp with T hermus thermophilus and Thermus aquaticus as outgroup. Values at branch points indicate the percentage of 100 replicate trees supporting that branch. Bootstrap values below 50% were omitted. Names of clones retrieved in this study begin with 16$ 1 (Washington Margin), 16$ 2 (Barrow Canyon Shallow), 16$ 3 (West of Juan de Fuca), 168 4 (Shallow Budd Inlet), and EH 163 (East Hanna Shoal Shallow). The scale bar represents 0.05 substitutions per sequence position. sed., sediment; 0]., clone; isol., isolate; hydroth., hydrothermal. 174 i J 1 r 1 519H,:LCalifomia marine sed. d. (AY193227) LoktaneI/a msea (AY682199) flLArctic sea ice isol. ARK10226 (AF468375) 163 4C H02 641.1L Sulfitobacter mediterraneus (Y17387) “sf—LL Raseabacter denrtnf cans (L01784) L16S 40 E09 :168 48 D04 1___163 18 B03 L—LParaooccus denitrificans (Y1 692 8) 1______1633A 3A802 Magnetospiril/um magnetotacticum (Y101 10) F 1 85L— 168 1A 805 761 1 — 168 4C E03 __ Methylosinus sporium (Y18946) LLL Methylocystis parvus (M29026) .521 LE ,__J—165 1A C09 3LLL— 163 3A E10 ‘ _rL LL- 1681B E02 1_1 __ EH1681 F09 —95—. 1,. - L BgréogglEfi/izabethae (L01260) 1 _ _ _ L L L 1 l 90;_____ .. L163 38 F05 moL Juan de Fuca Ridge basalt glass cl. JdFBGBact 41 (00070828) 168 2A F12 -EL 168 18 COB __ E Spin'llum volutans (M34131) ‘, L—LLL LLLLL Alca/igenes faecalis (M2 2508) LLLT ILL—LL Cupn'avr'dus necator(M32021) L {31 Thauera aromatica (X77118) 1 _ J L w 168 38 D10 ‘1001 Nitrosospira multifarmis (L35 509) Nitrosospira bn'ensis (AY123800) 168 28 804 F109 L Mi.d-Atl Ridge hydroth. vent sed cl le3BBOS(AY354174) 11 East Pac Rise basalt glass 0|. 9NBGBact 3 (00070790) ( , 191 ___ LLLLL— Wolinella succinogenes (M88159) 1 L1; LLLLL 16$ 2ACO4 ,- 64 ___ Thiomicrospira denitrificans (L40808) LL LLLLL Aroobacter cryaerophilus (L14624) 1 _1L LLLLLLLL Sulfurospirillum arcaohonense (Y11561) 1. _ Sulfurospirillum bamesii (AF038843) 861- 16S 48 808 55;“ '166 4C 605 168 4C 811 60' . South Korea tidal sed. d. (AY304364) , 9811 166 4c 801 1 L 166 4C 005 ' ' _ 168 4C (309 99 70 1682AH10 L“ 168 28 G02 [—1168 28 H10 Guaymas Basin hydroth. vent sed cl a1b001 (AF420344) 1V6 16S 28 G05 1 51_m 166 48 H03 L _99;— 166 2A H06 1 64131.1" '168 213 H06 I L 168 48 001 1 Ldeep-sea sed. d 807- 9 (A8015584) 168 1A 001 168 2A E03 95 L168 28A05 16S 48 010 __2, ILL Thermus thennophilus (X07998) 100 1.___ Thermus aquaticus (L09663) ._J 175 Figure 4.21. Phylogenetic tree showing the affiliation of 168 rRNA clone sequences from the Bacteroidetes, Chlarabi, Deferribacteres, Spirochaetes, and Acidobacteria retrieved from Shallow Budd Inlet (SB), Washington Margin (WM), West of Juan de Fuca (WJF), Barrow Canyon Shallow (BC), and East Hanna Shoal Shallow . (EHS) sediments to selected reference sequences. The tree was generated by neighbor-joining analysis of approximately 550 bp with Thermus thermophilus and T hermus aquaticus as outgroup. Values at branch points indicate the percentage of 100 replicate trees supporting that branch. Bootstrap values below 50% were omitted. Names of clones retrieved in this study begin with 168 1 (Washington Margin), 168 2 (Barrow Canyon Shallow), 16$ 3 (West of Juan de Fuca), 16$ 4 (Shallow Budd Inlet), and EH 168 (East Hanna Shoal Shallow). The scale bar represents 0.05 substitutions per sequence position. sed., sediment; cl., clone; Atl., Atlantic; hydroth., hydrothermal; Massach., Massachussetts; benzene-degr. nitrate-red. cons., benzene-degrading nitrate-reducing consortium. 176 _. 100 100.; saltrriarshssedB.E509 cl. LCP- 26 (AF286031) _ _ 1_Caldrthri'x abyssi BéAJ430587) 86751—1665 328 8A02 . 16$ 16 604 99861 ———---L16326A04 1L- ~ -— L163 26 001 168 40 D11 76 f— ‘99 Japan Trench cold- -seep sed. cl JT 8250 (A8015264) -— —1 1618 H09 HEH1681 D02 ___1_09r CaEsEfideia Mar%in hydrate ridge sed. d. Hyd89—65 (N535255) 511 'L——— 16828 509 -—L~--~- 16318812 L100 LLL - — L1632AH03 L_____Cfl1;topha a fermentans (M 5876 62 7 _J __ - __ 168 riniBiaF i/ia saimonioo or (M62 22) ,, V1 ”h, ___1 " “ “ 163 4C F11 100 "'— 163 2A F11 " _163 2AA08 , 163 3A H07 1 __ 1002163 2A 306 ‘11 ‘ n _‘LL1682AA07 1 V___ MId-Atl. 2ARidge hydroth. vent sed. cl. pIR38H02 (AY354148) 1001 _ JapanS Trench deep—sea sed. cl. NB1-m (A8013834) 100; 5C 168 38 804 168 28 DOS 95:,A 168 18 G10 _ [ -‘ "' ‘-‘—" Arctic sea ice isolate 118 (AF542202) 168 2A F 03 Flavobacterium tripods AJ557887) Flavpbacfgrsium agua rle ( 6279) L V1 168 4C D09 Massach. estuarine sed. cl. C319a- RBC- E2 (AY678530) 10011 ,. 163 ZAP L L L L L1608 28 H11 168 4B 810 LLLL168 4C DO3 __ LL 16S 48 C01 Tenacibaculum maritimum (M64629) P Sf:hiLBoscE)rpens bartonensis (U62913) 7956 Air L EH16S1 F02 006616 1_' __ 168 4C H05 L' __ 2 ..-._ __.. "_‘—— ‘L‘ ‘ Chlorobiumlimioola(Y10640) ,2 , L168 28 E07 100 L North Sea sed. cl. Belgim2005/10-120-13 (DQ351751) L .. LLL L 168 3A 606 L i f“. 100 .mw, 16S 18 001 1 . . "L- -__ LLLLL Sgirochaeta ba/am/rfornrensr‘s (N698859) L L H1681 GO4 Spiroch aeta halophi/a (M88722) 48 809 East Mediterranean mud volcano cl MilanoWF1 8- 06 (AY592849) 168 2A 609 L EH1681 H03 1681 18 E07 ..______. 68 3AC03 '61EH186?:1) 51%8 Acrdobacterr’um capsu/atum (026171) 168 3A E0 L L 1681 AF05 ‘ ‘ 1 16318 H01 1 L -L L LL- 16838005 005 99? ,1001 LL 168 1A H05 ‘L __hlan_ka_i T1rgtégh8d§egsea sed cl NK818 (A8013270) “L ‘“‘ 16$ 38 G12 ,- r163 3A (308 East Mediterranean Sea sed cl. S lonian- F04 (AY533909) Thermus theWnophrlus (x0799 98 ~—- Thermus aquaticus (L09 63) 177 Figure 4.22. Phylogenetic tree showing the affiliation of 16S rRNA clone sequences from the Chloroflaxi, Actinobacteria, Nitraspira, ODl group and OH] group retrieved from Shallow Budd Inlet (SB), Washington Margin (WM), West of Juan de Fuca (WJF), Barrow Canyon Shallow (BC), and East Hanna Shoal Shallow (EHS) sediments to selected reference sequences. The tree was generated by neighbor-joining analysis of approximately 550 bp with T hermus thermophilus and T hermus aquaticus as outgroup. Values at branch points indicate the percentage of 100 replicate trees supporting that branch. Bootstrap values below 50% were omitted. Names of clones retrieved in this study begin with 168 1 (Washington Margin), 168 2 (Barrow Canyon Shallow), 168 3 (West of Juan de Fuca), 16$ 4 (Shallow Budd Inlet), and EH 168 (East Hanna Shoal Shallow). The scale bar represents 0.05 substitutions per sequence position. sed., sediment; cl., clone; Pac., Pacific; Atl., Atlantic; hydroth., hydrothermal; Mich, Michigan; aq., aquatic; microb., microbial. 178 _, ' LL EH16S1 DO4 L - LL marinesed cl BMS43 (AY193161) 1681 18005 _ LLLLL EH1681 806 L 168 38 806 168 18 F07 _ LLLL—L Mich aquifersed. d. WCHB1- 26 AF050599) ___ ___- Mich. aquifer sed cl. WCH81- 6(AF050605) 168 18A06 168 1B F10 Yellowstone hot spring cl. 0P892 (AF027030) - ~ --—- - LL-L1683AG12 __ .L 1682A008 --L16818 D12 86 ,L ,W- 1 - Mid- Pac. deep-sea sed cl. M8AE49(AJ567599) .L—— 168 38 H09 70 L 168 3A F09 LEH1682 803 1 1 lL1681AEO7 6, EH1681 808 Mid- Atl Ridge hydroth ventsed d lR3BDOS AY354150) 99‘ _ . Guaymas Basin hydroth vent sed a2b024 F419678) 1 r hydrocarbonse 8PC063(AF1540 3) 1g dee -sea sed cl 8 2 e1% A(8015539) 97 71—LJE 1681F04 55‘ 1681AC02 168 38 C10 L L 168 3A H01 ' _ LLL-16S4CHO4 7 iL—LLLL—LLL 168 3AA097 “‘ wig—.4 L LL168188 100.. L LL LL Mid- Atl9 Ridge hydroth. ventsed d AT-sZ- 33 (AY225655) EH1681 1605 Cand Microrhnxéiarwoella (X82546) __,_ Ac/djmicrobium 1ferrooxidans (U75647) , _LLL1681ADO3 - 199: fltralia N1u6llgrggrgiaes aq microb formation cl. wb1 _P06 (AF 317769) ___ 1L Mid- -81? Rid%%2 hydroth vent sed. d le3BCOB (AY354165) 1L 100' L 556—16816F09 168 4C F 12 1 065,12— 1168 A10 2 1 10° 90... ”Neogthsw Sea04 sed cl. Belgica2005/10- ZG 15 (DQ351808) Nitraspira04 man'na (X82559) LLLL L L LLLLLLL Nitrospira mOSOOVIGNSlS (X82558) l j FL- L- L16S1AF02 ._ , . 100L 165 3AH03 T , LL East Mediterranean Sea sedol as Ionian A11 (AY534050) rL _ LL LLL 16 $8 agga/ocoocoides eihenogenes (AF 004928) 100 LLLPCB dtachlorinating enrichment culture d OTU-1O (AY559073) 3__1L 168 48 D1 12 ~—-—; - ~* “’0' 168 46 (309 168 4C A12 _JflL 168 48 DO7 1_—*-- 168 48 H06 j 168 aCEugymas Basin hydroth. vent sed d. C1 8046(AF419699) 94f 1 98—1 L Southggrgga tidal sed d 881 -()-115 (AY304372) 1 4 LA LL16S4BAO4 . , LLL LLLL LLL L 168 4B COS 93: L--LLL---16S 4CA06 : 95731- 16818 003 10096 L 168 18 E04 1 . LGuaymas Basinh droth vent sed d. 804RO15 (AY197416) 1_4 571—— L—-1__'hy1droc 4seFep0 sedd PC110(AF154084) 16$2ADO3 - --—~-L: 165 1A F12 ~ - 16S 48 COB 60; a ,- ~~— . ‘09. EH1681 F06 T {LLLLLLL16 60184CC L East Mediterranean mud volcano cl Milano-WF1B-48 (AY592889) 1 ___- _100 168 4C A02 1 _f L East Mediterranean mud volcano cl. Amsterdam -28- 06 (AY592366) ;f_ _ L168 48 804 116$ 18 E03 00 V1 168 18 D06 168 18 E12 22,2 LThermus thermophilus (X07 998 100L Thermus aquaticus (L0966 ) 179 1991-2 168 3A F12 9,9; 1. LL L- ----1--——- EH1681A04 1 ;______.___ 168 18 805 1 1 , 1 ' LLL L‘ _- LLL Propionigenium man's (X84049) 1 601-"LL —-- L 16848607 ;fl__74 _._ LL—LLLL L 168 3A 303 . 1 EH1682 004 1 881 L L LL Bacillus Iitora/is (AY608605) i I __ __ 'LLL Clostn'dium bowmanii(AJ506120) 1 A/lisonel/a histaminifonnans (A F548373) 1 ! 