ASSESSING THE BIODEGRADATION OF TOLUENE, ETHYLBENZENE AND RDX AND THE IDENTIFICATION OF THE MICROORGANISMS INVOLVED USING STABLE ISOTOPE PROBING AND HIGH THROUGHPUT AMPLICON SEQUENCING By Indumathy Jayamani A DISSERTATION Submitted to Michigan State University In partial fulfillment of the requirements for the degree of Environmental Engineering – Doctor of Philosophy 2015 ABSTRACT ASSESSING THE BIODEGRADATION OF TOLUENE, ETHYLBENZENE AND RDX AND THE IDENTIFICATION OF THE MICROORGANISMS INVOLVED USING STABLE ISOTOPE PROBING AND HIGH THROUGHPUT AMPLICON SEQUENCING BY Indumathy Jayamani Contamination of groundwater by organic pollutants is a worldwide environmental problem. Bioremediation is a viable option for cleaning and reclaiming sites contaminated with pollutants amenable to microbial transformation. However, our understanding of microorganisms playing key roles in bioremediation is still developing. In this study, specific aspects of biodegradation of three organic pollutants namely toluene, ethylbenzene and the nitramine explosive hexahydro1,3,5-trinitro-1,3,5-triazine (RDX) were investigated. The overall objectives of this research were to 1) assess the biodegradation potential of toluene, ethylbenzene and RDX using soils from various sources, 2) identify the major groups of microorganisms responsible for the degradation of ethylbenzene and RDX in these samples and 3) assess the effect of a potential cocontaminant, isobutanol, on toluene biodegradation. The first study examined the effect of isobutanol on the biodegradation of toluene under sulfate amended, nitrate amended or methanogenic conditions. The results indicated that toluene biodegradation was not greatly affected by isobutanol in five of the six experimental set-ups. However, toluene biodegradation was completely inhibited in one set of microcosms amended with sulfate and inocula from wastewater treatment plant activated sludge. This suggests that if co-contamination occurs, in some cases toluene degradation may be inhibited. In the second study, stable isotope probing (SIP) and high throughput sequencing were used with ethylbenzene degrading consortia to identify microorganisms benefiting from 13C label uptake from ethylbenzene (or metabolites). Several phylotypes were relatively more abundant in the heavy fractions from the labeled ethylbenzene amended soil microcosms compared to the controls indicating 13C label uptake. This included phylotypes within the families Oxalobacteraceae, Rhodospirillaceae, Xanthomonadaceae and Rhodocyclaceae (Proteobacteria) as well as the genus Gemmatimonas. This work indicates microorganisms not previously linked to ethylbenzene degradation could have significant roles in the carbon uptake from this pollutant. The third and fourth studies involved applying SIP and high throughput sequencing to investigate RDX degrading microbial communities. In the third study, microbial communities obtained from four soils previously unexposed to explosives were investigated. Sequences from the total DNA extracts of all soils illustrated an increase in abundance of Brevundimonas and/or unclassified Bacillaceae 1 compared to the microbial communities in the initial soil or no RDX treatments. The fourth study investigated a RDX degrading community obtained from a Navy Base previously contaminated with explosives with and without glucose. The microbial communities in the total DNA samples indicated phylotypes classing as Pseudomonadaceae and Acinetobacter was more abundant in the presence and absence of glucose respectively. The SIP study also found that unclassified Pseudomonadaceae were primarily responsible for label uptake in both treatments. When glucose was present, Comamonas also increased in abundance following RDX degradation and was enriched in the heavy fractions, suggesting, for the first time, that this phylotype is also important for RDX removal. Overall, these data suggest both novel RDX degraders and previously reported RDX degraders were associated with RDX removal in these soils. Dedicated to all the great people in my life iv ACKNOWLEDGEMENTS This journey was a life changing experience and was made possible by many kind hearts. While the space here might not allow me to list all, I will do my best to list most. First of all, I am very grateful for this opportunity to conduct doctoral research in this field and sincerely thank my research advisor and mentor Dr. Alison M. Cupples for that. She has been nothing less of supportive, encouraging and patient throughout the course of my studies. She has seen me through my ups and downs and I am grateful for her trust in my work. I also would like to thank my dissertation committee members Dr. Syed Hashsham, Dr. Irene Xagoraraki and Dr.Terrence Marsh, whose guidance, support and comments makes this work complete. I would also like to extend my thanks to my mentors from Dr. Cupples group, Dr. Weimin Sun, Dr. Chunlin Luo and Dr. Shuguang Xie, who helped me learn the basics of the work that I present here. I am also indebted to Dr. Weimin Sun for providing me enrichment culture samples for the work presented in chapter 2. I am grateful to Yanlyang Pan for his assistance with the analytical instruments. Special thanks to my friend and colleague Fernanda Paes for the numerous discussions and evenings of bouncing ideas off each other and for her constant support through my studies. I am also thankful to Hanna Miller, Alyse Way, Kathyleen Haynes and Jordyn Davis who helped me as part of their undergraduate research work. I owe my gratitude to members of Dr.Hashsham’s, Dr. Xagoraraki’s and Dr. Voice’s laboratory - Dr. Robert D. Stedfeld, Tiffany L. Stedtfeld, Dr. Amanda Herzog, Maggie Kronlein, Mariya Munir, Amira Oun and Chaiyanun Tangtong for letting me share their instruments and for the countless research discussions. I am also grateful to my teaching mentor Dr. Masten for her tremendous support. I would also like to extend my thanks to my colleagues Yang Song, Liang Ding, Jean Rene Thelusmond and Yogendra v Kanitkar. Thanks to Lori Larner, Margaret Connor, Mary Morz and Laura Taylor for their support during my studies. The research work presented here was funded by two different research grants funded to Dr. Cupples and I am grateful for that (National Science Foundation grant number 0853249 and Strategic Environmental Research and Development Program, Project ER1606). My sincere thanks to members of Dr. Patrick Schloss laboratory at University of Michigan, particularly Ms. Sarah Wescott for her advice during the Mothur analysis. Also thanks to Dr. Mark E. Fuller and Dr. Paul Hatzinger (CB&I Federal Services) for providing samples from the Navy Base Site for the work presented in chapter 5. Thanks also to all my friends and family in Michigan providing me a home away from home. Last but not least, I am grateful to my husband Purushotham Pakala for his love and support through my endeavors. vi TABLE OF CONTENTS LIST OF TABLES…………………………………………………………………………………..ix LIST OF FIGURES………………………………………………………………………………….x KEY TO ABBREVIATIONS………………………………………………………………………..xiii CHAPTER 1 INTRODUCTION ............................................................................................................ 1 GROUNDWATER POLLUTION BY BTEX AND RDX ....................................................................... 1 BIODEGRADATION OF TOLUENE, ETHYLBENZENE AND RDX ....................................................... 1 STABLE ISOTOPE PROBING AND HIGH THROUGHPUT SEQUENCING .............................................. 3 OBJECTIVES .................................................................................................................................. 3 REFERENCES ................................................................................................................................ 7 CHAPTER 2 EFFECT OF ISOBUTANOL ON TOLUENE BIODEGRADATION IN NITRATE AMENDED, SULFATE AMENDED AND METHONOGENIC ENRICHMENT MICROCOSMS.................................. 13 ABSTRACT .................................................................................................................................. 13 INTRODUCTION ........................................................................................................................... 14 METHODS ................................................................................................................................... 16 Microcosm preparation .......................................................................................................... 17 Analytical methods ................................................................................................................ 17 RESULTS AND DISCUSSION ......................................................................................................... 18 CONCLUSIONS ............................................................................................................................ 21 APPENDIX ................................................................................................................................... 23 REFERENCES .............................................................................................................................. 31 CHAPTER 3 STABLE ISOTOPE PROBING AND HIGH THROUGHPUT SEQUENCING IMPLICATES OXALOBACTERACEAE, RHODOSPIRILLACEAE, XANTHOMONADACEAE AND RHODOCYCLACEAE IN ETHYLBENZENE DEGRADATION .................................................................................................. 35 ABSTRACT .................................................................................................................................. 35 INTRODUCTION ........................................................................................................................... 36 METHODS ................................................................................................................................... 39 Chemicals .............................................................................................................................. 39 Development of Ethylbenzene Degrading Microcosms ........................................................ 39 Analytical Techniques ........................................................................................................... 40 DNA Extraction and Isopycnic Centrifugation ..................................................................... 41 High Throughput Amplicon Sequencing (Illumina MiSeq) .................................................. 42 Cloning and Sanger Sequencing ............................................................................................ 43 RESULTS ..................................................................................................................................... 44 Ethylbenzene Biodegradation ................................................................................................ 44 High Throughput Amplicon Sequencing Analysis................................................................ 44 Microbial Community in Ethylbenzene Degrading Microcosms .......................................... 45 16S rRNA Gene Clone Library ............................................................................................. 47 Phylotype Relative Abundance Across SIP Fractions........................................................... 48 DISCUSSION ................................................................................................................................ 49 vii APPENDIX ................................................................................................................................... 52 REFERENCES .............................................................................................................................. 61 CHAPTER 4 COMPARATIVE STUDY OF RDX DEGRADING MICROBIAL COMMUNITIES IN FOUR SOILS ............................................................................................................................................. 68 ABSTRACT .................................................................................................................................. 68 INTRODUCTION ........................................................................................................................... 69 METHODS ................................................................................................................................... 72 Chemicals .............................................................................................................................. 72 Experimental Design ............................................................................................................. 72 RDX Extraction and Analysis ............................................................................................... 73 DNA Extraction and Isopycnic Centrifugation ..................................................................... 74 Amplicon Sequencing and Data Analysis ............................................................................. 75 RESULTS AND DISCUSSION ......................................................................................................... 76 RDX Degradation .................................................................................................................. 76 Comparison at Phyla Level ................................................................................................... 76 Microbial Community Analysis ............................................................................................ 78 Phylotypes Increasing in Abundance Following RDX Degradation ..................................... 80 Phylotypes Indicated In Label Uptake .................................................................................. 82 APPENDIX ................................................................................................................................... 85 REFERENCES .............................................................................................................................. 97 CHAPTER 5 STABLE ISOTOPE PROBING REVEALS THE IMPORTANCE OF COMAMONAS AND PSEUDOMONADACEAE IN RDX DEGRADATION IN SAMPLES FROM A NAVY DETONATION SITE ..................................................................................................................................................... 105 ABSTRACT ................................................................................................................................ 105 INTRODUCTION ......................................................................................................................... 106 METHODS ................................................................................................................................. 109 Chemicals ............................................................................................................................ 109 Experimental Setup and DNA Extraction ........................................................................... 109 Analytical Methods ............................................................................................................. 110 Isopycnic Centrifugation ..................................................................................................... 111 Illumina MiSeq™ High Throughput Amplicon Sequencing .............................................. 112 Data Analysis....................................................................................................................... 113 RESULTS ................................................................................................................................... 114 DISCUSSION .............................................................................................................................. 117 APPENDIX ................................................................................................................................. 122 REFERENCES ............................................................................................................................ 129 viii LIST OF TABLES Table 2.1. Experimental setup for nitrate amended, sulfate amended and methanogenic toluene degrading samples. ........................................................................................................................ 24 Table 3.1. Buoyant density of heavy fractions from the labeled and unlabeled samples submitted for amplicon sequencing. .............................................................................................................. 53 Table 3.2. Summary MiSEQ Illumina data generated from the fractions as well as total DNA extracted from labeled and unlabeled ethylbenzene amended microcosms ................................. 53 Table 3.3. Classification of 16S rRNA gene clones in the clone library. .................................... 54 Table 4.1. The properties of the four agricultural soils used in this study. .................................. 86 Table 4.2. A summary of MiSEQ Illumina data generated from all genomic total DNA samples and fractions for Soil 9 and Soil 10. ............................................................................................. 86 Table 4.3. The classification of anaerobic RDX degrading isolates from previous studies. ....... 87 Table 5.1. Chao1 and Shannon and lower and upper confidence intervals (CI) values for all microbial communities................................................................................................................ 123 ix LIST OF FIGURES Figure 2.1. Toluene concentrations in nitrate amended microcosms derived from enrichments constructed from agricultural soil 1(a), agricultural soil 2 (b) and agricultural soil 3 (c). ........... 25 Figure 2.2. Toluene concentrations in sulfate amended microcosms derived from enrichments constructed from agricultural soil 1(a) wastewater treatment plant digester sludge with 2.5 ppm isobutanol (b), or with 5 ppm isobutanol (c)................................................................................. 26 Figure 2.3. Toluene concentrations in methanogenic microcosms derived from enrichments constructed from wastewater treatment plant digester sludge. ..................................................... 27 Figure 2.4. Isobutanol concentrations in nitrate amended toluene degrading microcosms derived from enrichments constructed from agricultural soil 1 (a), agricultural soil 2 (b) and agricultural soil 3 (c). ....................................................................................................................................... 28 Figure 2.5. Isobutanol concentrations in sulfate amended toluene degrading microcosms derived from enrichments constructed with agricultural soil 1 (a), and wastewater treatment plant digester sludge 2.5 ppm isobutanol (b) or 5 ppm isobutanol (c). ................................................. 29 Figure 2.6. Isobutanol concentrations in methanogenic toluene degrading microcosms derived from enrichments constructed with wastewater treatment plant digester sludge. ........................ 30 Figure 3.1. Concentration of ethylbenzene in microcosms over time ( × - abiotic controls, open triangles - amended with 12C6 ethylbenzene, closed circles - amended with 13C6 ethylbenzene). 55 Figure 3.2. DNA concentration over the ultracentrifugation gradient in fractions from replicate microcosms amended with labeled or unlabeled ethylbenzene. The dashed arrows indicate the fractions that were sequenced from the unlabeled EB amended samples and solid arrows indicate the fractions that were sequenced from the labeled EB amended samples................................... 56 Figure 3.3. Rarefaction curves generated from sequencing data from total DNA extracts for all six microcosms.............................................................................................................................. 57 Figure 3.4. Comparison of relative abundance of sequences in total genomic DNA extracted from six microcosms amended with 13C6 labeled or unlabeled ethylbenzene at phylum level classification ................................................................................................................................. 58 Figure 3.5. Relative abundance (%) of the most common phylotypes (>1%, at the lowest classification level) in all six microcosms (from total DNA extracts). The phylotypes are at the genus level, unless unclassified. ................................................................................................... 59 Figure 3.6. Relative abundance of the most common phylotypes in heavy fractions from 13C ethylbenzene amended microcosms compared to those from 12C ethylbenzene amended microcosms (SIP Label EB1 and SIP Unlabel EB1 on left, SIP Label EB2 and SIP Unlabel EB2 on right). ........................................................................................................................................ 60 x Figure 4.1. RDX in microcosms amended with labeled RDX (13C3, 15N3), unlabeled RDX (12C3, 14 N3), and killed abiotic controls for Soil 9 (A), Soil 10 (B), Soil 11 (C), Soil 12 (D). Error bars denote standard deviation from triplicate micrososms. Arrows indicate hypothetical RDX concentrations on the day of respike. ............................................................................................ 88 Figure 4.2. Phyla level comparison between the four soils of the communities in 1) the initial soils (No RDX, Time 0), 2) replicates of the labeled and unlabeled RDX amended microcosms (Labeled 1 & 2, Unlabeled 1 & 2, Time t) and 3) microcosms without RDX addition (No RDX, Time t)........................................................................................................................................... 89 Figure 4.3. MLTree Map (A) and heatmap analysis (B) of the phylotypes within the initial soils and in samples with and without RDX. To simplify the presentation, only one replicate was used for the RDX amended samples. .................................................................................................... 90 Figure 4.4. Rarefaction curves (A) and Principal Component Analysis (B) of sequencing data generated from all four soils, including total DNA extracts of the labeled RDX (Labeled 1 & 2, Time t), unlabeled RDX (Unlabeled 1 & 2, Time t), no RDX (No RDX, Time t) microcosms and the initial soils (No RDX, Time 0)................................................................................................ 92 Figure 4 5. A comparison of the relative abundance (%) of the most common phylotypes (>1% , at the lowest classification level) in all four soils in the no RDX samples, including the initial soils (No RDX, Time 0) (A) and the microcosms without RDX (No RDX, Time t) (B). ............ 93 Figure 4.6. A comparison of the relative abundance (%) of the most common phylotypes (>1% , at the lowest classification level) in all four soils (A-D) following RDX degradation. ............... 94 Figure 4.7. The relative abundance (%) of the eight most abundant phylotypes in the four RDX amended agricultural soils (Labeled 1 &2, Unlabeled 1 &2, Time t) compared to the RDX controls (initial soil and no RDX microcosms). Boxed phylotypes illustrate a greater abundance in RDX amended samples compared to the RDX controls for at least three soils. The phylotypes are at genus level, unless unclassified........................................................................................... 95 Figure 4.8. Relative abundance of the most common phylotypes in heavy fractions from the labeled RDX amended microcosms compared to those from the unlabeled RDX amended microcosms in Soil 9 (A) and Soil 10 (B). Bars illustrate standard deviations from triplicate sequencing data. ............................................................................................................................ 96 Figure 5.1. Relative abundance of the most common phylotypes in RDX amended microcosms (with and without glucose) compared to their abundance in the initial soil community and no RDX microcosms. ....................................................................................................................... 124 Figure 5.2. Rarefaction curves for the total microbial communities in the initial soil, RDX amended microcosms and the no RDX controls for both the no glucose amended (A) and glucose amended (B) microcosms. .......................................................................................................... 125 Figure 5.3. Relative abundance of phylotypes illustrating a difference in relative abundance between the RDX amended microcosms and the no RDX controls (initial soil and no RDX xi microcosm) in the glucose amended microcosms (A) and in the microcosms that were not amended with glucose (B). ......................................................................................................... 126 Figure 5.4. DNA concentration (ng/µL) in fractions across buoyant density gradients obtained from samples amended with labeled and unlabeled RDX, with (A) or without (B) the addition of glucose. Replicate lines represent duplicate DNA measurements. The shaded area represents the fractions analyzed with Illumina sequencing. ............................................................................. 127 Figure 5.5. The three most abundant phylotypes in the heavy fractions from the labeled RDX amended microcosms compared to their abundance in fractions of similar buoyant density from the unlabeled RDX amended microcosms. Data are shown from those amended with glucose (A, top three graphs) and those not amended with glucose (B, bottom three graphs). The y-axis represents relative abundance in each fraction normalized by the DNA mass in each fraction (relative abundance times the DNA mass (ng)). Error bars represent standard deviations from three replicates and may be too small to be seen. ....................................................................... 128 xii KEY TO ABBREVIATIONS BTEX - benzene, toluene, ethylbenzene and o-, p-, and m-xylenes CFR – Code of Federal Regulations DGGE – Denaturing gradient gel electrophoresis DNA - deoxyribonucleic acid LUST – leaking underground storage tanks OTU - operational taxonomic unit PCR - polymerase chain reaction RDX – royal demolition explosive (hexahydro-1,3,5-trintro-1,3,5-triazine) rRNA - ribosomal RNA SIP – stable isotope probing TRFLP - terminal restriction fragment length polymorphism xiii CHAPTER 1 INTRODUCTION GROUNDWATER POLLUTION BY BTEX AND RDX The chemicals benzene, toluene, ethylbenzene and the xylenes, collectively known as BTEX compounds, are a major source of groundwater contamination from leaking gasoline tanks. According the USEPA, as of March 2014, over half of a million releases have been confirmed in the United States from underground storage tanks (UST) containing gasoline (1). Groundwater contamination near military sites with the nitroamine explosive hexahydro-1,3,5-trinitro-1,3,5triazine (RDX) is widespread due to the explosives manufacturing and handling particularly after World War II (2). Groundwater contamination by such organic pollutants threatens drinking water sources due to their toxic health effects. Bioremediation is a viable remedy option for these contaminants as they are amenable to bacterial degradation. In situ biodegradation of organic pollutants can depend on numerous factors including the microbial community, nutrient availability, physical and chemical factors and co-contaminant effects. It is also more appropriate to consider degradation of such pollutants under anaerobic conditions (as groundwater is often depleted in oxygen). The research presented in this thesis relates to the microorganisms involved in the biodegradation of the organic contaminants toluene, ethylbenzene and RDX. BIODEGRADATION OF TOLUENE, ETHYLBENZENE AND RDX Microorganisms that are able to metabolize or mineralize toluene under nitrate reducing (3-6), iron reducing (7, 8) and sulfate reducing (9) conditions have been isolated thus far. However, 1 very little is known about the effects of a potential co-contaminant (isobutanol) on toluene degradation. To date, only three studies have investigated the effect of isobutanol on BTEX, two under aerobic conditions (10, 11) and one under anaerobic conditions(12). In this study, we extend on knowledge by investigating the effect of isobutanol on toluene (a model BTEX compound) in six different toluene degrading communities. In contrast to toluene, only five anaerobic ethylbenzene degrading isolates exist. Among these, four can degrade ethylbenzene under nitrate reducing conditions (13). These include two Azoarcus sp. strains (Aromatoleum aromaticum EbN1 and EB1) (14-16) , Dechloromonas aromatica RCB (6, 17), Georgfuchsia toluolica (18) and Desulfobacteraceae Strain EbS7 (19). Interestingly, all except Strain EbS7 (Deltaproteobacteria) belong to the family Rhodocyclaceae (Betaproteobacteria). Compared to the BTEX chemicals, many microorganisms have been associated with RDX degradation. Many anaerobic RDX degraders have been isolated from the phyla Proteobacteria (20-32) and Firmicutes (21, 29, 33-38) with one in the phylum Fusobacteria (28). Although contaminant degradation has been confirmed with pure cultures, it is still uncertain if these isolates can degrade ethylbenzene or RDX when present in a mixed culture. To address this unknown, stable isotope probing (SIP) was used in this research to link function to identity in mixed communities. 