1681A 301 i 1 - EH1681HO4 991 deep-sea cold seep cl. C8521 (A8069798) L 168 2A (301 951-- 16838 E09 1 . -- 100 1166 3A C10 1 1 Mid-Atl. Ridge hydroth. vent sed. cl. AT-sso (AY225657) 168 48 A06 1 71LLLLLL L __ ‘L Fibrobacter suocinogenes (M62696) 621 L- L- 168 2A 005 1 L Thermus thermophi/us (X07998) Thermus aqua ticus (L09663) Figure 4.23. Phylogenetic tree showing the affiliation of 168 rRNA clone sequences from the F irmicutes, F usobacteria, and F ibrobacteres retrieved from Shallow Budd Inlet (SB), Washington Margin (WM), West of Juan de Fuca (WJF), Barrow Canyon Shallow (BC), and East Hanna Shoal Shallow (EHS) sediments to selected reference sequences. The tree was generated by neighbor-joining analysis of approximately 550 hp with T hermus thermophilus and T hermus aquaticus as outgroup. Values at branch points indicate the percentage of 100 replicate trees supporting that branch. Bootstrap values below 50% were omitted. Names of clones retrieved in this study begin with 16$ 1 (Washington Margin), 16$ 2 (Barrow Canyon Shallow), 16$ 3 (West of Juan de Fuca), 16$ 4 (Shallow Budd Inlet), and EH 16S (East Hanna Shoal Shallow). The scale bar represents 0.05 substitutions per sequence position. sed., sediment; cl., clone; Atl., Atlantic; hydroth., hydrothermal. 180 Figure 4.24. Phylogenetic tree showing the affiliation of 168 rRNA clone sequences from the Planctomycetes, Verrucamicrabia and WS3 group retrieved from Shallow Budd Inlet (SB), Washington Margin (WM), West of Juan de Fuca (WJF), Barrow Canyon Shallow (BC), and East Hanna Shoal Shallow (EHS) sediments to selected reference sequences. ‘ The tree was generated by neighbor-joining analysis of approximately 550 hp with Thermus thermophilus and Thermus aquaticus as outgroup. Values at branch points indicate the percentage of 100 replicate trees supporting that branch. Bootstrap values below 50% were omitted. Names of clones retrieved in this study begin with 168 1 (Washington Margin), 16$ 2 (Barrow Canyon Shallow), 168 3 (West of Juan de Fuca), 168 4 (Shallow Budd Inlet), and EH 168 (East Hanna Shoal Shallow). The scale bar represents 0.05 substitutions per sequence position. sed., sediment; cl., clone; hydroth., hydrothermal; G. of Mex., Gulf of Mexico. 181 00413.1 EH1681 C07 [I ———fi , ool ___ EH1682 E06 "I Greenland lkka Fjord ikaite column cl. ikaite un-ch (N431347) _- ‘6‘){163 BB 012 :——_— “’5 ”533353 H1 1 LJ WWW168 3B H04 62L? W 16S1AA07 ‘ W Arctic sed. ct. Sva0800 (AJ297459) J""_——‘ --1681A(302 WW 168 28 DO7 I :WrW EH16S1 A02 L _ . W16S1AG07 ‘W Black Sea water column d. JK471 (00368274) " , 168 38 801 99 ___ Blastopinellula man'na (X62912) l WW' ' “ "WWW Pirellula staleyi (AJ231183) ‘ 168 1B 801 I 100I __7_ r— EH1682A03 100 L G. of Mex. sed. cl. GoM GB425 028—7 (AY542549) l I Gemmata obscuriglobus (AJ231191) 73 _f' , ’ 168 3AD10 «I 66 .. ‘95 Le? ——-—-—— Planctomyces brasiliensis (AJ231190) 85 ‘ “ ' Planctomyces man's(AJ231184) _100; ' W EH1681 GlO "WW ‘ —WW 168 38 C08 ‘ g," 168 3B F02 J 'WWQB ,_ I _II 16828 F11 100 168 4BE12 "“"M ‘ W "““1683ABO4 [ 75‘W ,, __ --WW16S4C F06 W "W" ‘W " ‘ ‘ 168 BABOG 76' -- ~——A~ EH16S1E06 mi . ——~' - 168 1A 808 J i-___ ____ 16S 28 006 88?: 100 Scotland anoxic marine sed. cl. LD1-PB2 (AY114333) L—«fi‘ 168 1A 803 I V I :“WWW— 168 2A 002 {g 100 ---~w~—H ~--——~ 16$ 28 (310 I | W “W —- -— 168 2A E12 " _ _Ioor- ~-— 168 38 I303 ' [___ #91 ~ --— North Sea sed cl. BeIgica2005/10-130 26 (DQ351768) r Opitutus terrae (AJ229235) , 62 .W 163 2A E09 ; * r ‘ L ----- W" W Verrucamicrabium spinosum (X90515) f 95.’ ,.__1QQ.I” 1631B A12 I l: ‘- Barents Sea mud volcano cl. HMMVPog-57 (AJ704715) *W- 168 28 608 ~ W 1 -- - "WW EH16$2 H05 I 168 1A C01 50: ,- {A _ II— 168 2A 804 100 t _ EH1682 B10 , f ’ inactive deep- sea hydroth ventdwimney cl lndB1- 5 (A8099995) as EH1682 H01 99[ _; EH1681 DO5 94W EH16S1 F08 ,5; EH1682 CO7 L 168 40 B10 1 [_.fli’ , r EH1681 E01 I I ' L ‘ "W WW WScotland anoxic marine sed. cl. LD1-PA13 (AY114311) : ! I “'W ‘WWWW‘W" W soil clone PBS-Ill-9 (AJ390450) ‘ 771 r— ' soil cione PBS-ll-5 (N390441) J #4 I 1QrW EH1682 A09 I 80L? L— 16828 CO? + 90! __-’—1BS4CG12 ‘ 100 W W East Mediterranean Sea sed. d. Therm01—68 (AY534052) _ " WW' .,_ “ Scotland anoxic marine sed. cl. LD1-PB19 (AY114332) ml ___ Thermus thermophilus (X07998) 100 " * Thermus aquaticus (L09663) I-— ~-————i 005 182 70—J 1682A F10 57? W 16828 F08 66 ‘163 2A C06 51' [_[r— 1682AGl1 1386 “W 16828 802 l’163 28 (307 168 28 G12 168 2AA02 W 168 28 (311 168 2A H02 '16SZADO4 55 i EH1682 008 168 2A F04 [ 1-9—3W 16S 28 012 {LBJ ’ EH1681 A11 .W 168 4C (304 W- Salt marsh cl. SIMO—823 (AY712360) 97‘ I", T“ 168 2A GOB l 73: Shepherds Creek sub-watershed d. 182up (AY212634) -- WW 168 28 H07 W~ - --W- W EH16S1 F03 I i ' I ' —_ Dermocarpe/Ia incrassata (AJ344559) l l ‘ sat-W W- Anabaena cylindrica(AF091150) __87 Pro chlorococcus marin us (AF180967) __ _ _} 7hermusthennophilus (X07998) 100‘ ___ Thermus aquaficus(L09663) Figure 4.25. Phylogenetic tree showing the affiliation of 168 rRNA clone sequences from the Cyanobacteria retrieved from Shallow Budd Inlet (SB), Washington- Margin (WM), West of Juan de Fuca (WJF), Barrow Canyon Shallow (BC), and East Hanna Shoal Shallow (EHS) sediments to selected reference sequences. The tree was generated by neighbor-joining analysis of approximately 550 bp with Thermus thermophilus and T hermus aquaticus as outgroup. Values at branch points indicate the percentage of 100 replicate trees supporting that branch. Bootstrap values below 50% were omitted. Names of clones retrieved in this study begin with 168 1 (Washington Margin), 16$ 2 (Barrow Canyon Shallow), 16$ 3 (West of Juan de Fuca), 168 4 (Shallow Budd Inlet), and EH 168 (East Hanna Shoal Shallow). The seale bar represents 0.05 substitutions per sequence position. cl., clone. 183 Table 4.5. Assignment of 16S rRNA clone sequences from Pacific Northwest and Arctic sediments into taxonomic groups. Taxonomic groupa Stationb BC EHS SB WM WJF Proteabacteria 30 42 35 30 25 Alpha Proteobacteria 1 1 4 4 3 Beta Proteabacteria 0 O O O l Gamma Proteabacteria 9 2O 13 l 1 14 Delta Proteobacteria 10 19 10 11 ‘ 3 Epsilon Proteabacteria 2 O 5 0 0 unclassified Proteabacteria 8 2 3 4 4 Bacteroidetes l4 4 12 4 3 F lavobacteria 4 1 7 0 O Sphingobacteria 2 0 O 2 2 Bacteroidetes l O O O O unclassified ' Bacteroidetes 7 3 5 2 1 Planctomycetes 1 4 0 4 5 Planctomycetacia 1 4 O 4 5 F irmicutes 1 1 0 l 0 C Iostridia 1 l O O 0 Caldithrix O 0 0 1 0 C yanobacteria 4 3 1 0 0 Cyanobacteria 4 3 1 O ' O Verrucamicrabia 3 3 2 l 0 Verrucomicrobiae 3 3 2 1 0 Nitraspira 0 0 0 l l Nitraspira O 0 0 1 1 Actinobacteria 0 0 1 l 2 Actinobacteria O O l 1 2 Genera_incertae_sedis_WS3 l l 1 0 0 Genera_incertae_sedis_OPll 0 0 0 0 ' l Unclassified Bacteria 27 25 28 38 38 aAssignment to taxonomic groups as determined by the placement of the sequences into the RDP Hierarchy with the RDP Classifier. Bootstrapping was performed to estimate the classification reliability. Sequences that could not be assigned to a taxon with a bootstrap confidence estimate above the 80% selected threshold were considered “unclassified”. bBC, Barrow Canyon Shallow; EHS, East Hanna Shoal Shallow; SB, Shallow Budd Inlet; WM, Washington Margin; WJF, West of Juan de Fuca. 184 Table 4.6. Significant differences in the taxonomic composition between 16S rRNA libraries from Pacific Northwest and Arctic sediments. Libraries comparedal Cloneiaisslllgn ed to Taxonomic group . . Significancec Library 1 Library 2 L‘blra’y L‘bgary BC EHS Bacteroidetes 14 4 0.017 Gamma Proteabacteria 9 20 0.029 BC SB Sulfurospirillum 0 5 0.030 BC WM Bacteroidetes 14 4 0.02 1 BC WJ F Bacteroidetes 14 3 0.012 EHS SB Epsilon Proteabacteria 0 5 0.028 Campylobacterales 0 5 0.028 C ampylobacteraceae 0 5 0.028 Sulfurospirillum 0 5 0.028 Bacteroidetes 4 1 2 0.04 l EHS WM Desulfobulbaceae 8 0 0.005 Proteabacteria 42 28 0.044 EHS W] F - Delta Proteabacteria 20 3 0.001 Desulfobulbaceae 8 0 0.006 Desulfobacterales 1 2 2 0.01 3 Proteabacteria 42 24 0.018 SB WM F lavobacteria 7 0 0.008 F lavobacteriales 7 0 0.008 F Iavobacteriaceae 7 0 0.008 Desulfobulbaceae 6 0 0.01 6 Desulfobacterales 1 0 2 0.023 Epsilon Proteabacteria 5 0 0.031 C ampylobacterales 5 0 0.03 l C ampylobacteraceae 5 0 0.03 1 Sulfurospirillum 5 0 0.03 1 Bacteroidetes 1 2 4 0.049 SB WJ F F lavobacteria 7 0 0.010 185 Table 4.6 (cont’d) Significant differences in the taxonomic composition between 16S rRNA libraries from Pacific Northwest and Arctic sediments. Clones assigned to Libraries compareda Taxonomic group taxon Significancec Library 1 Library 2 Library 1 Library 2 F lavobacteriales 7 0 0.010 F Iavobacteriaceae 7 0 0.01 0 Desulfobulbaceae 6 0 0.020 Planctomycetes 0 5 0.026 Planctomycetacia 0 5 0.026 Planctomycetales 0 5 0.026 Planctomycetaceae 0 5 0.026 Bacteroidetes 1 2 3 0.030 Desulfobacterales 1 0 2 0.030 Epsilon Proteobacteria 5 O 0.038 C ampylobacterales 5 0 0.038 Campylobacteraceae 5 0 0.038 Sulfurospirillum 5 0 0.038 WM WJF Delta Proteobacteria 12 3 0.030 aBC, Barrow Canyon Shallow; EHS, East Hanna Shoal; SB, Shallow Budd Inlet; WM, Washington Margin; WJF, West of Juan de Fuca. bAssignment to taxonomic groups as determined by the placement of the sequences into the RDP Hierarchy with the RDP Classifier, using a threshold value of 80% for the bootstrap confidence estimate. cOnly taxonomic groups presenting significant differences (P S 0.05) between the compared pair of clone libraries were included. 