2 STABLE ISOTOPE PROBING AND HIGH THROUGHPUT SEQUENCING SIP involves amending a stable isotope labeled (e.g. 13C, 15N) compound to a mixed microbial community, such as a soil or water sample (39). Isopycnic centrifugation of the extracted DNA in cesium chloride gradients allows the DNA to band based on it’s buoyant density. Thus, this technique allows the separation of DNA that has incorporated the label (heavy DNA) from the background community DNA (unlabeled or light DNA). SIP studies have been used to identify microorganisms involved in biogeochemical cycles (40, 41) and environmental pollutant remediation (42-45). These studies employed finger printing techniques such as denaturing gradient gel electrophoresis (DGGE), terminal restriction fragment length polymorphism (TRFLP) and cloning and sequencing methods. Recently, researchers have combined SIP with high throughput sequencing to obtain a greater depth of information than that achievable with fingerprinting techniques or Sanger sequencing. (46-48). In this study, we combined SIP with high throughput amplicon sequencing (MiSeq Illumina) to identify the microorganisms in mixed communities involved in ethylbenzene/RDX biodegradation. OBJECTIVES The overall objectives were, i) To assess toluene, ethylbenzene and RDX biodegradation potential in soils from various sources. ii) To identify the major groups of microorganisms responsible for degradation of ethylbenzene and RDX. iii) To assess the effect of a potential co-contaminant, isobutanol, on toluene biodegradation. 3 Specific objectives addressed in each chapter are as follows: 1. To assess the effect of a potential co-contaminant (isobutanol) on a model gasoline compound (toluene) under various electron acceptor conditions in laboratory microcosms (Chapter 2). 2. To identify the microorganisms responsible for carbon uptake from ethylbenzene under oxygen depleted conditions (Chapter 3). 3. To determine and compare the dominant RDX degraders in four uncontaminated soils and in a soil from a contaminated site (Chapters 4 and 5). The research work presented under Chapters 2 and 3 were funded by an NSF grant awarded to Dr. Cupples entitled “Stable isotope probing to assess bioremediation of LUST contaminants: addressing existing MTBE/BTEX and probably future ethanol/BTEX contamination”. Chapter 2 addresses specific objective 1 and involved the investigation of isobutanol on toluene degradation in six toluene degrading communities. The results indicated that isobutanol had varying effects on toluene biodegradation. A modified version of the work in Chapter 2 was published (Jayamani, I and A. M. Cupples. 2013. Effect of isobutanol on toluene biodegradation in nitrate amended, sulfate amended and methanogenic enrichment microcosms. Biodegradation, 24: 657-663). The study on ethylbenzene degradation under oxygen depleted conditions in agricultural soils is presented in Chapter 3. This study involved the investigation of ethylbenzene degradation in ~30 different experimental set-ups (different soils and redox conditions). From these, ethylbenzene degradation was noted only once. For this condition (nitrate reducing conditions in an 4 agricultural soil), SIP and amplicon sequencing were used to identify the dominant ethylbenzene degrading microorganisms. . Bacteria belonging to the families Oxalobacteraceae, Rhodospirillaceae, Xanthomonadaceae and Rhodocyclaceae and the genus Gemmatimonas were linked to the assimilation of the ring labeled carbon from ethylbenzene. The sequencing data from the total DNA extracts from the ethylbenzene degrading microcosms showed that phylotypes belonging to the genus Rhodanobacter were the most dominant microorganisms. The work presented under Chapter 3 has been submitted for publication to a peer reviewed journal. The research work presented in Chapters 4 and 5 were funded by a Strategic Environmental Research and Development Program (SERDP) grant entitled “Development of biomarkers for assessing in situ RDX biodegradation potential” awarded to Dr. Cupples. In Chapter 4, a comparative study of RDX degrading microbial communities obtained from four different uncontaminated soil sources is presented. Illumina sequencing revealed that phylotypes classifying either within the family Bacillaceae 1 (contains RDX isolates) or the genus Brevundimonas (no known RDX degrading isolates) were the most abundant phylotypes in all four soils. In addition, a SIP study was conducted with samples for two of the soils (soil 9 and soil 10). In soil 9, sequences classifying as Bacillaceae, Clostridiales and Tisserella were the most abundant in the heavy labeled fractions compared to the control unlabeled fractions. In soil 10, phylotypes belonging to Pusillimonas, Rhodococcus, Sedimentibacter were enriched in the heavy labeled fractions in comparison to the fractions obtained from the unlabeled samples. The work presented under this chapter is being prepared for publication. 5 Chapter 5 involves the use of SIP and high throughput sequencing to identify the microorganisms responsible for RDX degradation in soils obtained from an area used to test denotations (a Navy Base in Virginia, US). This work also explores the use of glucose as an amendment to promote RDX degradation. Members of the Pseudomonadaceae, Comamonas (with glucose samples) and Acinetobacter (no glucose samples) were observed as the most abundant phylotypes in the total DNA extracts. Members of the Pseudomonadaceae family were also linked to label uptake from RDX, both in the presences and absence of glucose. The work in Chapter 5 has been submitted for publication to a peer reviewed journal. 6 REFERENCES 7 REFERENCES 1. EPA 2014, posting date. Underground Storage Tanks. [Online.] 2. Pennington JC, Brannon JM. 2002. Environmental fate of explosives. Thermochimica Acta 384:163-172. 3. Dolfing J, Zeyer J, Bindereicher P, Schwarzenbach RP. 1990. Isolation and characterization of a bacterium that mineralizes toluene in the absence of molecularoxygen. Archives of Microbiology 154:336-341. 4. Schocher RJ, Seyfried B, Vazquez F, Zeyer J. 1991. Anaerobic degradation of toluene by pure cultures of denitrfying bacteria. Archives of Microbiology 157:7-12. 5. Evans P, Mang D, Kim KS, Young L. 1991. Anaerobic degradation of toluene by a denitrifying bacterium. Applied and Environmental Microbiology 57:1139-1145. 6. Chakraborty R, O'Connor SM, Chan E, Coates JD. 2005. Anaerobic degradation of benzene, toluene, ethylbenzene, and xylene compounds by Dechloromonas strain RCB. Applied and Environmental Microbiology 71:8649-8655. 7. Lovley DR, Lonergan DJ. 1990. Anaerobic oxidation of toluene, phenol, and paracresol by the dissimilatory iron-reducing organism, GS-15. Applied and Environmental Microbiology 56:1858-1864. 8. Kunapuli U, Jahn MK, Lueders T, Geyer R, Heipieper HJ, Meckenstock RU. 2010. Desulfitobacterium aromaticivorans sp nov and Geobacter toluenoxydans sp nov., ironreducing bacteria capable of anaerobic degradation of monoaromatic hydrocarbons. International Journal of Systematic and Evolutionary Microbiology 60:686-695. 9. Beller HR, Spormann AM, Sharma PK, Cole JR, Reinhard M. 1996. Isolation and characterization of a novel toluene-degrading, sulfate-reducing bacterium. Applied and Environmental Microbiology 62:1188-1196. 10. Schaefer CE, Yang XM, Pelz O, Tsao DT, Streger SH, Steffan RJ. 2010. Aerobic biodegradation of iso-butanol and ethanol and their relative effects on BTEX biodegradation in aquifer materials. Chemosphere 81:1104-1110. 11. Ding L, Cupples AM. 2014. The effect of the potential fuel additive isobutanol on benzene, toluene, ethylbenzene, and p-xylene degradation in aerobic soil microcosms. Environmental Technology:1-8. 12. Schaefer CE, Yang XM, Pelz O, Tsaob DT, Streger SH, Steffan RJ. 2010. Anaerobic biodegradation of iso-butanol and ethanol and their relative effects on BTEX biodegradation in aquifer materials. Chemosphere 81:1111-1117. 8 13. Weelink SAB, van Eekert MHA, Stams AJM. 2010. Degradation of BTEX by anaerobic bacteria: physiology and application. Rev Environ Sci Bio 9:359-385. 14. Rabus R, Widdel F. 1995. Anaerobic degradation of ethylbenzene and other aromatic hydrocarbons by new denitrifying bacteria. Arch Microbiol 163:96-103. 15. Ball HA, Johnson HA, Reinhard M, Spormann AM. 1996. Initial reactions in anaerobic ethylbenzene oxidation by a denitrifying bacterium, strain EB1. Journal of Bacteriology 178:5755-5761. 16. Wöhlbrand L, Kallerhoff B, Lange D, Hufnagel P, Thiermann J, Reinhardt R, Rabus R. 2007. Functional proteomic view of metabolic regulation in “Aromatoleum aromaticum” strain EbN1. Proteomics 7:2222-2239. 17. Coates JD, Chakraborty R, Lack JG, O'Connor SM, Cole KA, Bender KS, Achenbach LA. 2001. Anaerobic benzene oxidation coupled to nitrate reduction in pure culture by two strains of Dechloromonas. Nature 411:1039-1043. 18. Weelink SAB, van Doesburg W, Saia FT, Rijpstra WIC, Roling WFM, Smidt H, Stams AJM. 2009. A strictly anaerobic betaproteobacterium Georgfuchsia toluolica gen. nov., sp nov degrades aromatic compounds with Fe(III), Mn(IV) or nitrate as an electron acceptor. Fems Microbiol Ecol 70:575-585. 19. Kniemeyer O, Fischer T, Wilkes H, Glockner FO, Widdel F. 2003. Anaerobic degradation of ethylbenzene by a new type of marine sulfate-reducing bacterium. Applied and Environmental Microbiology 69:760-768. 20. Kwon MJ, Finneran KT. 2008. Biotransformation products and mineralization potential for hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) in abiotic versus biological degradation pathways with anthraquinone-2,6-disulfonate (AQDS) and Geobacter metallireducens. Biodegradation 19:705-715. 21. Kwon MJ, Finneran KT. 2008. Hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) and octahydro-1,3,5,7-tetranitro-1,3,5,7-tetrazocine (HMX) biodegradation kinetics amongst several Fe(III)-reducing genera. Soil & Sediment Contamination 17:189-203. 22. Kitts CL, Cunningham DP, Unkefer PJ. 1994. Isolation of three hexahydro-1,3,5trinitro-1,3,5-triazine-degrading species of the family Enterobacteriaceae from nitriamine explosive-contaminated soil. Applied and Environmental Microbiology 60:4608-4611. 23. Kitts CL, Green CE, Otley RA, Alvarez MA, Unkefer PJ. 2000. Type I nitroreductases in soil enterobacteria reduce TNT (2,4,6-trinitrotoluene) and RDX (hexahydro-1,3,5-trinitro-1,3,5-triazine). Canadian Journal of Microbiology 46:278-282. 24. Young DM, Unkefer PJ, Ogden KL. 1997. Biotransformation of hexahydro-1,3,5trinitro-1,3,5-triazine (RDX) by a prospective consortium and its most effective isolate Serratia marcescens. Biotechnology and Bioengineering 53:515-522. 9 25. Zhao JS, Halasz A, Paquet L, Beaulieu C, Hawari J. 2002. Biodegradation of hexahydro-1,3,5-trinitro-1,3,5-triazine and its mononitroso derivative hexahydro-1nitroso-3,5-dinitro-1,3,5-triazine by Klebsiella pneumoniae strain SCZ-1 isolated from an anaerobic sludge. Applied and Environmental Microbiology 68:5336-5341. 26. Zhao JS, Manno D, Beaulieu C, Paquet L, Hawari J. 2005. Shewanella sediminis sp nov., a novel Na+-requiring and hexahydro-1,3,5-trinitro-1,3,5-trinitro-degrading bacterium from marine sediment. International Journal of Systematic and Evolutionary Microbiology 55:1511-1520. 27. Zhao JS, Manno D, Leggiadro C, O'Neill D, Hawari J. 2006. Shewanella halifaxensis sp nov., a novel obligately respiratory and denitrifying psychrophile. International Journal of Systematic and Evolutionary Microbiology 56:205-212. 28. Zhao JS, Spain J, Thiboutot S, Ampleman G, Greer C, Hawari J. 2004. Phylogeny of cyclic nitramine-degrading psychrophilic bacteria in marine sediment and their potential role in the natural attenuation of explosives. Fems Microbiology Ecology 49:349-357. 29. Singh R, Soni P, Kumar P, Purohit S, Singh A. 2009. Biodegradation of high explosive production effluent containing RDX and HMX by denitrifying bacteria. World Journal of Microbiology & Biotechnology 25:269-275. 30. Binks PR, Nicklin S, Bruce NC. 1995. Degradatoion of hexahydro-1,3,5-trinitro-1,3,5triazine (RDX) BY Stenotrophomonas maltophilia Pb1. Applied and Environmental Microbiology 61:1318-1322. 31. Arnett CM, Adrian NR. 2009. Cosubstrate independent mineralization of hexahydro1,3,5-trinitro-1,3,5-triazine (RDX) by a Desulfovibrio species under anaerobic conditions. Biodegradation 20:15-26. 32. Boopathy R, Gurgas M, Ullian J, Manning JF. 1998. Metabolism of explosive compounds by sulfate-reducing bacteria. Current Microbiology 37:127-131. 33. Regan K, Crawford R. 1994. Characterization of Clostridium bifermentans and its biotransformation of 2, 4, 6-trinitrotoluene (TNT) and 1, 3, 5-triaza-1, 3, 5trinitrocyclohexane (RDX). Biotechnology Letters 16:1081-1086. 34. Zhang CL, Hughes JB. 2003. Biodegradation pathways of hexahydro-1,3,5-trinitro1,3,5-triazine (RDX) by Clostridium acetobutylicum cell-free extract. Chemosphere 50:665-671. 35. Zhao J-S, Paquet L, Halasz A, Hawari J. 2003. Metabolism of hexahydro-1, 3, 5trinitro-1, 3, 5-triazine through initial reduction to hexahydro-1-nitroso-3, 5-dinitro-1, 3, 5-triazine followed by denitration in Clostridium bifermentans HAW-1. Applied Microbiology and Biotechnology 63:187-193. 36. Bhushan B, Halasz A, Thiboutot S, Ampleman G, Hawari J. 2004. Chemotaxismediated biodegradation of cyclic nitramine explosives RDX9 HMX, and CL-20 by 10 Clostridium sp EDB2. Biochemical and Biophysical Research Communications 316:816821. 37. Adrian NR, Arnett CM. 2004. Anaerobic biodegradation of hexahydro-1,3,5-trinitro1,3,5-triazine (RDX) by Acetobacterium malicum strain HAAP-1 isolated from a methanogenic mixed culture. Current Microbiology 48:332-340. 38. Sherburne LA, Shrout JD, Alvarez PJJ. 2005. Hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) degradation by Acetobacterium paludosum. Biodegradation 16:539-547. 39. Radajewski S, McDonald IR, Murrell JC. 2003. Stable-isotope probing of nucleic acids: a window to the function of uncultured microorganisms. Current Opinion in Biotechnology 14:296-302. 40. Radajewski S, Ineson P, Parekh NR, Murrell JC. 2000. Stable-isotope probing as a tool in microbial ecology. Nature 403:646-649. 