186 70 168 rRNA (11 O | I 1 1 1 1 | .h 0 1 (a) O l I No. of OTUs "J ‘1' “I ".' '5" '1' “I "I "I "f "f "f "f ".’ "f "I "I"! ".' 't' f‘f'f'f‘f' 7O ,. - ___ __ ____ __ . ___. . -_ ..._ ”In _.__HLMLJ 50 n ., __ ._._ __ __ .__..___.__ .___ L____-_.._2__‘_. No. of OTUs 13 5 7 91113 15 1719 2123 25 27 Abundance EBCIEHS EISB IWM IWJF Figure 4.26. Relationship between the number of OTUs and their abundance, as measured by the number of clones corresponding to each OTU, for 168 rRNA and nirS clone libraries from Barrow Canyon Shallow (BC), East Hanna Shoal Shallow (EHS), Shallow Budd Inlet (SB), Washington Margin (WM), and West of Juan de Fuca (WJF) sediments. 187 so LL#,W,, “77 rem—em , , «— 1 70 183 rRNA W 60 . Egg;- ‘ @1111 i i 30 - ggggfig ‘ 20 Number of OTUs o g 1 ‘ 1 I 1 1 0 10 20 30 40 5O 60 70 80 90 Number of clones analyzed WnirS 70* 60* 50‘ 40- Number of OTUs 20 WW 10 W 0 25 50 75 100 125 150 175 Number of clones analyzed EBC_'-giE_I?. SB WW Figure 4.27. Rarefaction curves for 16S rRNA and nirS clone libraries from Barrow Canyon Shallow (BC), East Hanna Shoal Shallow (EHS), Shallow Budd Inlet (SB), Washington Margin (WM), and West of Juan de Fuca (WJF) sediments. Error bars represent 95% confidence intervals. 188 Table 4.7 . Diversity and predicted richness of 16S rRNA and nirS gene fragments in sediments from different sites based on OTU assignment of clones by DOTUR. Gene and No. of No. of Shannon Index of Chaol richness stationa clones OTUSb diversityc estimatec 16S rRNA BC 81 65 3.99 (3.78; 4.20) 493 (232; 1160) EHS 83 62 3.99 (3.81; 4.17) 215 (129; 414) SB 80 62 4.01 (3.83; 4.19) 221 (132; 427) WM 80 70 4.18 (4.01; 4.35) 574 (269; 1348) WJF 75 62 4.02 (3.83; 4.21) 266 (150; 537) nirS BC 153 60 3.55 (3.36; 3.74) 112 (82; 183) EHS 124 53 3.52 (3.33; 3.72) 98 (71; 164) SB 137 42 3.08 (2.87; 3.29) 93 (59; 192) WM 161 87 3.94 (3.74; 4.14) 348 (209; 643) WJF 74 30 3.07 (2.85; 3.30) 41 (33; 71) aBC, Barrow Canyon Shallow; EHS, East Hanna Shoal Shallow; SB, Shallow Budd Inlet; WM, Washington Margin; WJF, West of Juan de Fuca. bDistance threshold applied for OTU assignment was 3% and 5% for 168 rRNA and nirS, respectively. cLower und upper 95% confidence intervals are indicated in parenthesis. Table 4.8. Determination of significant differences between libraries by the application of the integral form of the Cramér—von Mises statistic with 1- LIBSHUFF. Sampling site P value for comparison of heterologous with Gene homgldgous homologous libraryb “bran?" BC EHS SB WM WJF 1 6S rRNA BC 0.0036 0.0104 0.0000 0.0000 EHS 0.0000 0.0000 0.0003 0.0000 SB 0.0000 0.0000 0.0000 0.0000 WM 0.0000 0.3476 0.0012 0.0000 W] F 0.0000 0.0003 0.0000 0.0002 nirS BC 0.0009 0.0000 0.0000 0.0000 EHS 0.01 75 0.0000 0.0000 0.0000 SB 0.0000 0.0000 0.0000 0.0000 WM 0.0000 0.0000 0.0000 0.0000 WJ F 0.0000 0.0000 0.0000 0.0000 aBC, Barrow Canyon Shallow; EHS, East Hanna Shoal Shallow; SB, Shallow Budd lnlet; WM, Washington Margin; W] F, West of Juan de F uca. bSignificant P values (P < 0.0026; (1 = 0.05) are indicated in bold. Each pair of libraries is compared twice, by switching the library used as homologous and heterologous for the second comparison. Libraries are significantly different from each other if at least one of the two comparisons between them leads to a significant P value. 190 Table 4.9. Similarity between libraries described by the abundance-based Sorenson similarity index (Labund) and with a nonparametric richness estimator of shared OTUs analogous to Chaol (SM; cm) as determined with SONS. Libraries compareda 16S rRNA Labund SA,B Chao Labund SA,B Chao EHS WM 0.239 45 0.494 28 EHS BC 0.194 14 0.676 47 EHS WJF 0.020 2 0 0 EHS SB 0.049 3 0.011 _ 2 WM BC 0.083 13 0.158 10 WM WJF 0 0 0.011 2 WM SB 0.044 5 0.012 2 BC WJF 0.019 2 0 0 BC SB 0 0 0 0 WJF SB 0 0 0 i 0 aEHS, East Hanna Shoal Shallow; BC, Barrow Canyon Shallow, WM, Washington Margin; WJF, West of Juan de Fuca; SB, Shallow Budd Inlet. 191 Discussion In previous studies (Chapter 3) a strong conservation of denitrifier community structure was observed at different sediment depths along redox gradients within the bioturbated layer. Clear community differences were detected between sites within one geographic location (Puget Sound) and increased when sites from two different geographic locations were compared (Puget Sound and Washington Margin). This pattern was also observed in the same area by Braker et al. (2000; 2001) when various sites with differing water depths at the Washington Margin were compared to a Puget Sound site, suggesting that geographic location, water depth, and sediment depth were involved in the appearance of community differences with decreasing importance in that order. To further describe this phenomenon, the denitrifier, as well as the bacterial community diversity and distribution were studied in two distant geographic areas, the Pacific Northwest and the Arctic Ocean. The former included the previously studied locations in Puget Sound and the continental slope at the Washington Margin (Chapter 3), in addition to an abyssal sea floor site separated from the rest by a geographic barrier. The Arctic sites included high and low productivity locations with various water depths. All studied sites showed a high bacterial and denitrifier diversity. The former was especially high at the continental slope of the Washington Margin, based on 16S rRNA clone libraries (Table 4.7). This site also presented, as well as the Arctic, a high denitrifier diversity, based on nirS gene clone libraries and T-RFLP (Table 4.7, Figure 4.6B). The abyssal sea floor (West of Juan de Fuca) presented the highest and lowest nirS diversity, based on T-RFLP and clone library results, respectively. This incongruence might be due to the different resolution and coverage levels of both techniques. 192 Furthermore, although the T-RF richness at this site was lower than in other areas (Figure 4.6A), the even distribution of T-RFs probably led to an increase of the measured diversity. Surprisingly, Puget Sound samples (Shallow Budd Inlet, Turning Basin, and Carr Inlet) consistently showed a lower denitrifier diversity (Table 4.7, Figure 4.68) than the other sites, which is opposite the results obtained by Braker et al. (2000; 2001) who detected a higher diversity in Puget Sound compared to the Washington margin area. Samples analyzed in the two studies were, however, not retrieved from the same sites, so local differences might be responsible for the observed pattern. The low diversity determined in Puget Sound samples, similar to the one observed at West of Juan de Fuca, might be related to the fact that these sites had the highest and lowest productivity of the sampled areas, respectively. It has been observed that the richness of certain bacterial taxonomic groups was affected by primary productivity, leading in some cases to higher richness at intermediate productivity levels (Homer-Devine et al., 2003), as often observed for plants and animals (Mittelbach et al., 2001). Denitrifiers, although not a taxonomic group, might, as a functional group behave in a similar manner and present higher richness and diversity at intermediate productivity levels. In addition, the same pattern was observed also on a vertical scale, as maximum richness and diversity of T- RF 5 were generally measured in subsurface sediments, rather than in the top layer (0-0.5 cm depth), which receives the highest amount of degradable organic matter from the overlying water and lies within the depth at which oxygen and nitrate are detected in the pore water. This same phenomenon was also observed in the Mediterranean Sea, where the highest bacterial richness, as studied by T-RFLP, was detected at 1-2 and 2-3 cm deep sediments and not at the top layer (Luna et al., 2004). 193 As presented previously in Chapter 3, site-specific grouping of the nirS T-RFLP profiles was observed by cluster analysis (Figure 4.7) and PCA ordination (Figure 4.8), except for deep sediments, indicating that sediment depth within the bioturbated zone did not play a major role in differentiating denitrifier communities. Water depth along a transect or slightly differing biogeochemical characteristics within one geographic location had a stronger effect, while the most differentiated communities were from different geographic areas. This is consistent with observations by Braker et al. (2001) on nirS and 168 rRNA distributions in sediments from Washington Margin and Puget Sound. The distribution of another denitrification gene, nosZ, showed the same pattern, with sediment depth, meter scale and kilometer scale horizontal distances having increasing influence on the differentiation of the community (Scala and Kerkhof, 2000). Differences in sulfate-reducing bacterial communities were also detected. at the Washington Margin related to water depth, which has been associated with decreasing carbon quantity and bioavailability at increasing depth (Liu et al., 2003). Water depth and organic carbon were also identified by CCA as main environmental factors influencing the denitrifier community structure with opposite effects, in addition to geographic position (latitude and longitude), oxygen, nitrate and ammonium (Figure 4.9). Nitrate and oxygen have both been suggested as key factors affecting the sediment (Liu et al., 2003) and water column (Castro-Gonzalez et al., 2005) denitrifier community structure. In the Baltic Sea water column ammonium was also identified in addition to oxygen and nitrate for its role in shaping the denitrifier community (Hannig et al., 2005). Interestingly, although higher salinity has been related with lower denitrifier diversity in other studies (Santoro et al., 2006), it didn’t seem to 194 have a major role in this study. Furthermore, the lowest salinity area (Puget. Sound) presented the lowest richness and diversity. Clone libraries from all sites, except the abyssal sea floor (West of Juan de Fuca), suggested a dominance of Gamma and Delta Proteabacteria (Table 4.5), as has been observed in Arctic (Ravenschlag et al., 1999) and Antarctic (Bowman and McCuaig, 2003) continental shelf sediments. The abyssal sea floor, as expected for deep sea sediments which generally show a lower sulfate reduction rate than shallow sediments (Hartnett and Devol, 2003), presented a high abundance of Gamma Proteabacteria, but significantly lower abundance of Delta Proteabacteria than other sites. The high abundance of Delta Proteabacteria in Puget Sound and Washington Margin sediments is consistent as well with the detection of a broad range of dissimilatory (bi)sulfite reductase gene sequences (dsrAB) in the former (Tiquia et al., 2006) and latter site (Liu et al., 2003), respectively. Furthermore, sulfate reduction was reported to be responsible for an average of 50% of the total carbon oxidation rate in Washington shelf and upper slope sediments (Hartnett and Devol, 2003). The widely distributed Bacteroidetes, were also detected at all sites and were particularly abundant in Barrow Canyon Shallow and Puget Sound sediments. This phylum had been identified as a major contributor. to the community incorporation of tritium-labelled thymidine in Puget Sound water column (Van Mooy et al., 2004), suggesting an important role in this coastal environment. Bacteroidetes seem to be highly adapted to metabolize high molecular weight compounds (Kirchman, 2002), which are probably highly abundant in the sediments of the highly productive and close to shore Puget Sound and Barrow Canyon Shallow sites. Planctomycetes were detected at all sites, except at Shallow Budd Inlet in Puget Sound, 195 though none of the retrieved sequences were similar to anammox-related bacteria. These have been, however, detected with specific primers targeting this group in Shallow Budd Inlet, Turning Basin, Washington Margin, West of Juan de Fuca, East Hanna Shoal Shallow and Deep sediments (Penton et al., 2006). This difference is expected when general bacterial vs. group-specific primers are used and might also be related to the low coverage of the 16S rRNA clone libraries in this study of the total bacterial diversity present at these sites. The clone libraries of nirS as well as 16S rRNA genes were significantly different between the studied locations (Table 4.8), suggesting that the nirS-containing denitrifier populations as well as the general bacterial community were significantly different at each site. This is in agreement with the detection of significantly different denitrifier communities even along a 40 m nitrate and salinity gradient in a beach aquifer (Santoro, 2006), suggesting that denitrifier communities are highly site-specific and vary at relatively small spatial scales. The Arctic (Barrow Canyon Shallow and East Hanna Shoal Shallow) nirS clone libraries, though significantly different, presented the highest similarity and estimated number of shared OTUS between these two sites (Table 4.9). Although on different transects with differing productivity levels, these two sites share several biogeochemical characteristis and are geographically closer together than the other sites, which could favor the development of similar denitrifier communities. Interestingly, the Washington Margin nirS library was more similar to the Arctic libraries than to the other Pacific Northwest libraries. Despite the larger geographic distance, biogeochemical similarities detected between these two continental margin environments (Devol et al., 2005) might influence the development of more similar communities. 196 Castro-Gonzalez et al. (2005) also observed a closer, though still distant, clustering of water column nirS sequences from the oxygen minimum zone in the eastern South Pacific with Arabian Sea oxygen minimum zone water column clones, as with sediment clones from other Pacific Ocean areas. Clustering on the nirS tree further supported the significant differences observed between libraries. Clusters of highly similar sequences generally included clones retrieved at only one location, however, several clusters including sequences from both Arctic locations, as well as clusters including Arctic and Washington Margin clones, were also observed, consistent with the similarities detected between those libraries. Shallow Budd Inlet and West of Juan de Fuca nirS clones generally formed individual clusters, consistent with their low similarity to other libraries. Despite individual site- specific clustering of highly similar clones, no general separation of clones from different locations was observed. Furthermore, most clones (95.1%) were more than 80% similar to previously retrieved environmental clones from various geographic locations, indicating an increasing coverage of nirS diversity by growing sequence databases. Interestingly, the Shallow Budd Inlet nirS clusters previously identified and quantified by real-time PCR (Chapter 3) did not include any sequences from other areas, consistent with the lower or lack of detection of most of them at the other locations (Table 4.3). This might suggest a specific adaptation to this environment. Decrease in number of specific nirS gene and transcript phylotypes was also observed along the Colne estuary from the head to the mouth (Smith et al., 2007). P. stutzeri nirS, on the other hand, was detected at all studied locations, although it represented at most 0.66% of the community in Washington Margin sediments and 1 to 2 orders of magnitude less at 197 the other sites. This is consistent with the frequent detection of this bacterium in marine samples, although at varying abundances (Ward and Cockcroft, 1993; Griintzig et al., 2001) Shallow Budd Inlet and West of Juan de Fuca were also the most divergent sites based on their bacterial community, while the Arctic and Washington Margin, presented the most similar bacterial communities, coincident with the pattern observed for nirS. However, the highest similarity was detected between East Hanna Shoal Shallow and Washington Margin, rather than between the two Arctic sites, as observed for nirS. This further supports a probable similarity in environmental conditions favoring the development of similar communities at these two sites. The lower evolutionary rate of the 16S rRNA gene compared to nirS could be responsible for the establishment of more similar communities regardless of the larger geographic distances. This pattern was also observed in the 16S rRNA phylogenetic tree, with clusters of highly similar sequences generally including clones from only one location, but no specific clustering of clones from close geographic locations beyond that. Regretfully, it is not possible to draw a clear correlation between 168 rRNA and nirS phylogenies, as strains with the same 16S rRNA phylotype have been often observed to have notable differences in nirS phylotype (Goregues et al., 2005). In conclusion, the differentiation of denitrifier communities, as studied by the nirS gene, seems to be mainly influenced by geographic location, followed by varying environmental factors within one area, and least affected by sediment depth. This can be correlated to the hypothesis proposed by Hughes Martiny et al. (2006) of multiple microbial provinces and habitats, representing, respectively, historical and environmental 198 factors affecting the microbial community structure. Based on this hypothesis, if biotic similarities between samples are correlated with environmental similarity and geographic distance, multiple habitats and provinces can be defined. On the other hand, for 16S rRNA, due to its lower mutation rate, biotic similarities were less correlated to geographic distance, therefore, provinces seem less differentiated and tend to form one province with multiple habitats. 