41. Whitby C, Bailey M, Whiteley A, Murrell C, Kilham K, Prosser J, Lappin-Scott H. 2005. Stable isotope probing links taxonomy with function in microbial communities Microbial ecologists have a culture-independent means for analyzing sources of metabolic activities among complex mixtures of microbes. Asm News 71:169-173. 42. DeRito CM, Pumphrey GM, Madsen EL. 2005. Use of field-based stable isotope probing to identify adapted populations and track carbon flow through a phenoldegrading soil microbial community. Applied and Environmental Microbiology 71:78587865. 43. Singleton DR, Powell SN, Sangaiah R, Gold A, Ball LM, Aitken MD. 2005. Stableisotope probing of bacteria capable of degrading salicylate, naphthalene, or phenanthrene in a Bioreactor treating contaminated soil. Applied and Environmental Microbiology 71:1202-1209. 44. Luo C, Xie S, Sun WM, Li X, Cupples AM. 2009. Identification of a novel toluenedegrading bacterium from the candidate phylum TM7, as determined by DNA stable isotope probing. Applied and Environmental Microbiology 75:4644-4647. 45. Sun W, Cupples AM. 2012. Diversity of five anaerobic toluene-degrading microbial communities investigated using stable isotope probing. Applied and environmental microbiology 78:972-980. 46. Xia Y, Cai L, Zhang T, Fang HHP. 2012. Effects of substrate loading and co-substrates on thermophilic anaerobic conversion of microcrystalline cellulose and microbial communities revealed using high-throughput sequencing. International Journal of Hydrogen Energy 37:13652-13659. 47. Bell TH, Yergeau E, Martineau C, Juck D, Whyte LG, Greer CW. 2011. Identification of nitrogen-incorporating Bacteria in petroleum-contaminated Arctic soils 11 by using [15N]DNA-based stable isotope probing and pyrosequencing. Applied and Environmental Microbiology 77:4163-4171. 48. Pilloni G, von Netzer F, Engel M, Lueders T. 2011. Electron acceptor-dependent identification of key anaerobic toluene degraders at a tar-oil-contaminated aquifer by Pyro-SIP. Fems Microbiology Ecology 78:165-175. 12 CHAPTER 2 EFFECT OF ISOBUTANOL ON TOLUENE BIODEGRADATION IN NITRATE AMENDED, SULFATE AMENDED AND METHONOGENIC ENRICHMENT MICROCOSMS ABSTRACT Isobutanol is an alternate fuel additive that is being considered because of economic and lower emission benefits. However, future gasoline spills could result in co-contamination of isobutanol with gasoline components such as benzene, toluene, ethyl-benzene and xylene (BTEX). Hence, isobutanol could affect the degradability of gasoline components thereby having an effect on contaminant plume length and half-life. In this study, the effect of isobutanol on the biodegradation of a model gasoline component (toluene) was examined in laboratory microcosms. For this, toluene and isobutanol were added to six different toluene degrading laboratory microcosms under sulfate amended, nitrate amended or methanogenic conditions. While toluene biodegradation was not greatly affected in the presence of isobutanol in five out of the six different experimental sets, toluene degradation was completely inhibited in one set of microcosms. This inhibition occurred in sulfate amended microcosms constructed with inocula from wastewater treatment plant activated sludge. Our data suggest that toluene degrading consortia are affected differently by isobutanol addition. These results indicate that, if cocontamination occurs, in some cases the in situ half-life of toluene could be significantly extended. 13 INTRODUCTION Various additives have been added to gasoline in an effort to minimize the emissions and to reduce non-renewable fuel dependency. Approximately 77% of all gasoline in the US contained alcohol up to 10% v/vol as of the year 2009 (1). Currently, biologically produced isobutanol is being considered as an alternative gasoline additive as it convenient to use with existing car engines and pipelines, can be used pure or at any blend ratio and also can be mixed offsite (2). A recent success in obtaining microbially produced isobutanol from cellulosic material such as corn stove waste by genetically engineered bacterium (3) adds to the benefits of switching to isobutanol as a gasoline additive. However, assuming that future spills are inevitable, it would be wise to have an understanding of the fate of isobutanol in the environment as well as any effect it might have on the biodegradation of other gasoline components. Previous research has revealed ethanol affects the fate of gasoline components in the environment. For example, due to its high hygroscopic nature, ethanol was found to produce a co-solvency effect resulting in increased BTEX aqueous concentrations (4, 5) and plume lengths (6). Ethanol is also preferentially degraded, causing depletion of oxygen, resulting in the inhibition of benzene degradation (7, 8). Further, ethanol biodegradation can consume subsurface nutrients and electron acceptors that would otherwise have been available for BTEX degradation (9). This in turn alters the composition of the local microbial population which could further result in reduced BTEX degradation (10). The concentration of ethanol also appears to affect BTEX degradation. At low ethanol concentrations, it was found to enhance BTEX biodegradation, in particular both aerobic and anaerobic benzene biodegradation (11, 12). In 14 contrast, at higher concentrations, ethanol decreased the degradation rates BTEX compounds (9, 11, 13). As ethanol exerts a high biochemical oxygen demand on the system, it is more relevant to discuss the effect of ethanol on BTEX biodegradation under anaerobic conditions (7). Corseiul et al.(1998) reported ethanol’s detrimental effect on toluene degradation under iron and methanogenic conditions, but enhanced toluene degradation under sulfate reducing conditions (14). Others have reported only slight toluene degradation under denitrifying conditions (15). In contrast, Da Silva et al. (2005) reported that though ethanol was inhibitory under methanogenic conditions, it enhanced BTEX biodegradation with the addition of electron acceptors (12). At higher ethanol concentrations most studies report slow or no BTEX biodegradation (9, 11, 12, 15-17). Similar contradicting results were obtained in biodegradation studies with n-butanol, an isomer of isobutanol. N-butanol while enhancing biodegradation of gasoline (18), was also found to be toxic to cell growth even at a concentration of 1% v/v (19). To date, only three studies (three aerobic and one anaerobic) have investigated the impact of isobutanol on BTEX biodegradation (20-22). Two of the studies used soil and groundwater samples collected from the same contaminated site in CA (20, 21) and found isobutanol had varying effects on BTEX degradation. The third study (22) used several agricultural soils degrading one or more of the BTEX compounds under aerobic conditions and concluded that isobutanol does not affect aerobic BTEX degradation. Here, we expand on this research by investigating the effect of isobutanol in six microbial communities derived from four different sources. The objective was to determine if the effect of isobutanol varied between samples from 15 different sources (i.e. different microbial communities). More specifically, the effect of isobutanol on toluene degradation was investigated under nitrate amended, sulfate amended and methanogenic conditions, using samples from different sources. Toluene was selected as the representative BTEX compound because previous work in our laboratory identified the toluene degrading species in a number of these communities using stable isotope probing (SIP) (23). Thus, this enabled us to link the effect of isobutanol of specific toluene degrading species. METHODS The effect of isobutanol on toluene degradation was investigated in seven different experiments prepared using inocula from enrichment cultures previously developed in our laboratory. Of these, toluene was degraded under either methanogenic (one experiment), sulfate amended (three experiments), or nitrate amended (three experiments) conditions (Table 2.1). Each experiment consisted of 12 sample (live) microcosms and 2 control (killed) microcosms. The microcosms for the methanogenic experiment were prepared using wastewater treatment plant digester sludge. The microcosms for the three nitrate amended experiments were prepared using either one of the three agricultural soils (corn crop) obtained from three different sources, two with low organic matter (agricultural soil 1 and 2) and one with high organic matter (agricultural soil 3). The microcosms for two of the sulfate amended experiments were prepared using wastewater treatment plant digester sludge and the microcosm for the third sulfate amended experiment was prepared using agricultural soil 1. A low concentration of isobutanol (5 mg/L) was chosen for six out of the seven experiments, following previous research (21). For the 16 seventh set (sulfate amended experiment prepared with agricultural soil 1) the isobutanol concentration was still lower (~2.5 mg/L). Microcosm preparation All microcosms were prepared under strictly anaerobic conditions in an anaerobic chamber (Coy Laboratory Products, Grass Lake, MI). The microcosms were prepared in 60 mL clear wheaton serum bottles (Fisher scientific, Pittsburgh, PA) by adding 0.5 to 1 mL of enrichment culture to 40 mL of basal minimal media (24). When appropriate, sterile potassium nitrate or magnesium sulfate solution were added (final concentration of 1 g/L) as electron acceptors. The microcosms were sealed tight with a butyl rubber stopper with PTFE coating and an aluminum seal to maintain anaerobic conditions. Seven experiments were performed to investigate the effect of isobutanol on toluene degradation (one methanogenic, three sulfate amended and three nitrate amended). For each experiment, 14 microcosms were prepared, including 2 abiotic controls (heat sterilized), 6 no isobutanol replicate microcosms and 6 replicate microcosms with isobutanol. The concentration of isobutanol in the isobutanol containing replicate microcosms, was ~5 mg/L in all experiments except one sulfate amended experiment prepared with wastewater sludge enrichment, which was prepared with lower isobutanol concentration(~ 2.5 mg/L). The rationale for this extra experiment is discussed later. The microcosms were stored at room temperature without shaking. Analytical methods Toluene and isobutanol concentrations were investigated over time in all microcosms. The timing of sample removal varied depending on preliminary data collected on toluene 17 degradation. Headspace toluene concentrations were measured using a Perkin Elmer gas chromatography equipped with a flame ionization detector (GC-FID). 200 µL headspace samples were sampled using a gas-tight syringe and injected through a DB-624 (Agilent) capillary column (length 30m, internal diameter 0.53mm and a film thickness of 0.5 um). Helium was used as the carrier and at 14 psi head pressure the elution time for ethylbenzene was 1.8 minutes. The aqueous phase toluene concentrations were calculated by applying Henry’s law. Aqueous isobutanol concentrations were estimated using headspace isobutanol concentrations and mass balance, using the same GC protocol. RESULTS AND DISCUSSION The effect of isobutanol on toluene degradation was examined in seven experiments involving microbial communities from four different sources. As expected, toluene was degraded in all live samples but was not significantly depleted in the controls, indicating a biological removal mechanism. The time for toluene depletion varied, primarily depending on the electron accepting process. While the varying levels of organic matter in the different soils could affect sorption of toluene and thereby its bioavailability and biodegradation rates, previous studies show that it also desorbs rapidly when biodegradation occurs (25). In microcosms amended with nitrate and constructed with the agricultural soil 1 enrichment, toluene degraded in just over 20 days in all 12 live samples (Figure 2.1a). There was no significant difference in the time for toluene degradation between the samples amended and those unamended with isobutanol. Toluene was degraded more rapidly (under 5 days) in the microcosms derived from agricultural soil 2 and soil 3 enrichments (Figures 2.1b and 2.1c). Again, no effect of isobutanol on toluene degradation was noted for these samples. 18 Toluene degraded at a much slower rate in sulfate amended experiments compared to the nitrate amended experiments. In sulfate amended microcosms constructed from agricultural soil 1 enrichments, (Figure 2.2a), toluene degradation was completed in all live samples after approximately 90 days. In these samples, toluene degradation rates varied between replicates and between the treatments, with no clear trend being apparent. In contrast, in the sulfate amended microcosms constructed from the digester sludge enrichments (Figure 2.2b and 2.2c), toluene degradation was completely inhibited in the presence of isobutanol. Also, among the two different sulfate amended experiments constructed from digester sludge enrichments, the time for toluene degradation was slower in those amended with a lower concentration of isobutanol (Figure 2.2b), which was likely because these were prepared at a later date and the enrichment culture was less active. Toluene degradation was also slow in the methanogenic microcosms constructed from wastewater digester sludge enrichments. In these microcosms, toluene degradation occurred between day 60 and 137. As no measurements were taken between these timepoints, it is difficult to conclude if isobutanol had an effect on toluene degradation during this time (Figure 2.3). Also in these microcosms, isobutanol was in fact completely degraded by day 34 (see below). Isobutanol degradation varied between the different experiments. Among the nitrate amended samples, it was slightly depleted in two of the three experiments to ~2-3 ppm (Figure 2.4a and 2.4c). In the third set of nitrate amended samples, the isobutanol concentration did not significantly change (Figure 2.4b). In the sulfate amended samples, isobutanol degradation was varied, taking between 21 and 118 days to degrade to ~1 ppm in the samples constructed with agricultural soil 1 enrichments (Figure 2.5a). In contrast, in the two experiments with the 19 activated sludge samples, isobutanol degradation was rapid at both the lower (Figure 2.5b) and higher (Figure 2.5c) isobutanol concentrations. In these samples, isobutanol was degraded by approximately day 15. In the methanogenic toluene degrading microcosms, isobutanol degraded rapidly and was completely depleted by day 34 (Figure 2.6). Our results show that isobutanol is completely biodegradable in both the sulfate and nitrate amended microcosms and in the methanogenic microcosms. Isobutanol degraded much faster and along with toluene in the nitrate amended microcosms compared to the sulfate amended microcosms. The presence of isobutanol had varying effects on toluene degradation in the six different microbial communities. In all three nitrate amended toluene degrading microbial consortia, the presence of isobutanol did not significantly alter the rate of toluene degradation. Schaefer et al (2010b) (21) reported that isobutanol actually increased the rate of BTEX biodegradation. They reported that isobutanol enhanced TEX biodegradation under nitrate reducing conditions and both ethanol and isobutanol enhanced benzene biodegradation under sulfate reducing conditions. Whereas, in the current study, in the sulfate amended samples derived from wastewater digester sludge, isobutanol actually inhibited toluene degradation. As previously mentioned, such mixed effects have been noted before for ethanol, where ethanol inhibited toluene degradation under methanogenic, iron reducing and nitrate reducing conditions, yet enhanced toluene degradation under sulfate reducing conditions (14). Schaefer et al (2010b) (21) also found that isobutanol could have a mixed effect on toluene degradation particularly under the presence of excess sulfate. Other studies have reported ethanol slowed TEX biodegradation rate in the absence of any electron acceptors (12, 21). These results show that the effect of isobutanol on BTEX biodegradation is likely not electron acceptor specific. 