199 References Armstrong, F. A. J., C. R. Stearns, and J. D. H. Strickland. 1967. The measurement of upwelling and subsequent biological processes by means of the Technicon Autoanalyser and associated equipment. Deep-Sea Res. 14:381-389. Bates, N. R., M. H. P. Best, and D. A. Hansel]. 2005. Spatio-temporal distribution of dissolved inorganic carbon and net community production in the Chukchi and Beaufort Seas. Deep-Sea Res. 11 52:3303-3323. Bowman, J. P. and R. D. McCuaig. 2003. Biodiversity, community structural shifts, and biogeography of prokaryotes within Antarctic continental shelf sediment. Appl. Environ. Microbiol. 69:2463-2483. Braker, G., H. A. Ayala-deI-Rio, A. H. Devol, A. Fesefeldt, and J. M. Tiedje. 2001. Community structure of denitrifiers, Bacteria, and Archaea along redox gradients in Pacific Northwest marine sediments by terminal restriction fragment length polymorphism analysis of amplified nitrite reductase (nirS) and 16S rRNA genes. Appl. Environ. Microbiol. 67: 1893—1901 . Braker, G., J. Zhou, L. Wu, A. H. Devol, J. M. Tiedje. 2000. Nitrite reductase genes (nirK and nirS) as functional markers to investigate diversity of denitrifying bacteria in Pacific Northwest marine sediment communities. Appl. Environ. Microbiol. 66: 2096— 2104. Brandes, J. A., and A. H. Devol. 1995. Simultaneous nitrate and oxygen respiration in coastal sediments: Evidence for discrete diagenesis. Journal of Marine Research. 53:771- 797. Brewer, P. G. and D. W. Spencer. 1974. Colorimetric determination of manganese in anoxic waters. Limnology and Oceanography 16:107-110. Castro-Gonzalez, M., G. Braker, L. Farias, and O. Ulloa. 2005. Communities ‘of nirS- type denitrifiers in the water column of the oxygen minimum zone in the eastern South Pacific. Environ. Microbiol. 7: 1298-1306. Cho, J.—C. and J. M. Tiedje. 2000. Biogeography and degree of endemicity of fluorescent Pseudomonas strains in soil. Appl. Environ. Microbiol. 66:5448-5456. Cole, J. R., B. Chai, R. J. Farris, Q. Wang, A. S. Kulam-Syed-Mohideen, D. M. McGarrell, A. M. Bandela, E. Cardenas, G. M. Garrity, and J. M. Tiedje. 2006. The ribosomal database project (RDP-II): introducing myRDP space and quality controlled public data. Nucleic Acids Res. 35:Dl69-D172. Cole, J. R., B. Chai, R. J. Farris, Q. Wang, S. A. Kulam, D. M. McGarrell, G. M. Garrity, and J. M. Tiedje. 2005. The Ribosomal Database Project (RDP 11): sequences and tools for high-throughput rRNA analysis. Nucleic Acids Res. 33:D294-D296. 200 Devol, A., B. Chang, and J. Christensen. 2005. Denitrification and sulfate reduction in Arctic continental margin sediments. Goldschmidt Conference Abstracts. A581. Dhillon, A., A. Teske, J. Dillon, D. A. Stahl, and M. L. Sogin. 2003. Molecular characterization of sulfate-reducing bacteria in the Guaymas Basin. Appl. Environ. Microbiol. 69:2765-2772. Dojka, M. A., P. Hugenholtz, S. K. Haack, and N. R. Pace. 1998. Microbial diversity, in a hydrocarbon- and chlorinated-so[vent-contaminated aquifer undergoing intrinsic bioremediation. Appl. Environ. Microbiol. 64: 3869-3877. Fossing, H. and B. B. Jorgensen. 1989. Measurement of bacterial sulfate-reduction in sediments: evaluation of single-step chromium reduction method. Biogeochemistry 8:205-222. Goregues, C. M., V. D. Michotey, and P. C. Bonin. Molecular, biochemical, and physiological approaches for understanding the ecology of denitrification. Microb. Ecol. 49:198-208. Green, J. and B. J. M. Bohannan. 2006. Spatial scaling of microbial biodiversity. Trends Ecol. Evol. 21:501-507. Griintzig, V., S. C. Nold, J. Zhou, and J. M. Tiedje. 2001. Pseudomonas stutzeri nitrite reductase gene abundance in environmental samples measured by real-time PCR. Appl. Environ. Microbiol. 67:760-768. Hall, T. A. 1999. A user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucl. Acids Symp. Ser. 41:95-98. Hannig, M., G. Braker, J. Dippner, and K. Jiirgens. 2006. Linking denitrifier community structure and prevalent biogeochemical parameters in the pelagial of the central Baltic Proper (Baltic Sea). FEMS Microbiol. Ecol. 57:260-271. Hartnett, H. E. and A. H. Devol. 2003. Role of a strong oxygen-deficient zone in the preservation and degradation of organic matter: A carbon budget for the continental margins of northwest Mexico and Washington State. Geochim. Cosmochim. Acta 67 :247-264. Hedges, J. and J. Stern. 1984. Carbon and nitrogen determinations of carbonate containing solids. Limnology and Oceanography 29:657-663. Heijs, S. K., J. S. Sinninghe-Damste, and L. J. Forney. 2005. Characterization of a deep-sea microbial mat from an active cold seep at the Milano mud volcano in the Eastern Mediterranean Sea. FEMS Microbiol. Ecol. 54:47-56. Hill, T. C. J., K. A. Walsh, J. A Harris, B. F. Moffet. 2003. Using ecological diversity measures with bacterial communities. FEMS Microbiol. Ecol. 43:1-11. 201 Hill, V. and G. Cota. 2005. Spatial patterns of primary production on the shelf, slope and basin of the Western Arctic in 2002. Deep-Sea Res. 52:3344-3354. Horner—Devine, M. C., M. Lage, J. B. Hughes, and B. J. M. Bohannan. 2004. A taxa- area relationship for bacteria. Nature 432:750-753. Horner-Devine, M. C., M. A. Leibold, V. H. Smith, and B. J. M. Bohannan. 2003. Bacterial diversity patterns along a gradient of primary productivity. Ecol. Lett. 6:613- 622. Houghton, J. T., Y. Ding, D. J. Griggs, M. Noguer, P. J. van der Linden, X. Dai, K. Maskell, and C. A. Johnson. 2001. Climate change 2001: the scientific basis. Cambridge University Press, Cambridge. Hugenholtz, P., C. Pitulle, K. L. Hershberger, and N. R. Pace. 1998. Novel division level bacterial diversity in a Yellowstone hot spring. J. Bacteriol. 180:366—376. Hughes Martiny, J. B., B. J. M. Bohannan, J. H. Brown, R. K. Colwell, J. A. Fuhrman, J. L. Green, M. C. Horner-Devine, M. Kane, J. Adams Krumins, C. R. Kuske, P. J. Morin, S. Naeem, L. Ovreas, A.-L. Reysenbach, V. H. Smith, and J. T. Staley. 2006. Microbial biogeography: putting microorganisms on the map. Nature 42102-112. . Ivanova, E. P., N. V. Zhukova, A. M. Lysenko, N. M. Gorshkova, A. F. Sergeev, V. V. Mikhailov, and J. P. Bowman. 2005. Loktanella agm'ta sp. Nov. and Loktanella rosea sp. Nov., from the north-west Pacific Ocean. Int. J. Syst. Evol. Microbiol. 55:2203- 2207. Jayakumar, D. A., C. A. Francis, S. W. A. Naqvi, and B. B. Ward. 2004. Diversity of nitrite reductase genes (nirS) in the denitrifying water column of the coastal Arabian Sea. Aquat. Microb. Ecol. 34:69-78. Jukes, T. H. and C. R. Cantor. 1969. Evolution of protein molecules. In Munro, H. N. (ed.). Mammalian Protein Metabolism. Academic Press, New York. Kirchman, D. L. 2002. The ecology of Cytophaga—Flavobacteria in aquatic environments. FEMS Microbiol. Ecol. 39:91-100. Kumar, S., K. Tamura, I. B. Jakobsen, and M. Nei. 2001. MEGA2: molecular evolutionary genetics analysis Software. Bioinformatics. 17 : 1244-1245. Li, L., C. Kato, and K. Horikoshi. 1999. Microbial diversity in sediments collected from the deepest cold-seep area, the Japan Trench. Mar. Biotechnol. 1:391-400. Liu, X., C. E. Bagwell, L. Wu,A. H. Devol, andJ. Zhou. 2003. Molecular diversity of sulfate-reducing bacteria from two different continental margin habitats. Appl. Environ. Microbiol. 69: 6073- 6081. 202 Liu, X., S. M. Tiquia, G. Holguin, L. Wu, S. C. Nold, A. H. Devol, K. Luo, A. V. Palumbo, J. M. Tiedje, and J. Zhou. 2003. Molecular diversity of denitrifying genes in continental margin sediments within the oxygen-deficient zone off the Pacific coast of Mexico. Appl. Environ. Microbiol. 69:3549-3560. Lopez-Garcia, P., S. Duperron, P. Philippot, J. F oriel, J. Susini, and D. Moreira. 2003. Bacterial diversity in hydrothermal sediment and epsilonproteobacterial dominance in experimental microcolonizers at the Mid-Atlantic Ridge. Environ. Microbioll 5:961- 976. Luna, G. M., A. Dell’Anno, L. Giuliano, and R. Danovaro. 2004. Bacterial diversity in deep Mediterranean sediments: relationship with the active bacterial fiaction and substrate availability. Environ. Microbiol. 6:745-753. Mittelbach, G. G., C. F. Steiner, S. M. Scheiner, K. L. Gross, H. L. Reynolds, R. B. Waide, M. R. Willig, S. I. Dodson, and L. Gough. 2001. What is the observed relationship between richness and productivity? Ecology 82:2381-2396. Mussmann, M., K. Ishii, R. Rabus, and R. Amann. 2005. Diversity and vertical distribution of cultured and uncultured Deltaproteobacteria in an intertidal mud flat of the Wadden Sea. Environ. Microbiol. 7 2405-418. Nei, M. and S. Kumar. 2000. Molecular evolution and phylogenetics. Oxford University Press, New York. Nercessian, O., Y. Fouquet, C. Pierre, D. Prieur, and C. Jeanthon. 2005. Diversity of Bacteria and Archaea associated with a carbonate-rich metalliferous sediment sample from the Rainbow vent field on the Mid-Atlantic Ridge. Environ. Microbiol. 7 :698-714. Nogales, B., K. N. Timmis, D. B. Nedwell and A. M. Osborn. 2002. Detection and diversity of expressed denitrification genes in estuarine sediments after reverse transcription-PCR amplification from mRNA. Appl. Environ. Microbiol. 68:5017-5025. Nold, S. C., J. Zhou, A. H. Devol, and J. M. Tiedje. 2000. Pacific Northwest marine sediments contain ammonia-oxidizing bacteria in the [3 subdivision of the Proteabacteria. Appl. Environ. Microbiol. 66:4532-4535. Penton, C. R., A. H. Devol, and J. M. Tiedje. 2006. Molecular evidence for the broad distribution of anaerobic ammonium-oxidizing bacteria in freshwater and marine sediments. Appl. Environ. Microbiol. 72:6829-6832. Polymenakou, P. N., S. Bertilsson, A. Tselepides, and E. G. Stephanou. 2005. Bacterial community composition in different sediments from the Eastern Mediterranean Sea: a comparison of four 16S ribosomal DNA clone libraries. Microb. Ecol. 50:447-462. Ramette, A. and J. M. Tiedje. 2007. Biogeography: An emerging cornerstone for understanding prokaryotic diversity, ecology, and evolution. Microbiol. Ecol. 53:197- 207. 203 Ramette, A. and J. M. Tiedje. 2007. Multiscale responses of microbial life to spatial distance and environmental heterogeneity in a patchy ecosystem. Proc. Nat. Acad. Sci. 10422761-2766. Ravenschlag, K., K. Sahm, J. Pernthaler, and R. Amann. 1999. High bacterial diversity in permanently cold marine sediments. Appl. Environ. Microbiol. 65:3982- 3989. Saitou, N. and M. Nei. 1987. The neighbour joining method: a new tool for reconstructing phylogenetic trees. Mol. Biol. Evol. 41406-425. Sambrook, J. and D. W. Russell. 2001. Molecular cloning: a laboratory manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. Santoro, A. E., A. B. Boehm and C. A. Francis. 2006. Denitrifier community composition along a nitrate and salinity gradient in a coastal aquifer. Appl. Environ. Microbiol. 72: 2102-2109. Scala, D. J. and L. J. Kerkhof. 1999. Diversity of nitrous oxide reductase (nosZ) genes in continental shelf sediments. Appl. Environ. Microbiol. 65: 1681-1687. Scala, D. J. and L. J. Kerkhof. 2000. Horizontal heterogeneity of denitrifying bacterial communities in marine sediments by terminal restriction fragment length polymorphism analysis. Appl. Environ. Microbiol. 66:1980-1986. Schloss, P. D., B. R. Larget, and J. Handelsman. 2004. Integration of microbial ecology and statistics: a test to compare gene libraries. Appl. Environ. Microbiol. 70:5485-5492. Schloss, P. D. and J. Handelsman. 2005. Introducing DOTUR, a computer program for defining operational taxonomic units and estimating species richness. Appl. Environ. Microbiol. 71:1501-1506. Schloss, P. D. and J. Handelsman. 2006. Introducing SONS, a tool for operational taxonomic unit-based comparisons of microbial community memberships and structures. Appl. Environ. Microbiol. 72:6773-6779. Smith, C. J., D. B. Nedwell, L. F. Dong, and A. M. Osborn. 2007. Diversity and abundance of nitrate reductase genes (narG and napA), nitrite reductase genes (nirS and nrfA), and their transcripts in estuarine sediments. Appl. Environ. Microbiol. 73:3612- 3622. ' Stevens, H., M. Stubner, M. Simon, and T. Brinkhoff. 2005. Phylogeny of Proteobacteria and Bacteroidetes from oxic habitats of a tidal flat ecosystem. FEMS Microbiol. Ecol. 54:351-365. Stookey, L. L. 1970. Ferrozine —a new spectrophotometric reagent for iron. Analytical Chemistry 42:779-781. 204 Strickland, J. D. H. and T. R. Parsons. 1972. A practical handbook of seawater analysis. Fisheries Research Board of Canada, Ottawa, Ontario. Tabatabai, M. A. 1974. A rapid method for determination of sulfate in water samples. Environ. Letters 7 2237-243. Tarnegai, H., R. Aoki, S. Arakawa, and C. Kato. 2007. Molecular analysis of the nitrogen cycle in deep- s-ea microorganisms from the Nankai Trough: genes for nitrification and denitrification from deep-sea environmental DNA. Extremophiles 11:269-275. Tamegai, H., L. Li, N. Masui, and C. Kato. 1997. A denitrifying bacterium from the deep sea at 11,000 m depth. Extremophiles 1:207-211. ter Braak, C. J. F. 1986. Canonical correspondence analysis: a new eigenvector technique from multivariate direct gradient analysis. Ecology 67 :1 167-1 179. Teske, A., K. U. Hinrichs,K, V. Edgcomb, A. de Vera Gomez, D. Kysela, S. P. Sylva, M. L. Sogin, and H. W. Jannasch. 2002. Microbial diversity of hydrothermal sediments in the Guaymas Basin: evidence for anaerobic methanotrophic communities. Appl. Environ. Microbiol. 68: 1994-2007. Tiquia, S. M., S. Gurczynski, A. Zholi, and A. Devol. 2006. Diversity of biogeochemical cycling genes from Puget Sound sediments using DNA microarrays. Environmental Technology. 27: 1377-1 389. Torsvik, V., L. Ovreas, and T. F. Thingstad. 2002. Prokaryotic diversity-magnitude, dynamics, and controlling factors. Science 296: 1 064-1066. Tuomainen, J. M., S. Hietanen, J. Kuparinen, P. J. Martikainen, and K. Servomaa. 2003. Baltic Sea cyanobacterial bloom contains denitrification and nitrification genes, but has negligible denitrification activity. FEMS Microbiol. Ecol. 45:83-96. Van Mooy, B. A. s., A. H. Devol, and R. G. Keil. 2003. Quantifying 3H-thymidine incorporation rates by a phylogenetically defined group of marine planktonic bacteria (Bacteroidetes phylum). Environ. Microbiol. 6: 1061-1069. Walsh, J. J. 1991. Importance of continental margins in the marine biogeochemical cycling of carbon and nitrogen. Nature 350:53-55. Ward, B. B. and A. R. Cockcroft. 1993. Immunofluorescence detection of the denitrifying strain Pseudomonas stutzeri (ATCC 14405) in seawater and intertidal sediment environments. Microb. Ecol. 25: 233- 246. Whitman, W. B., D. C. Coleman, and W. J. Wiebe. 1998. Prokaryotes: The unseen majority. Proc. Natl. Acad. Sci. 95:6578-6583. 205 CHAPTER 5 CONCLUSIONS AND FUTURE PERSPECTIVES At the time this study was started, microbial ecology was lacking an accurate way to rapidly quantify genes and the corresponding organisms in the environment, especially when specific groups that may not be abundant or active in the studied environment needed to be targeted. Real-time PCR effectively filled that niche and this study provided the background testing necessary before implementing it on a wide scale for the quantification of functional genes in the environment. It was proved that real-time PCR was specific, sensitive, precise and accurate on a wide array of environmental samples, which paved the way for its broad use. The detection of P. stutzeri nirS by this technique in a broad range of habitats, although at highly varying abundances, suggests that this commonly isolated denitrifier, though fairly cosmopolitan, might not be as dominant a member of the denitrifying community as most people had thought based on its frequent isolation. However, its persistence even in deep unmixed sediments, not exposed to oxygen or nitrate for years, in addition to its relative increase within the bacterial community due to sample manipulation, indicates a high level of resistance to various and changing environmental conditions and its responsiveness to favorable. growth conditions, probably explaining its ease of isolation. On the other hand, most of the detected nirS genes, as usually observed in diversity studies, were unrelated to any cultured denitrifier. Their relative decrease in the bacterial community after sample manipulation suggests that they might be sensitive to factors not considered when their isolation is attempted, explaining their absence from the cultured 206 strain collections. Some of these novel nirS genes, however, might also be part of alternative pathways in non-denitrifying organisms, as the nirS gene detected in the genome of the uncultured anammox bacterium Kuenenia stuttgartiensis, which has been suggested to play a role in an alternative anammox pathway (Strous et al., 2006). The question remains if these novel genes are all functional in the environment, or if at least some of them are remnants of nirS genes which have lost their functionality due to mutation in the gene itself or its regulatory system. Only isolation of denitrifiers with one of these novel genes will answer this question. The vertical distribution of denitrifiers was correlated to the bioturbation of the sediments and not to the redox gradients as had been often suggested. A highly conserved community structure associated also to denitrification capacity was detected throughout the mixed layer, but with its sharp shift in the deep unmixed sediments. Therefore, the eventual exposure of these denitrifiers to oxygen and nitrate by the particle mixing activity of burrowers as well as the increase of oxygen and nitrate exchange surface by the burrows themselves apparently led to the maintenance of an active denitrifying community far below the nitrate penetration depth into bulk sediments. The rapid initiation of denitrification after exposure to nitrate suggests that the denitrifying enzymes must already have been present in the organisms, although most of the bulk sediment bacteria were not exposed to nitrate in situ at the time of retrieval. Expression of these enzymes regardless of the presence of nitrate may give them a competitive advantage in an electron acceptor-limited environment. On the other hand, some of