20 The current study found that the effect of isobutanol on toluene degradation varied between different toluene degrading communities. The major toluene degraders in three of the six inocula tested have been previously identified (23). The key degraders in agricultural soil 1 were identified as members of the family Comamonadaceae. In the microcosms derived from agricultural soil 1 enrichments, where isobutanol did not affect toluene biodegradation in sulfate amended microcosms, the major toluene degrader was identified as a member of the genus Desulfosporosinus. In the sulfate amended microcosms derived from digester sludge enrichments in which toluene degradation was inhibited in the presence of isobutanol the major toluene degrading organism was identified as a member of the family Syntrophobacteraceae. Site specific active microbial communities are often not considered as a input parameter in numerical models used to predict BTEX plume length (7). However, different microbial communities might be affected differently by isobutanol as seen here, and such numerical models may provide only partial information. Thus like ethanol the effect of isobutanol could be specific to the system in which contamination occurs (26). To our knowledge, this is the first study that has studied the effect of isobutanol on more than one toluene degrading consortia under different electron acceptor conditions. CONCLUSIONS Isobutanol illustrated a varying effect on toluene biodegradation in the different toluene degrading microbial communities. While the presence of isobutanol did not affect toluene degradation in five of the six experiments under nitrate reducing, sulfate reducing or methanogenic conditions, toluene degradation was inhibited in the digester sludge amended sulfate reducing microcosms. Also, in the experiments in which isobutanol did not affect toluene 21 degradation, there was no apparent difference in the rate of degradation, suggesting it neither enhanced nor inhibited toluene degradation in these microcosms. Our results suggest that the effect of isobutanol on BTEX biodegradation could differ based on the site specific conditions, particularly the microbial consortia and their assimilative capacity. Thus, this study supports the need for further site specific studies on the effect of isobutanol on BTEX contaminants to further understand the overall impact of switching to isobutanol as a gasoline additive. 22 APPENDIX 23 Table 2.1. Experimental setup for nitrate amended, sulfate amended and methanogenic toluene degrading samples. 1 2 3 4 5 6 7 Electron acceptor Methanogenic Nitrate amended Sulfate amended Enrichment culture source Wastewater treatment plant activated sludge Agricultural soil 1 Agricultural soil 2 Agricultural soil 3 Agricultural soil 1 Wastewater treatment plant activated sludge Wastewater treatment plant activated sludge No. of microcosms Abiotic Live (no isobutanol) 2 6 Live (with isobutanol) 6 Isobutanol concentration (mg/L) 5 2 2 2 2 2 6 6 6 6 6 6 6 6 6 6 5 5 5 5 5 2 6 6 2.5 24 25 (a) 20 15 Toluene concentration in solution (mg/L) 10 5 0 25 (b) 20 15 10 5 0 25 20 (c) 15 10 5 0 0 5 10 15 20 25 Time (Days) Figure 2.1. Toluene concentrations in nitrate amended microcosms derived from enrichments constructed from agricultural soil 1(a), agricultural soil 2 (b) and agricultural soil 3 (c). The symbols represent abiotic control [ ♦ ], toluene only [ ] and toluene and isobutanol [ amended microcosms. 25 ] 30 25 (a) 20 15 10 Toluene concentration in solution (mg/L) 5 0 30 25 20 (b) 15 10 5 0 30 25 20 (c) 15 10 5 0 0 20 40 60 80 Time (Days) 100 120 140 Figure 2.2. Toluene concentrations in sulfate amended microcosms derived from enrichments constructed from agricultural soil 1(a) wastewater treatment plant digester sludge with 2.5 ppm isobutanol (b), or with 5 ppm isobutanol (c). The symbols represent abiotic control [ ], toluene only [ ] and toluene and isobutanol [ amended microcosms. 26 ] Toluene concentration in solution (mg/L) 35 30 25 20 15 10 5 0 0 25 50 75 100 Time (Days) 125 150 Figure 2.3. Toluene concentrations in methanogenic microcosms derived from enrichments constructed from wastewater treatment plant digester sludge. The symbols represent abiotic control [ ], toluene only [ ] and toluene and isobutanol [ amended microcosms. 27 ] Isobutanol concentration in solution (mg/L) 16 14 12 10 8 6 4 2 0 16 14 12 10 8 6 4 2 0 16 (a) (b) (c) 14 12 10 8 6 4 2 0 0 5 10 15 20 25 Time (Days) Figure 2.4. Isobutanol concentrations in nitrate amended toluene degrading microcosms derived from enrichments constructed from agricultural soil 1 (a), agricultural soil 2 (b) and agricultural soil 3 (c). 28 12 10 8 (a) 6 Isobutanol concentration in solution (mg/L) 4 2 0 12 10 8 (b) 6 4 2 0 12 10 8 (c) 6 4 2 0 0 20 40 60 80 Time (Days) 100 120 140 Figure 2.5. Isobutanol concentrations in sulfate amended toluene degrading microcosms derived from enrichments constructed with agricultural soil 1 (a), and wastewater treatment plant digester sludge 2.5 ppm isobutanol (b) or 5 ppm isobutanol (c). 29 Isobutanol concentration in solution (mg/L) 8 7 6 5 4 3 2 1 0 0 25 50 75 100 125 150 Time (Days) Figure 2.6. Isobutanol concentrations in methanogenic toluene degrading microcosms derived from enrichments constructed with wastewater treatment plant digester sludge. 30 REFERENCES 31 REFERENCES 1. Weaver JW, Exum LR, Prieto LM. 2010. Gasoline Composition Regulations Affecting LUST Sites. In Laboratory NER (ed.). U.S. Environmental Protection Agency, Ecosystems Research Division, Athens, GA 30605. 2. Duerre P. 2007. Biobutanol: An attractive biofuel. Biotechnology Journal 2:1525-1534. 3. Higashide W, Li YC, Yang YF, Liao JC. 2011. Metabolic Engineering of Clostridium cellulolyticum for Production of Isobutanol from Cellulose. Applied and Environmental Microbiology 77:2727-2733. 4. Corseuil HX, Kaipper BIA, Fernandes M. 2004. Cosolvency effect in subsurface systems contaminated with petroleum hydrocarbons and ethanol. Water Research 38:1449-1456. 5. Heermann SE, Powers SE. 1998. Modeling the partitioning of BTEX in waterreformulated gasoline systems containing ethanol. Journal of Contaminant Hydrology 34:315-341. 6. Deeb RA, Sharp JO, Stocking A, McDonald S, West KA, Laugier M, Alvarez PJJ, Kavanaugh MC, Alvarez-Cohen L. 2002. Impact of ethanol on benzene plume lengths: Microbial and modeling studies. Journal of Environmental Engineering-Asce 128:868875. 7. Powers SE, Rice D, Dooher B, Alvarez PJJ. 2001. Will ethanol-blended gasoline affect groundwater quality? Using ethanol instead of MTBE as a gasoline oxygenate could be less harmful to the environment. Environmental Science & Technology 35:24A-30A. 8. Ruiz-Aguilar GML, O'Reilly K, Alvarez PJJ. 2003. A comparison of benzene and toluene plume lengths for sites contaminated with regular vs. ethanol-amended gasoline. Ground Water Monitoring and Remediation 23:48-53. 9. Da Silva MLB, Alvarez PJJ. 2002. Effects of ethanol versus MTBE on benzene, toluene, ethylbenzene, and xylene natural attenuation in aquifer columns. Journal of Environmental Engineering-Asce 128:862-867. 10. Feris K, Mackay D, de Sieyes N, Chakraborty I, Einarson M, Hristova K, Scow K. 2008. Effect of ethanol on microbial community structure and function during natural attenuation of benzene, toluene, and o-xylene in a sulfate-reducing aquifer. Environmental Science & Technology 42:2289-2294. 11. Lovanh N, Hunt CS, Alvarez PJJ. 2002. Effect of ethanol on BTEX biodegradation kinetics: aerobic continuous culture experiments. Water Research 36:3739-3746. 32 12. Da Silva MLB, Ruiz-Aguilar GML, Alvarez PJJ. 2005. Enhanced anaerobic biodegradation of BTEX-ethanol mixtures in aquifer columns amended with sulfate, chelated ferric iron or nitrate. Biodegradation 16:105-114. 13. Mackay DM, De Sieyes NR, Einarson MD, Feris KP, Pappas AA, Wood IA, Jacobson L, Justice LG, Noske MN, Scow KM, Wilson JT. 2006. Impact of ethanol on the natural attenuation of benzene, toluene, and o-xylene in a normally sulfate-reducing aquifer. Environmental Science & Technology 40:6123-6130. 14. Corseuil HX, Hunt CS, Dos Santos RCF, Alvarez PJJ. 1998. The influence of the gasoline oxygenate ethanol on aerobic and anaerobic BTX biodegradation. Water Research 32:2065-2072. 15. Chen YD, Barker JF, Gui L. 2008. A strategy for aromatic hydrocarbon bioremediation under anaerobic conditions and the impacts of ethanol: A microcosm study. Journal of Contaminant Hydrology 96:17-31. 16. Osterreicher-Cunha P, Guimaraes JRD, Vargas ED, da Silva MIP. 2007. Study of biodegradation processes of BTEX-ethanol mixture in tropical soil. Water Air and Soil Pollution 181:303-317. 17. Wu YG, Li YF, Hui L, Tan Y, Jin S. 2009. Effects of Ethanol on Benzene Degradation Under Denitrifying Conditions. Bulletin of Environmental Contamination and Toxicology 82:145-152. 18. Mariano AP, Tomasella RC, Di Martino C, Maciel Filho R, Regali Seleghim MH, Contiero J, de Angelis DdF. 2009. Aerobic biodegradation of butanol and gasoline blends. Biomass & Bioenergy 33:1175-1181. 19. Mariano AP, Tomasella RC, Di Martino C, Morais EB, Maciel Filho R, Regali Seleghim MH, Contiero J, Tauk Tornisielo SM, de Angelis DdF. 2010. Aerobic biodegradation of butanol and diesel oil blends. African Journal of Biotechnology 9:7094-7101. 20. Schaefer CE, Yang XM, Pelz O, Tsao DT, Streger SH, Steffan RJ. 2010. Aerobic biodegradation of iso-butanol and ethanol and their relative effects on BTEX biodegradation in aquifer materials. Chemosphere 81:1104-1110. 21. Schaefer CE, Yang XM, Pelz O, Tsaob DT, Streger SH, Steffan RJ. 2010. Anaerobic biodegradation of iso-butanol and ethanol and their relative effects on BTEX biodegradation in aquifer materials. Chemosphere 81:1111-1117. 22. Ding L, Cupples AM. 2014. The effect of the potential fuel additive isobutanol on benzene, toluene, ethylbenzene, and p-xylene degradation in aerobic soil microcosms. Environmental Technology:1-8. 33 23. Sun W, Cupples AM. 2012. Diversity of five anaerobic toluene-degrading microbial communities investigated using stable isotope probing. Applied and environmental microbiology 78:972-980. 24. Yang YR, Zeyer J. 2003. Specific detection of Dehalococcoides species by fluorescence in situ hybridization with 16S rRNA-targeted oligonucleotide probes. Applied and Environmental Microbiology 69:2879-2883. 25. Robinson KG, Farmer WS, Novak JT. 1990. Availability of sorbed toluene in soils for biodegradation by acclimated bacteria. Water Research 24:345-350. 26. Alvarez PJJ, Hunt CS. 2002. The effect of fuel alcohol on monoaromatic hydrocarbon biodegradation and natural attenuation. Rev Latinoam Microbiol 44:83-104. 