these bacteria may also survive based on alternative metabolic pathways, such as fermentation, which has been detected in the denitrifying bacterium Pseudovibrio denitrificans recently isolated 207 from shallow sea water (Shieh et al., 2004). In any case, this indicates that denitrification or at least denitrification capacity, extends well beyond the top few centimeters to which nitrate is often restricted, therefore suggesting the necessity of considering the whole mixed layer when modeling or predictions are made about denitrification and the responsible organisms. Denitrifier communities presented a much higher differentiation on the horizontal than on the vertical dimension, however different scales are involved for each. Sediments from different sites within one area with gradually changing environmental conditions revealed more distantly related communities than sediments from different depths within one core, but lower differentiation than sediments from distant geographic locations. Clone libraries suggested the existence of significantly different bacterial and denitrifier communities at Shallow Budd Inlet, Washington Margin, West of Juan de Fuca,Barrow Canyon Shallow, and East Hanna Shoal Shallow sediments, represented as well by the site-specific clustering of highly similar sequences observed for both genes. However, while nirS exhibited a higher similarity between the closely located Arctic transects, no clear grouping of 16S rRNA gene sequences from different sites was observed, and a higher similarity between the Arctic East Hanna Shoal Shallow and Washington Margin bacterial communities was suggested. This discrepancy could result from the lower coverage of the 16S rRNA gene libraries compared to the nirS gene libraries, but it might also result from the lower mutation rate of the 16S rRNA gene, allowing the further dispersal of highly similar 16S rRNA gene sequences. Hughes Martiny et al. (2006) suggested four alternative hypotheses about the effect of environmental, represented as habitats, and historical, represented as provinces, effects on communities. Depending on 208 the presence of single or multiple habitats and provinces, different distributions of communities would be observed. Correlation of biotic similarity with environmental similarity as well as with geographic distance would be indicative of the existence of multiple habitats as well as provinces for the distribution of microorganisms. The distribution of the nirS genes seems to correlate well with this view, while the distribution of the 16S rRNA genes suggests a tendency of certain provinces to fuse into bigger provinces, moving towards, although not reaching, the Baas-Becking hypothesis, “everything is everywhere, the environment selects”. Hence, the way in which we define “everything” e.g. a species or strain becomes an important issue, as the results Will vary according to that definition and the characters measured. If we concentrate on all genes in a genome as used by Konstantinidis and Tiedje (2005), more differences, and therefore more provinces will be established. This tendency will increase even further if intergenic regions are also considered. However, the question remains as to how important these differences are in the functioning of the organism and if they actually represent adaptations to the environment or are the result of neutral ecological drift and other factors as has been recently suggested (Ramette and Tiedje, 2007). In this study, the analyzed sites were not located along a transect or some clear gradient of environmental conditions, but rather formed a patchy array. This makes the correlation of the genotypic patterns seen with environmental variables more difficult, as several variables vary in unrelated ways from one site to the other. However, a general trend towards higher diversity in intermediate productivity areas was detected, a pattern also observed on the vertical axis. If this results from a balance between competitive ability and resistance to predators remains to be established. Understanding the effect that 209 productivity has on the development of the denitrifier and general bacterial community is of great importance to better predict changes that could result from increased productivity in areas such as the Arctic, which has been suggested will be most highly affected by global warming. Future perspectives As stated above, most of the nirS gene sequences retrieved in this study, as well as in other diversity studies, are unrelated to known denitrifiers. However, their high abundance suggests that the bacteria that harbor them might play an important role in the community. This remains questionable until the functionality of these novel genes can be proven. The detection of mRNA sequences (Nogales et al., 2002) closely related to some of the novel nirS genes detected in this and other studies, and, furthermore, their high abundance in environmental samples (Smith et al., 2007) indicates that at least some of these novel genes are being expressed in nature, constituting a further evidence of their functionality. However, the translation into a functional protein of one of these abundant sequences has not yet been demonstrated. Several approaches have been tried in this study (Appendix) to retrieve an integral novel nirS gene in order to test its functionality in a complementation assay. The complexity of the environmental DNA sample was, however, a limiting factor and impeded the successful retrieval of a complete nirS gene from marine sediments. Enrichment of the target novel nirS gene sequence with a magnetic capture-hybridization assay (Jacobsen, 1995) might help overcome this problem. However, the regulation of the expression of this gene and the ultimate function in the original organism can only be elucidated by isolation of a denitrifying organism 210 that carries the gene. This has proven to be a difficult task and the apparent sensitivity of some of these novel denitrifiers to manipulation of the samples might be a cause as well as our inability to find the right selective conditions. Therefore, screening for the presence of these novel genes along with careful dissection of the various steps involved in the attempt to isolate a novel denitrifier might help decipher critical factors leading to the loss of these organisms during the isolation process. Recently, several new SARll and other uncultured strains have been isolated from seawater with an improved high- throughput dilution-to-extinction protocol (Stingl et al., 2007). Although not necessarily effective for the more complex sediment community, some of the improvements could be considered to aid in the isolation of novel denitrifiers from this environment. A common denominator for all the clone libraries presented in this study was the low coverage of the diversity present, especially for the 16S rRNA gene. Traditional cloning and sequencing techniques limited the number of sequences that could be obtained to several tens or hundreds. However, with the development of pyrosequencing (Ronaghi et al., 1996) and its improvement to 454 pyrosequencing (Margulies et al., 2005) for high-throughput generation of sequencing data, tens of thousands of sequences can be generated in only a few hours, significantly improving our ability to cover the diversity present in the environment of interest. Roesch et al. (2007) recently studied the microbial diversity in soil from four different areas with this technique, which allowed them to obtain between 26,140 and 53,533 sequences from each site, more than two orders of magnitude more than the number generated in this study. Most of the taxonomic groups identified in these soil samples, were, however, present at less than 0.5% relative abundance in the community, suggesting that they most probably would not have been 211 detected in a clone library of a similar size to the ones used in this study. Sediments have been shown to be as diverse as soil, therefore, applying 454 pyrosequencing to the sediments used here would as well give us further information about the broad diversity present to better understand the function of the community and the players involved. Furthermore, larger coverage would also help to better establish if the differences we are seeing between samples are real or artifacts due to undersampling of the diversity present. In any case, a deeper knowledge on the fine tuning of microbial communities is beginning to be possible. 212 References Baas-Becking, L. G. M. 1934. Geobiologie of Inleiding Tot de Milieukunde. W. P. van Stockum & Zoon N. V., The Hague, The Netherlands. Hughes Martiny, J. B., B. J. M. Bohannan, J. H. Brown, R. K. Colwell, J. A. Fuhrman, J. L. Green, M. C. Horner-Devine, M. Kane, J. Adams Krumins, C. R. Kuske, P. J. Morin, S. Naeem, L. flvrer‘is, A.-L. Reysenbach, V. H. Smith, and J. T. Staley. 2006. Microbial biogeography: putting microorganisms on the map. Nature 4:102-112. Jacobsen, C. S. 1995. Microscale detection of specific bacterial DNA in soil with a magnetic capture-hybridization and PCR amplification assay. Appl. Environ. Microbiol. 