34 CHAPTER 3 STABLE ISOTOPE PROBING AND HIGH THROUGHPUT SEQUENCING IMPLICATES OXALOBACTERACEAE, RHODOSPIRILLACEAE, XANTHOMONADACEAE AND RHODOCYCLACEAE IN ETHYLBENZENE DEGRADATION ABSTRACT The overall aim was to identify the microorganisms responsible for carbon uptake from ethylbenzene (or ethylbenzene breakdown products) in soil microcosms using culture independent methods. Ethylbenzene degradation was observed in only one (nitrate amended agricultural soil) of the thirty-one different experimental setups (different amendments and inocula sources) tested. To determine which microorganisms were responsible for 13C uptake, stable isotope probing (SIP) was combined with high throughput sequencing Illumina amplicon sequencing. For this, total genomic DNA and heavy fraction samples (following ultracentrifugation to separate light from heavy DNA) from the live control microcosms (amended with 12C6 ethylbenzene) and sample microcosms (amended with 13C6 ethylbenzene) were subject to Illumina amplicon sequencing. Several phylotypes were relatively more abundant in the heavy fractions from the 13C6-labeled ethylbenzene amended soil microcosms compared to the heavy fractions from the live controls (from unlabeled ethylbenzene amended microcosms). This included unclassified phylotypes within the families Oxalobacteraceae, Rhodospirillaceae, Xanthomonadaceae and Rhodocyclaceae (Proteobacteria) as well as the genus Gemmatimonas (phylum Gemmatimonadetes). In addition, the total microbial community was characterized in six ethylbenzene degrading microcosms and was dominated by the genus Rhodanobacter 35 (relative abundance 11.7 – 15.6%) (family Xanthomonadaceae), although this phylotype did not illustrate 13C uptake. Azoarcus (family Rhodocyclaceae), a known ethylbenzene degrader, was dominant in only two of the six microcosms (relative abundance 14.6 and 24.1%) and it also did not illustrate label uptake. The work indicates microorganisms not previously linked to ethylbenzene degradation have significant roles in the carbon uptake from this contaminant. INTRODUCTION Hydrocarbons such as benzene, toluene, ethylbenzene and xylene (BTEX) contained in gasoline have been released into the environment by accidental spillage or from leaking underground storage tanks. These compounds cause groundwater and sediment pollution posing an environmental and health threat due to their toxicity and mobility. BTEX remediation using microorganisms that can utilize these compounds as a carbon or an energy source is a feasible solution. However, under anaerobic conditions, BTEX biological removal is typically slow. From the BTEX chemicals, toluene is the most studied and perhaps the most readily amenable to anaerobic biodegradation (1). Several microorganisms have been isolated that are able to metabolize or mineralize toluene under nitrate reducing (2, 3), iron reducing (4) and sulfate reducing (5) conditions (the majority belong to the phylum Proteobacteria and a few to Firmicutes). Isolates have also been shown to degrade m- or o-xylene under nitrate reducing (2), sulfate reducing (6) and iron reducing (7) conditions. In contrast, there have been reports of only three nitrate reducing, benzene degrading isolates (8, 9). After benzene, ethylbenzene is perhaps the next most difficult to degrade among the BTEX compounds. To date, anaerobic ethylbenzene degradation has been shown in several nitrate reducing or sulfate reducing bacteria (1). These 36 include two Azoarcus sp. strains (Aromatoleum aromaticum EbN1 and EB1) (10-12) , Dechloromonas aromatica RCB (8, 13), Georgfuchsia toluolica (14) and Desulfobacteraceae Strain EbS7 (15). Interestingly, all except Strain EbS7 (Deltaproteobacteria) belong to the family Rhodocyclaceae (Betaproteobacteria). More information concerning the diversity of microorganisms involved in ethylbenzene degradation would contribute to our understanding of the biodegradation potential of this chemical at contaminated sites. The current limited information on ethylbenzene degraders can, in part, be attributed to the limitations associated with traditional microbiological methods (only a minor % of existing bacteria can be cultured in the laboratory). Culture independent methods, such as stable isotope probing (SIP), can often circumvent these limitations. For example, SIP studies have identified novel microorganisms involved in geochemical cycles (16, 17) and in remediation of environmental pollutants (18-21) that were not previously identified using culture based techniques. SIP involves the addition of a labeled substrate (e.g. 13C labeled toluene) to a mixed community, ultracentrifugation of the extracted DNA, then the analysis of the heavy fractions of the SIP gradient to identify the phylotypes responsible for label uptake. Previously, we have used terminal restriction fragment length polymorphism (TRFLP) and clone libraries (with traditional Sanger sequencing) to identify the organisms enriched in the labeled treatments during SIP experiments (22-27). However, extensive 16S rRNA gene clone libraries are needed to identify the enriched TRFLP fragments. In fact, researchers have previously reported they were unable to identify 25% (two out of eight) TRFLP fragments thought to be responsible for label uptake (28). Replacing TRFLP and clone libraries with next generation sequencing has the potential to provide a greater depth of information on the occurrence of specific phylotypes in SIP gradients. 37 For example, in 2011, researchers have used pyrosequencing on total DNA extracts in an SIP study to investigate the microorganisms involved in toluene degradation at a tar-oil-contaminated aquifer (29). Others have integrated pyrosequencing libraries with SIP identified sequences to determine the abundance of microorganisms in samples from the Deepwater Horizon soil spill (30) or to investigate ammonia oxidizers (31). Further, researchers have also used pyrosequencing on the SIP fractions (rather than on total DNA extracts) to, for example, investigate nitrogen-incorporating bacteria in petroleum contaminated artic soils (32), to study organic carbon uptake by bacterioplankton (33), to investigate ammonia and nitrite oxidation (31, 34-36), to study nonylphenol degraders in activated sludge (37), to study the gut microbiota of cotton leafworm (Spodoptera littoralis) (38) to investigate toluene degradation (39) and to study biological soil crusts (40). In the current study, the overall aim was to identify microorganisms able to uptake 13C from ethylbenzene (or ethylbenzene degradation products) in soil based microcosms, using culture independent methods (SIP and high throughput sequencing). For this, the relative abundance of phylotypes present in the heavy fractions of two labeled and two unlabeled ethylbenzene amended microcosms from one soil were compared. The phylotypes with a high relative abundance in the heavy fractions of the labeled ethylbenzene amended microcosm (compared to similar fractions in the unlabeled amended microcosms) were deemed responsible for incorporating the labeled carbon from ethylbenzene (or ethylbenzene degradation products). An underlying hypothesis of this work was that these techniques would produce a more diverse group of ethylbenzene degraders than those previously reported. 38 METHODS Chemicals Unlabeled and labeled ethylbenzene (Phenyl -13C6 , 99% ) were purchased from Sigma-Aldrich (St. Louis, MO, USA). Other reagents were purchased from one or more of the following vendors: Fisher Bioreagent (Thermo Fisher Scientific, NJ, USA), Invitrogen (Life Technologies, Grand Island, NY, USA) and Sigma-Aldrich (St.Louis, MO, USA), unless stated otherwise. Development of Ethylbenzene Degrading Microcosms A wide variety of inocula and electron acceptor conditions were tested for ethylbenzene degradation using microcosms prepared under strict anaerobic conditions in an anaerobic chamber (Coy Laboratory Products, Grass Lake, MI). The inocula sources consisted of 12 different agricultural soil samples (collected from Michigan), activated sludge samples (from two wastewater treatment plants) and digested manure samples (from an anaerobic digester at Michigan State University). These sources were selected because previous research indicated BTEX degradation potential (23-27). The inocula sources were tested under nitrate amended, sulfate amended, iron amended and/or methanogenic conditions in the presence of an anaerobic minimal basal media (41). For this, triplicate microcosms were prepared with 6 to 10 g (wet weight) of the inocula source, 30 mL of the anaerobic basal media and 20 mg L-1 of unlabeled ethylbenzene in a 160 mL serum bottle. Electron acceptors were amended (1 g L-1 of SO42- or NO3- or Fe3+) as required. Microcosms were sealed using rubber stoppers and aluminum seals in the anaerobic chamber. All incubations were monitored for ethylbenzene removal for up to 8 weeks (data not shown). Among the 31 different incubations, only the triplicate microcosms prepared with an agricultural soil inocula amended with nitrate showed ethylbenzene 39 degradation. This soil was used to inoculate the microcosms for the SIP study (prepared as above). For the SIP study, triplicate killed controls and four microcosms were prepared with 20 mg L-1 of unlabeled ethylbenzene (hereafter, called Unlabel EB 1, Unlabel EB 2, SIP Unlabel EB 1 and SIP Unlabel EB 2) and two live sample microcosms were prepared with 20 mg L-1 of labeled ethylbenzene (phenyl -13C6 , 99% ) (hereafter, called SIP Label EB 1 and SIP Label EB 2). From these, DNA extracted from SIP Unlabel EB 1, SIP Unlabel EB 2, SIP Label EB 1 and SIP Label EB 2 was used for SIP analysis (ultracentrifugation and fractionation) and total microbial community analysis. Whereas, DNA extracted Unlabel EB 1 and Unlabel EB 2 was only used for microbial community analysis. This allowed the comparison of the microbial community across six microcosms. All microcosms were incubated at room temperature (~20 °C), without shaking. Analytical Techniques Ethylbenzene biodegradation was monitored over time by measuring headspace ethylbenzene concentrations in the microcosms. Sampling days for the SIP microcosms were determined based on the results of the initial screening microcosms. For this, 200 µL of headspace samples were injected into a gas chromatograph (Perkin Elmer) attached to a flame ionization detector and an AT-624 capillary column (J & W Scientific; 30 m, 0.53 mm ID, 3.0 µm) with helium as carrier gas. The oven temperature was set to 150 °C while the injector port and detector temperature were set to 200 °C with an ethylbenzene elution time of 2.3 minutes. Nitrite and nitrate measurements were made only on day 1 and 22 using ion chromatography. 40 DNA Extraction and Isopycnic Centrifugation Total genomic DNA was extracted from the six microcosms (two labeled and four unlabeled ethylbenzene amended) using the Power Soil DNA extraction kit following the manufacturer’s instruction (MO BIO Laboratories, Inc. Carlsbad, CA) following the degradation of ethylbenzene. The extracted DNA from SIP Unlabel EB 1, SIP Unlabel EB 2, SIP Label EB 1 and SIP Label EB 2 was ultracentrifuged in cesium chloride gradients separately to obtain density-resolved gradients and fractions. Approximately 10 µg of total genomic DNA was mixed with a Tris-EDTA (pH 8.0) buffer and CsCl solution (no glycogen was added to the gradient). This mixture was added to a 5.1 mL Quick-Seal polyallomer tubes (1.3 x 5.1 cm, Beckman Coulter) the buoyant density (BD) of this mixture was adjusted to ~1.72 g mL-1 using a model AR200 digital refractometer (Leica Microsystems Inc.) and then sealed using a tube topper (Cordless quick-seal tube topper, Beckman). The tubes were then centrifuged at 178,000 x g for 46 hours at 20 °C in a Wx Sorvall Ultra 80 ultracentrifuge fitted with a Stepsaver 70 V6 Vertical Titanium Rotor (Thermo Scientific). The samples were separated into 20 fractions (each 250 µL) by displacing the samples with molecular grade water. A syringe pump attached to a needle (BD, 23G and 1 inch) was used to displace samples from the top of the tube, as fractions were collected from the bottom by gravity, from heavy to light fractions. The BD of each fraction was calculated from the refractive index obtained using a refractometer. DNA from each of the fraction was recovered using a glycogen and ethanol precipitation. Precipitated DNA was then re-suspended in 30 µL PCR grade water and stored at -20 °C for further analysis. The DNA concentration in was quantified in these fractions using replicate measurements with the Qubit assay. 41 High Throughput Amplicon Sequencing (Illumina MiSeq) Samples were subject to amplicon sequencing, following the protocol described elsewhere (42, 43) (Research Technology Support Facility or RTSF at Michigan State University). The fractions were analyzed to determine which microorganisms were enriched in the heavy fractions and were therefore responsible for carbon uptake from ethylbenzene (or ethylbenzene degradation products). Also, the total DNA samples were analyzed with high throughput sequencing to characterize the total microbial community present. The analysis included four heavy fractions from SIP Unlabel EB 1, SIP Unlabel EB 2, SIP Label EB 1 and SIP Label EB 2. To enable a comparison between fractions, the BD of the selected fractions from the labeled and unlabeled amended microcosms were similar (maximum difference ±0.01 g mL-1) (Table 3.1). These fractions were selected because their BD values were similar to those previously reported by our group for SIP label enrichment (22-27). In addition, the fractions were selected based on the concentration of DNA in each fraction. Illumina specific fusion primers were used to amplify the V4 region of the 16S rRNA (42) and to add unique barcodes to samples in each well to enable pooling and sequencing. After the amplicons were checked on 1% agarose gel, equimolar amounts of the sample were pooled to normalize results, purified and then sequenced on the Illumina MiSeq™ Personal Sequencing System. The data obtained from amplicon sequencing in the fastq file format was analyzed using Mothur (version 1.32.1) developed by Schloss (44) and the MiSeq standard operating procedure (SOP) developed by the same laboratory(45, 46). Briefly, barcodes were removed from the sequence data and contigs were made using the forward and reverse reads. Samples from one labeled and one unlabeled replica each containing 4 fractions and one total genomic DNA sample (thus a 42 total of 10 samples) were analyzed together. The paired-end reads were used to make contigs, were analyzed for errors and then classified. Sequences that did not assemble well were removed from further analysis. Reads were checked for ambiguous bases and homopolymer length greater than 8 (following the Mothur SOP) to remove sequencing errors. These sequences were then aligned with the SILVA bacteria database (47) for the V4 region. Chimeras, mitochondrial and chloroplast lineage sequences were removed and then the sequences were classified into OTU’s. The OTUs were then grouped into taxonomical levels with corresponding confidence levels. Rarefaction curves were generated using Mothur. Cloning and Sanger Sequencing A 16S rRNA gene clone library was constructed by amplifying, cloning and sequencing the total genomic DNA. The 16S rRNA gene was amplified using the universal primers: 27F 5’AGAGTTTGATCMTGGCTCAG) and 1492R (5’-GGTTACCTTGTTACGACTT) (Operon Biotechnologies) using the following cycle program: initial denaturation step (94 °C for 5 min), 30 cycles of amplification (94 °C for 30 sec, 55 °C for 30 sec, 72 °C for 1.5 min) and final extension step (72 °C for 5 min). The amplicons were run on a 1% agarose gel and were then cut and purified using a gel purification kit (Qiagen). The purified amplicons were then cloned using the TOPO TA kit as per manufacturer’s instruction. The cloned Escherichia coli cells were then plated on agar plates with 50 µg mL-1 ampicillin and grown over night. The colonies were screened for inserts using blue-white screening and PCR (M13F 5’TGTAAAACGACGGCCAGT-3’ and M13R 5’-AACAGCTATGACCATG-3’). The positive cells were submitted to RTSF for Sanger sequencing. The sequences obtained were then classified using the Ribosomal Database Project (RDP) “Classifier” (48) . 