61:3347-3352. Konstantinidis, K. T. and J. M. Tiedje. 2005. Genomic insights that advance the species definition for prokaryotes. Proc. Natl. Acad. Sci. 102:2567-2572. Margulies, M., M. Egholm, W. E. Altman, S. Attiya, J. S. Bader, L. A. Bemben, J. Berka, M. S. Braverman, Y.-J. Chen, Z. Chen, S. B. Dewell, L. Du, J. M. Fierro, X. V. Gomes, B. C. Godwin, W. He, S. Helgesen, C. H. Ho, G. P. Irzyk, S. C. Jando, M. L. I. Alenquer, T. P. Jarvie, K. B. Jirage, J.-B. Kim, J. R. Knight, J. R. Lanza, J. H. Leamon, S. M. Lefkowitz, M. lei, J. Li, K. L. Lohman, H. Lu, V. B. Makhijani, K. E. McDade, M. P. McKenna, E. W. Myers, E. Nickerson, J. R. Nobile, R. Plant, B. P. Puc, M. T. Ronan, G. T. Roth, G. J. Sarkis, J. F. Simons, J. W. Simpson, M. Srinivasan, K. R. Tartaro, A. Tomasz, K. A. Vogt, G. A. Volkmer, S. H. Wang, Y. Wang, M. P. Weiner, P. Yu, R. F. Begley, and J. M. Rothberg. 2005. Nature 437 :376- 380. Nogales, B., K. N. Timmis, D. B. Nedwell and A. M. Osborn. 2002. Detection and diversity of expressed denitrification genes in estuarine sediments after .reverse transcription-PCR amplification from mRNA. Appl. Environ. Microbiol. 68:5017-5025. Ramette, A. and J. M. Tiedje. 2007. Multiscale responses of microbial life to spatial distance and environmental heterogeneity in a patchy ecosystem. Proc. Natl. Acad. Sci. 104:2761-2766. Roesch, L. F. W., R. R. Fulthorpe, A. Riva, G. Casella, A. K. M. Hadwin, A. D. Kent, S. H. Daroub, F. A. O. Camargo, W. G. Farmerie, and E. W. Triplett. 2007. Pyrosequencing enumerates and contrasts soil microbial diversity. The ISME Journal 1:283-290. Ronaghi, M., S. Karamohamed, B. Petterson, M. Uhlén, and P. Nyrén. 1996. Real- time DNA sequencing using detection of pyrophosphate release. Anal. Biochem. 242:84- 89. 213 Shieh, W. Y., Y.-T. Lin, and W. D. Jean. 2004. Pseudovibrio denitrificans gen. nov., sp. nov., a marine, facultatively anaerobic, ferrnentative bacterium capable of denitrification. Int. J. Syst. Evol. Microbiol. 54:2307-2312. Smith, C. J., D. B. Nedwell, L. F. Dong, and A. M. Osborn. 2007. Diversity and abundance of nitrate reductase genes (narG and napA), nitrite reductase genes (nirS and nrfA), and their transcripts in estuarine sediments. Appl. Environ. Microbiol. 73:3612- 3622. Stingl, U., H. J. Tripp, and S. J. Giovannoni. 2007. Improvements of high-throughput culturing yielded novel SARll strains and other abundant marine bacteria from the Oregon coast and the Bermuda Atlantic Time Series study site. The ISME Journal 1:361- 371. Strous, M., E. Pelletier, S. Mangenot, T. Rattei, A. Lehner, M. W. Taylor, M. Horn, H. Daims, D. Bartol-Mavel, P. Wincker, V. Barbe, N. Fonknechten, D. Vallente, B. Segurens, C. Schenowitz-Truong, C. Médigue, A. Collingro, B. Snel, B. E. Dutilh, H. J. M. Op den Camp, C van der Drift, I. Cirpus, K. T. van de Pas-Schoonen, H. R. Harhangi, L. van Niftrik, M. Schmid, J. Keltjens, J. van de Vossenberg, B. Kartal, H. Meier, D. Frishman, M. A. Huynen, H.-W. Mewes, J. Weissenbach, M. S. M. Jetten, M. Wagner, and D. Le Paslier. 2006. Deciphering the evolution and metabolism of an anammox bacterium from a community genome. Nature 440:790-794. 214 APPENDIX RECOVERY OF NOVEL nirS GENES FROM NATURE AND TESTING OF THEIR FUNCTIONALITY Molecular techniques have led to the retrieval from nature of an enormous number of novel gene sequences which might play an important role in the chemistry of the biosphere. However, the functionality of these sequences has generally not been tested and therefore their in situ function remains questionable. This is also true for denitrification genes, like the nirS gene coding for the heme ed, nitrite reductase. In previous studies, three clusters of nirS clones which did not include any nirS from a cultured denitrifier were detected in sediment samples from the Pacific Northwest (Braker et al., 2000). Furthermore, in Chapter 3, some clusters of novel nirS sequences were identified in Shallow Budd Inlet sediments, which were more abundant than commonly isolated denitrifiers such as Pseudomonas stutzeri at this site. These novel sequences could correspond to novel denitrifiers, however their functionality is unknown. A complementation assay was therefore applied in order to test the ability of these sequences to restore the denitrification pathway in a NirS' mutant. The complementation assay involved the cloning of a novel nirS gene into the pSUP104 vector, a broad host range vector which can be maintained in Pseudomonas (Priefer et al., 1985). Zumft et al. (1988) developed a Tn5-induced mutant strain of P. stutzeri (mutant MK202), which lacks the heme cdl nitrite reductase and is therefore defective in nitrite reduction. This mutant strain was used to test the ability of the cloned nirS gene to complement the defect in nitrite reduction after the introduction of the vector pSUPlO4 with the potential nirS insert into the mutant by conjugation following the protocol by Zumfi et al. (1985). The restoration of the denitrification pathway was tested by growth in presence of nitrite as the only electron acceptor. The previously retrieved nirS clones represent only part of the nirS gene sequence. Therefore, a complete gene sequence had to be retrieved first in order to test its functionality by complementation of the nirS mutant. Several different approaches were used 1n my attempts to achieve this. - Direct cloning of partially restricted environmental DNA (~2000-4500 bp long) retrieved from Shallow Budd Inlet sediments into pSUPlO4. - Inverse PCR to retrieve the ends and flanking regions of a novel nirS gene, which would allow the design of specific primers for the amplification of a complete novel nirS gene. Inverse PCR involved partial restriction of environmental genomic DNA followed by self-ligation and amplification by primers representing reverse complements of general nirS primers nirSlF and nirS6R (Braker et al., 1998). 215 - Inverse PCR, as above, but using primers representing reverse complements of cluster—specific primers designed for the quantification by real-time PCR of abundant novel clusters retrieved from the same site (Chapter 3). - Amplification of ends and flanking regions of novel nirS genes, by combination of general or cluster-specific nirS primers with ERIC primers for repetitive elements (Rademaker et al., 1998). Regretfully, none of the approaches led to the retrieval of a novel nirS gene fiom environmental samples. Direct cloning and the use of ERIC primers seem to be the least appropriate, due to the large number of steps with loss of genetic material involved in the former, and the significantly higher amplification of fragments by ERIC primers alone rather than combined with nirS primers in the latter. Inverse PCR, especially with cluster- specific primers, seems to be the more reasonable approach, however, some improvements could be done, such as the use of restriction enzymes leaving protruding ends, rather than blunt ends (AluI and RsaI were used) for the partial restriction of the DNA, which would increase the ligation efficiency. The complexity of the sample was probably the most detrimental factor. Therefore, previous enrichment of the target gene, for example by the use of a magnetic capture-hybridization assay (Jacobsen, 1995), might help overcome this problem. 216 References Braker, G., A. F esefeldt, and K.-P. Witzel. 1998. Development of PCR primer systems for amplification of nitrite reductase genes (nirK and nirS) to detect denitrifying bacteria in environmental samples. Appl. Environ. Microbiol. 64:3769-3775. Braker, G., J. Zhou, L. Wu, A. H. Devol, J. M. Tiedje. 2000. Nitrite reductase genes (nirK and nirS) as functional markers to investigate diversity of denitrifying bacteria in Pacific Northwest marine sediment communities. Appl. Environ. Microbiol. 66: 2096— 2104. Jacobsen, C. S. 1995. Microscale detection of specific bacterial DNA in soil with a magnetic capture-hybridization and PCR amplification assay. Appl. Environ. Microbiol. 61:3347-3352. Priefer, U. B., R. Simon, and A. Piihler 1985. Extension of the host range of Escherichia coli vectors by incorporation of RSF 1010 replication and mobilization functions. J. Bacteriol. 163:324-330. Rademaker, J. L. W., F. J. Louws, and F. J. de Bruijn. 1998. Characterization of the diversity of ecologically important microbes by rep-PCR genomic fingerprinting. In A. D. L. Akkermans, J. D. van Elsas, and F. J. de Bruijn (eds), Molecular Microbial Ecology Manual. Kluwer Academic Publishers, Dordrecht, The Netherlands. Zumft, W. G., K. Dohler, and H. Kiirner 1985. Isolation and characterization of transposon Tn5-induced mutants of Pseudomonas perfectomarina defective in nitrous oxide respiration. J. Bacteriol. 163:918-924. Zumft, W. G., K. Dohler, H. Kiirner, S. Lochelt, A. Viebrock, and K. Frunzke 1988. Defects in cytochrome cdl-dependent nitrite respiration of transposon TNS-induced mutants from Pseudomonas stutzeri. Arch.Microbiol. 149:492-4 217 IIIjlljljjjijilflljjljjl[WilliI