43 RESULTS Ethylbenzene Biodegradation Ethylbenzene degradation was observed in only one of the 31 different incubations over an 8 week period. The lack of ethylbenzene degradation may be a result of the rare occurrence of ethylbenzene degrading microorganisms (however, no molecular analysis of those samples not illustrating ethylbenzene degradation was performed). In the microcosms amended with agricultural soil and nitrate, >90% of the added ethylbenzene was degraded in 29 days in the sample microcosms, with no significant degradation in the killed abiotic controls (data not shown). There was an initial decrease of approximately 5 mg L-1 in all of the microcosms (live and killed) between day 1 and day 7. This was possibly due to sorption as no further decrease was noted in the killed control microcosms. Similar degradation rates were observed in the SIP microcosms prepared with this inoculum (Figure 3.1). However, replicate microcosms exhibited slightly different rates of ethylbenzene degradation. No significant reductions in ethylbenzene concentrations were observed in the abiotic killed controls (after the initial sorption decrease in all microcosms). An initial lag in degradation was observed in all microcosms and following this, most degradation occurred in 7 to 10 days. There was also an 80% decrease in the nitrate concentration between day 1 and day 22 in the labeled and unlabeled samples with no significant decrease in the abiotic controls (data not shown). High Throughput Amplicon Sequencing Analysis Samples from fractions (Table 3.1) as well as total DNA were submitted for high throughput amplicon sequencing. The fractions selected were at the heavy end of the ultracentrifugation 44 gradient (Figure 3.2). Approximately 90,000 final sequences were obtained per sample. Sequencing data were obtained for both the fractions as well as total DNA. The number of sequences obtained per run and the final number of sequences used for the OTU and phylotypes classification are listed (Table 3.2). Sequences that did not assemble well (>275 bp) were dropped from further analysis. Further, an average of approximately 3.3 % of the remaining sequences were also removed as they were either marked as chimeric by the Uchime algorithm (49) or as belonging to mitochondria or chloroplast lineage by the Bayesian classifier. The remaining sequences were used for analysis. These sequences were classified into OTU’s and taxons by splitting them into bins and clustering them at the order level within each bin at 97% similarity cutoff level. Sequences of clones from this study were submitted to Genbank under accession number KM362736-KM362822. The high throughput sequence data was submitted to SRA under Bioproject number PRJNA257333. Microbial Community in Ethylbenzene Degrading Microcosms Rarefaction curves indicated an adequate level of sequence coverage for the total community analysis of the six microcosms (Figure 3.3). As expected, no differences were noted between the coverage in the labeled compared to the unlabeled ethylbenzene amended microcosms. The OTU classification data of the total DNA samples into phylotypes was used to determine the community composition in each of the six microcosms. Figure 3.4 shows the phyla level classification of the sequences in the total genomic DNA samples from the six microcosms. A small percentage (4.3-8%) of the sequences were unclassified Bacteria. Proteobacteria was the most abundant phyla in all six microcosms (relative abundance of 58.7-79.1%). The remaining dominant phyla included Firmicutes (6.3-13.17%), Actinobacteria (2.9-8.4 %), Acidobacteria 45 (2.0-13.3%), and Bacteroidetes (0.8-4.6%). Microorganisms from both phyla Gemmatimonadetes and Verrucomicrobia were present in all six microcosms at levels between 0.9 and 1.9 %. Microorganisms in the phyla TM7, Planctomycetes, Chloroflexi, DeinococcusThermus, Armatimonadetes, Chlamydiae and Nitrospira were present on all six microcosms, but only at lower levels (0.007-0.3%). Whereas, microorganisms classifying within BRC1, OD1, WS3, Chlorobi, Lentisphaerae were present in some microcosms, but not in others (when present, <0.006%). The relative abundance of all phylotypes with a relative abundance of > 1 % were compared across all six microcosms (Figure 3.5). This resulted in a comparison of fifteen phylotypes. In four of the six microcosms, Rhodanobacter (Gammaproteobacteria) was the most abundant phylotype (14.1-15.6%) and in the other two microcosms it was the second (14.2%) or third (11.7%) most abundant phylotype. In two microcosms, Azoarcus was the most abundant phylotype (24.1 and 14.6%), however, in the other four microcosms it was present at very low levels (<0.006 %). In four microcosms, unclassified Rhodospirillaceae was the second most dominant phylotype (10.7-14.1%) a family within the Alphaproteobacteria. In the other two microcosms, it was present at lower levels (2.7 and 5.1%). In all six microcosms, unclassified Oxalobacteraceae (Betaproteobacteria) was present at high levels (5.5-14.1%). Unclassified Xanthomonadaceae (the family containing Rhodanobacter) was also present in all six microcosms, but at lower levels (1.9-6.3%). Other unclassified families occurring in all six microcosms at lower levels include Bacillaceae 1 (Firmicutes), Chitinophagaceae (Bacteroidetes) and Comamonadaceae (Betaproteobacteria). Three unclassified orders (Bacillales, Rhodospirillales and Rhizobiales), unclassified Bacteria and unclassified 46 Betaproteobacteria were also present in all six microcosms. In addition, two phylotypes classifying as Thiobacillus (Betaproteobacteria) and Gemmatimonas (phylum Gemmatimonadetes) were found in all six microcosms. 16S rRNA Gene Clone Library A clone library was generated so that a comparison could be performed between traditional Sanger sequencing and high throughput sequencing data. The clone library consisted of a 120 clones (Table 3.3) belonging to eight different phyla: Proteobacteria (73.1%), Firmicutes (10.9%), Bacteriodetes (8.4%), Acidobacteria (5.0%), Armatimonadetes (0.8%), Gemmatimonadetes (0.8%), TM7 (0.8%), and Verrucomicrobia (0.8%). Among the Proteobacteria, the majority belonged to Betaproteobacteria (45.8% of the total). The major phyla present in the clone library were relatively consistent with the high throughput sequencing results. That is, the Illumina results indicated the communities in the six microcosms were also dominated by Proteobacteria (58.7-79.1%), followed by Firmicutes (6.3-13.17%), Acidobacteria (2.0-13.3%), and Bacteroidetes (0.8-4.6%). Notably, no clones belonging to Actinobacteria were present in the clone library, however, this phylum exhibited a relative abundance of between 2.9 and 8.4 % in high throughput sequencing results. Further, the clone library did not contain representatives from 11 additional phyla that were found in the high throughput sequencing data. Consistent with the high throughput sequencing data, the clone library did contain a high number of Rhodanobacter (relative abundance of 10.0%) and unclassified Oxalobacteraceae sequences (25.0%). In contrast, the number of sequences of unclassified Rhodospirillaceae was much lower in the library (3.3%) compared to the Illumina data. Azoarcus was not present in the clone library, however, Gemmatimonas was present (0.83%). As expected, the Illumina results 47 provided a more in depth representation of the microorganisms present in these ethylbenzene degrading communities. Phylotype Relative Abundance Across SIP Fractions The relative abundance of phylotypes in heavy fractions were compared between the labelled and unlabeled ethylbenzene amended microcosms to determine which phylotypes could be linked to label uptake. For this, the analysis focused on the most abundant phylotypes in fractions from the labeled ethylbenzene amended microcosms. The relative abundance of these phylotypes was then compared to their relative abundance in fractions from the unlabeled ethylbenzene amended microcosms. Overall, seven phylotypes were dominant in the heavy fractions from both of the labeled ethylbenzene amended microcosms (Figure 3.6). Rhodanobacter was the most abundant in these fractions, however, the relative abundance was similar in fractions from the labeled and unlabeled ethylbenzene amended samples, indicating no label uptake. Following this, unclassified Oxalobacteraceae and Rhodospirillaceae were also dominant. A comparison of the relative abundance of these phylotypes indicates label uptake likely occurred in one labeled ethylbenzene amended microcosm (SIP Label EB 2), but not the other (SIP Label EB 1). Unclassified Xanthomonadaceae and Rhodocyclaceae both illustrated a higher relative abundance in the fractions from the labeled ethylbenzene amended microcosms compared to the unlabeled amended microcosms. No difference in relative abundance was noted for unclassified Bacillales between treatments. Finally, the phylotype Gemmatimonas was more abundant in fractions from one labeled ethylbenzene amended microcosm (SIP Label EB 1). Azoarcus was not dominant in the heavy fractions of either labeled ethylbenzene amended microcosm. 48 DISCUSSION Biodegradation studies using SIP aim to identify active organisms involved in chemical transformations that are typically not identified using culture based techniques (21, 50, 51). Combining SIP with high throughput sequencing has the added advantage of a greater depth of analysis (compared to Sanger sequencing). Although Illumina amplicon sequencing produces shorter reads even with paired reads (250 bp) compared to Sanger sequencing (up to ~1000 bp) (52, 53), previous studies have shown the method can still resolve the differences between microbial communities (42). The data generated in the current study indicated that one clone library (120 clones) with Sanger sequencing provided a weaker representation of the microbial community compared to the Illumina data. For example, Actinobacteria, a major phylum in the Illumina data, was absent from the clone library, as were 11 other phyla (that were present in the Illumina data). The contrasting results from these two sequencing methods are consistent with previous studies that used both techniques. The large amount of sequencing information from amplicon sequencing is an improved predictor of phylotype relative abundance (29, 32, 35) and thus has great potential for SIP studies. Although Rhodanobacter was the most abundant phylotype in the majority of the six microcosms, it’s relative abundance was similar in the fractions obtained from the labelled and unlabeled ethylbenzene amended microcosms, indicating it was not responsible for label uptake. Phylotypes from the families Oxalobacteraceae and Rhodospirillaceae appeared to be involved in carbon uptake from ethylbenzene in one microcosm (SIP Label EB 2). According to the Ribosomal Database Project (http://rdp.cme.msu.edu/index.jsp), Oxalobacteraceae (order Burkholderiales, class Betaproteobacteria) contains 11 genera and Rhodospirillaceae (order 49 Rhodospirillales, class Alphaproteobacteria) contains 27 genera. Any of these genera could be responsible for label uptake in this microcosm. The phylotype Gemmatimonas was more abundant in the fractions from one labeled ethylbenzene amended microcosm (SIP Label EB 1). This microorganism belongs to the phylum Gemmatimonadetes, which only contains one genus (with 15 known isolates), with the first isolate being obtained only relatively recently (2003) (54). The family Xanthomonadaceae (order Xanthomonadales, class Gammaproteobacteria) contains 21 genera and phylotypes within this family appeared to responsible for label uptake in both label amended microcosms. Again, any of these genera could be responsible for label uptake. Phylotypes within the family Rhodocyclaceae were also more abundant in the fractions from both the labeled ethylbenzene amended microcosms compared to similar fractions from the unlabeled ethylbenzene amended microcosms. This pattern indicates phylotypes classifying as Rhodocyclaceae were involved in carbon uptake from ethylbenzene (or ethylbenzene degradation products). These results are consistent with pure culture studies, as the majority of anaerobic ethylbenzene degrading isolates belong to this family. Previous research has indicated phylotypes classifying as Rhodocyclaceae (Azoarcus, Dechloromonas, Georgfuchsia spp.) are capable of ethylbenzene degradation (8, 10, 11, 14). Also, previous studies have linked organisms from this family to transformation of several organic and inorganic substances such as phenantharene (55-57), naphthalene (58), other petroleum products (32), arsenic (59), uranium (60), chromium (61) and alkanes (62). Members of this family are also shown to degrade almost all of the BTEX compounds, particularly under nitrate reducing conditions (8-11, 15, 29, 63, 64). These findings indicate these microorganisms play an important role in monoaromatic 50 hydrocarbon degradation in isolation as well as in mixed communities. Also, two ethylbenzene degrading strains, Aromatoleum aromaticum EbN1 and Dechloromonas aromatica RCB, were isolated from uncontaminated fresh water sediments. The other sulfate reducing ethylbenzene degrading bacterium was isolated from uncontaminated marine sediment, emphasizing the importance of ethylbenzene degrading bacteria in communities not previously exposed to monoaromatic or polyaromatic hydrocarbons (as in the current study). In summary, combining the culture independent technique SIP with high throughput amplicon sequencing can be a useful approach for investigating the biodegradation of xenobiotics by minority organisms or by a community of organisms. The current work illustrates the utility of these methods for examining carbon pathways in soil communities. Although it was not possible to identify enriched phylotypes at the genus level, important families involved in carbon assimilation were identified. Interestingly, all classified within the Proteobacteria. 51 APPENDIX 52 Table 3.1. Buoyant density of heavy fractions from the labeled and unlabeled samples submitted for amplicon sequencing. Microcosm name SIP Label EB 1 SIP Label EB 2 SIP Unlabel EB 1 SIP Unlabel EB 2 1.780 1.778 1.778 1.780 Buoyant Density of Fractions (g mL-1) 1.755 1.745 1.768 1.756 1.759 1.747 1.765 1.755 1.742 1.748 1.740 1.744 Table 3.2. Summary MiSEQ Illumina data generated from the fractions as well as total DNA extracted from labeled and unlabeled ethylbenzene amended microcosms. SIP Label EB 1 and SIP unlabeled EB 1 with 4 fractions from each and total DNA samples from both SIP Label EB 2 and SIP unlabeled EB 2 with 4 fractions from each and total DNA samples from both Unlabel EB 1 and Unlabel EB 2 total DNA # of Sequences Following Make Contigs command 1758568 Final # of Unique Sequences Final # of Sequences % Chimeric 68926 869977 3.09 OTUs Per Fraction or Sample (Average ± Std dev) 2622 ± 719 2181584 81557 1075361 3.20 2758 ± 743 829470 26696 409141 3.13 1495±254 53 Table 3.3. Classification of 16S rRNA gene clones in the clone library. Phyla Class Acidobacteria Acidobacteria Acidobacteria Acidobacteria Armatimonadetes Bacteriodetes Bacteriodetes Bacteriodetes Firmicutes Firmicutes Firmicutes Firmicutes Gemmatimonadetes Proteobacteria Proteobacteria Proteobacteria Proteobacteria Proteobacteria Proteobacteria Proteobacteria Proteobacteria Proteobacteria Proteobacteria Proteobacteria Proteobacteria Proteobacteria Proteobacteria Proteobacteria Proteobacteria Acidobacteria GP1 Acidobacteria GP3 Acidobacteria GP4 Proteobacteria Proteobacteria TM7 Verrucomicrobia Order Family Flavobacteria Sphingobacteria Flavobacteriales Sphingobacteriales Flavobacteriaceae Chitinophagaceae Bacilli Bacilli Bacilli Bacilli Gemmatimonadetes Alphaproteobacteria Alphaproteobacteria Alphaproteobacteria Alphaproteobacteria Alphaproteobacteria Alphaproteobacteria Betaproteobacteria Betaproteobacteria Betaproteobacteria Betaproteobacteria Betaproteobacteria Betaproteobacteria Betaproteobacteria Betaproteobacteria Gammaproteobacteria Gammaproteobacteria Bacilllales Bacilllales Bacilllales Bacilllales Gemmatimonadales Rhizobiales Rhizobiales Rhizobiales Rhizobiales Rhizobiales Rhodospirillales Burkholderiales Burkholderiales Burkholderiales Burkholderiales Burkholderiales Burkholderiales Rhododcyclales Alicyclobacillaceae Bacillaceae Paenibacillaceae 1 Xanthomonadales Xanthomonadales Xanthomonadaceae Xanthomonadaceae Gammaproteobacteria Xanthomonadales Xanthomonadaceae Spartobacteria Gemmatimonadaceae Brucellaceae Methylobacteriaceae Phyllobacteriaceae Xanthobateraceae Rhodospirillaceae Alcaligenaceae Comamonadaceae Oxalobacteraceae Oxalobacteraceae Oxalobacteraceae Rhodocyclaceae Unclassified TM7 Unclassified Spartobacteria 54 Full name No. of Clones Gp1 Gp3 Gp4 Gp6 Armatimonadetes_Gp4 Flavobacterium Ferruginibacter Unclassified Bacteriodetes Tumebacilus Bacillus Unclassified Bacillales Unclassified Bacillales Gemmatimonas Unclassified Brucellaceae Microvirga Aminobacter Unclassified Rhizobiales Pseudolabrys Unclassified Rhodospirillaceae Pusilimonas Ramlibacter Herbaspirillum Massilia Unclassified Oxalobacteraceae Unclassified Burkholderiales Unclassified Rhodocyclaceae Unclassified Betaproteobacteria Rhodanobacter Unclassified Xanthomonadaceae Luteimonas Unclassified Proteobacteria TM7 genera incertae sedis Spartobacteria genera incertae 3 1 1 1 1 3 6 1 1 5 1 6 1 1 1 1 2 1 4 1 1 4 1 30 10 4 4 12 1 1 8 1 1 Ethylbenzene concencentration mg/L 25 20 Abiotic controls 15 Unlabel EB 2 10 SIP Unlabel EB 2 Unlabel EB 1 5 SIP Label EB 2 SIP Unlabel EB 1 SIP Label EB 1 0 0 5 10 15 20 25 30 Time (days after ethylbenzene addition) Figure 3.1. Concentration of ethylbenzene in microcosms over time ( × - abiotic controls, open triangles - amended with 12C6 ethylbenzene, closed circles - amended with 13C6 ethylbenzene). 55 30 SIP Unlabel EB 1 SIP Label EB 1 25 20 -1 DNA Concentration (ng µL ) 15 10 5 0 1.70 30 1.72 1.74 1.76 1.78 1.80 SIP Unlabel EB 2 SIP Label EB 2 25 20 15 10 5 0 1.70 1.72 1.74 1.76 1.78 1.80 -1 Buoyant Density (g mL ) Figure 3.2. DNA concentration over the ultracentrifugation gradient in fractions from replicate microcosms amended with labeled or unlabeled ethylbenzene. The dashed arrows indicate the fractions that were sequenced from the unlabeled EB amended samples and solid arrows indicate the fractions that were sequenced from the labeled EB amended samples. 56 4000 3500 Number of OTUs 3000 Unlabel EB 1 2500 Unlabel EB 2 SIP Label EB 1 2000 SIP Label EB 2 1500 SIP Unlabel EB 1 SIP Unlabel EB 2 1000 500 0 0 20000 40000 60000 80000 100000 120000 140000 Number of sequences Figure 3.3. Rarefaction curves generated from sequencing data from total DNA extracts for all six microcosms. 57 100 Lentisphaerae Chlorobi WS3 OD1 80 BRC1 Relative Abundance (%) Nitrospira Chlamydiae 60 Armatimonadetes Deinococcus-Thermus Chloroflexi Planctomycetes 40 TM7 Verrucomicrobia Gemmatimonadetes Bacteroidetes 20 Acidobacteria Actinobacteria unclassified 0 Firmicutes Unlabel EB 1 Unlabel EB 2 SIP Label EB 1 SIP Label EB 2 SIP Unlabel EB SIP Unlabel EB 1 2 Proteobacteria Figure 3.4. Comparison of relative abundance of sequences in total genomic DNA extracted from six microcosms amended with 13C6 labeled or unlabeled ethylbenzene at phylum level classification. 58 Unlabel EB 1 SIP Label EB 1 Unlabel EB 2 Rhodanobacter Azoarcus unclassified Rhodospirillaceae unclassified Oxalobacteraceae Thiobacillus unclassified Bacteria unclassified Xanthomonadaceae unclassified Bacillales unclassified Rhodospirillales unclassified Bacillaceae 1 unclassified Betaproteobacteria unclassified Chitinophagaceae unclassified Comamonadaceae unclassified Rhizobiales Gemmatimonas SIP Unlabel EB 1 SIP Unlabel EB 2 SIP Label EB 2 Figure 3.5. Relative abundance (%) of the most common phylotypes (>1%, at the lowest classification level) in all six microcosms (from total DNA extracts). The phylotypes are at the genus level, unless unclassified. 59 Abundance in SIP Label EB 1, SIP Unlabel 25 20 20 15 15 10 10 5 5 0 1.74 1.75 1.76 1.77 1.78 0 1.74 1.79 15 15 10 10 5 5 0 1.74 1.75 1.76 1.77 1.78 0 1.74 1.79 25 20 20 15 15 10 10 5 5 1.75 1.76 1.77 1.78 0 1.74 1.79 6 1.78 1.79 1.75 1.76 1.77 1.78 1.79 Unclassified Rhodospirillaceae 1.75 1.76 1.77 1.78 1.79 Unclassified Xanthomonadaceae 4 2 1.75 1.76 1.77 1.78 1.79 0 1.74 6 6 4 4 2 2 1.75 1.76 1.77 1.78 1.79 0 1.74 10 10 8 8 6 6 4 4 2 2 1.75 1.76 1.77 1.78 1.79 0 1.74 4 4 3 3 2 2 1 1 0 1.74 1.77 12C EB 2 0 1.74 1.76 6 13C EB 4 0 1.74 1.75 8 8 0 1.74 Rhodanobacter Unclassified Oxalobacteraceae 25 0 1.74 Relative Abundance (%) Abundance in SIP Label EB 2, SIP Unlabel EB 2 25 1.75 1.76 1.77 1.78 1.79 0 1.74 1.75 1.76 1.77 1.78 1.79 Unclassified Rhodocyclaceae 1.75 1.76 1.77 1.78 1.79 Unclassified Bacillales 1.75 1.76 1.77 1.78 1.79 Gemmatimonas 1.75 1.76 1.77 1.78 1.79 Buoyant density (g mL-1) Figure 3.6. Relative abundance of the most common phylotypes in heavy fractions from 13C ethylbenzene amended microcosms compared to those from 12C ethylbenzene amended microcosms (SIP Label EB1 and SIP Unlabel EB1 on left, SIP Label EB2 and SIP Unlabel EB2 on right). 60 REFERENCES 61 REFERENCES 1. Weelink SAB, van Eekert MHA, Stams AJM. 2010. Degradation of BTEX by anaerobic bacteria: physiology and application. Rev Environ Sci Bio 9:359-385. 2. Dolfing J, Zeyer J, Bindereicher P, Schwarzenbach RP. 1990. Isolation and characterization of a bacterium that mineralizes toluene in the absence of molecularoxygen. Archives of Microbiology 154:336-341. 3. Schocher RJ, Seyfried B, Vazquez F, Zeyer J. 1991. Anaerobic degradation of toluene by pure cultures of denitrfying bacteria. Archives of Microbiology 157:7-12. 4. Lovley DR, Lonergan DJ. 1990. Anaerobic oxidation of toluene, phenol, and paracresol by the dissimilatory iron-reducing organism, GS-15. Applied and Environmental Microbiology 56:1858-1864. 5. Beller HR, Spormann AM, Sharma PK, Cole JR, Reinhard M. 1996. Isolation and characterization of a novel toluene-degrading, sulfate-reducing bacterium. Applied and Environmental Microbiology 62:1188-1196. 6. Morasch B, Schink B, Tebbe CC, Meckenstock RU. 2004. Degradation of o-xylene and m-xylene by a novel sulfate-reducer belonging to the genus Desulfotomaculum. Archives of Microbiology 181:407-417. 7. Kunapuli U, Jahn MK, Lueders T, Geyer R, Heipieper HJ, Meckenstock RU. 2010. Desulfitobacterium aromaticivorans sp nov and Geobacter toluenoxydans sp nov., ironreducing bacteria capable of anaerobic degradation of monoaromatic hydrocarbons. International Journal of Systematic and Evolutionary Microbiology 60:686-695. 8. Chakraborty R, O'Connor SM, Chan E, Coates JD. 2005. Anaerobic degradation of benzene, toluene, ethylbenzene, and xylene compounds by Dechloromonas strain RCB. Applied and Environmental Microbiology 71:8649-8655. 9. Kasai Y, Takahata Y, Manefield M, Watanabe K. 2006. RNA-based stable isotope probing and isolation of anaerobic benzene-degrading bacteria from gasolinecontaminated groundwater. Applied and Environmental Microbiology 72:3586-3592. 10. Rabus R, Widdel F. 1995. Anaerobic degradation of ethylbenzene and other aromatic hydrocarbons by new denitrifying bacteria. Arch Microbiol 163:96-103. 11. Ball HA, Johnson HA, Reinhard M, Spormann AM. 1996. Initial reactions in anaerobic ethylbenzene oxidation by a denitrifying bacterium, strain EB1. Journal of Bacteriology 178:5755-5761. 62 12. Wöhlbrand L, Kallerhoff B, Lange D, Hufnagel P, Thiermann J, Reinhardt R, Rabus R. 2007. Functional proteomic view of metabolic regulation in “Aromatoleum aromaticum” strain EbN1. Proteomics 7:2222-2239. 13. Coates JD, Chakraborty R, Lack JG, O'Connor SM, Cole KA, Bender KS, Achenbach LA. 2001. Anaerobic benzene oxidation coupled to nitrate reduction in pure culture by two strains of Dechloromonas. Nature 411:1039-1043. 14. Weelink SAB, van Doesburg W, Saia FT, Rijpstra WIC, Roling WFM, Smidt H, Stams AJM. 2009. A strictly anaerobic betaproteobacterium Georgfuchsia toluolica gen. nov., sp nov degrades aromatic compounds with Fe(III), Mn(IV) or nitrate as an electron acceptor. Fems Microbiol Ecol 70:575-585. 15. Kniemeyer O, Fischer T, Wilkes H, Glockner FO, Widdel F. 2003. Anaerobic degradation of ethylbenzene by a new type of marine sulfate-reducing bacterium. Applied and Environmental Microbiology 69:760-768. 16. Radajewski S, Ineson P, Parekh NR, Murrell JC. 2000. Stable-isotope probing as a tool in microbial ecology. Nature 403:646-649. 17. Whitby C, Bailey M, Whiteley A, Murrell C, Kilham K, Prosser J, Lappin-Scott H. 2005. Stable isotope probing links taxonomy with function in microbial communities Microbial ecologists have a culture-independent means for analyzing sources of metabolic activities among complex mixtures of microbes. Asm News 71:169-173. 18. DeRito CM, Pumphrey GM, Madsen EL. 2005. Use of field-based stable isotope probing to identify adapted populations and track carbon flow through a phenoldegrading soil microbial community. Applied and Environmental Microbiology 71:78587865. 19. Singleton DR, Powell SN, Sangaiah R, Gold A, Ball LM, Aitken MD. 2005. Stableisotope probing of bacteria capable of degrading salicylate, naphthalene, or phenanthrene in a Bioreactor treating contaminated soil. Applied and Environmental Microbiology 71:1202-1209. 20. Luo C, Xie S, Sun WM, Li X, Cupples AM. 2009. Identification of a novel toluenedegrading bacterium from the candidate phylum TM7, as determined by DNA stable isotope probing. Applied and Environmental Microbiology 75:4644-4647. 21. Sun W, Cupples AM. 2012. Diversity of five anaerobic toluene-degrading microbial communities investigated using stable isotope probing. Applied and environmental microbiology 78:972-980. 22. Luo CL, Xie SG, Sun WM, Li XD, Cupples AM. 2009. Identification of a novel toluene-degrading bacterium from the candidate phylum TM7, as determined by DNA stable isotope probing. Applied and Environmental Microbiology 75:4644-4647. 63 23. Sun W, Cupples AM. 2012. Diversity of five anaerobic toluene degrading microbial communities investigated using stable isotope probing (SIP). Applied and Environmental Microbiology 78:972-980. 24. Sun W, Sun X, Cupples AM. In press. Presence, diversity and enumeration of functional genes (bssA and bamA) relating to toluene degradation across a range of redox conditions and inoculum sources. Biodegradation. 25. Sun WM, Xie SG, Luo CL, Cupples AM. 2010. Direct Link between Toluene Degradation in Contaminated-Site Microcosms and a Polaromonas Strain. Applied and Environmental Microbiology 76:956-959. 26. Xie SG, Sun WM, Luo CL, Cupples AM. 2010. Stable Isotope Probing Identifies Novel m-Xylene Degraders in Soil Microcosms from Contaminated and Uncontaminated Sites. Water Air and Soil Pollution 212:113-122. 27. Xie SG, Sun WM, Luo CL, Cupples AM. 2011. Novel aerobic benzene degrading microorganisms identified in three soils by stable isotope probing. Biodegradation 22:7181. 28. Andeer P, Stahl DA, Lillis L, Strand SE. 2013. Identification of microbial populations assimilating nitrogen from RDX in munitions contaminated military training range soils by high sensitivity stable isotope probing. Environmental Science & Technology 47:10356-10363. 29. Pilloni G, von Netzer F, Engel M, Lueders T. 2011. Electron acceptor-dependent identification of key anaerobic toluene degraders at a tar-oil-contaminated aquifer by Pyro-SIP. Fems Microbiology Ecology 78:165-175. 30. Gutierrez T, Singleton DR, Berry D, Yang TT, Aitken MD, Teske A. 2013. Hydrocarbon-degrading bacteria enriched by the Deepwater Horizon oil spill identified by cultivation and DNA-SIP. Isme J 7:2091-2104. 31. Wu YC, Ke XB, Hernandez M, Wang BZ, Dumont MG, Jia ZJ, Conrad R. 2013. Autotrophic growth of bacterial and archaeal ammonia oxidizers in freshwater sediment microcosms incubated at different temperatures. Appl Environ Microb 79:3076-3084. 32. Bell TH, Yergeau E, Martineau C, Juck D, Whyte LG, Greer CW. 2011. Identification of nitrogen-incorporating Bacteria in petroleum-contaminated Arctic soils by using [15N]DNA-based stable isotope probing and pyrosequencing. Appl Environ Microb 77:4163-4171. 33. Nelson CE, Carlson CA. 2012. Tracking differential incorporation of dissolved organic carbon types among diverse lineages of Sargasso Sea bacterioplankton. Environ Microbiol 14:1500-1516. 34. Lu L, Jia ZJ. 2013. Urease gene-containing Archaea dominate autotrophic ammonia oxidation in two acid soils. Environ Microbiol 15:1795-1809. 64 35. Xia WW, Zhang CX, Zeng XW, Feng YZ, Weng JH, Lin XG, Zhu JG, Xiong ZQ, Xu J, Cai ZC, Jia ZJ. 2011. Autotrophic growth of nitrifying community in an agricultural soil. Isme Journal 5:1226-1236. 36. Wang BZ, Zheng Y, Huang R, Zhou X, Wang DM, He YQ, Jia ZJ. 2014. Active ammonia oxidizers in an acidic soil are phylogenetically closely related to neutrophilic Archaeon. Appl Environ Microb 80:1684-1691. 37. Zemb O, Lee M, Gutierrez-Zamora ML, Hamelin J, Coupland K, Hazrin-Chong NH, Taleb I, Manefield M. 2012. Improvement of RNA-SIP by pyrosequencing to identify putative 4-n-nonylphenol degraders in activated sludge. Water Research 46:601610. 38. Shao YQ, Arias-Cordero E, Guo HJ, Bartram S, Boland W. 2014. In vivo pyro-SIP assessing active gut microbiota of the cotton leafworm, Spodoptera littoralis. Plos One 9. 39. Kim SJ, Park SJ, Jung MY, Kim JG, Madsen EL, Rhee SK. 2014. An uncultivated nitrate-reducing member of the genus Herminiimonas degrades toluene. Appl Environ Microb 80:3233-3243. 40. Angel R, Conrad R. 2013. Elucidating the microbial resuscitation cascade in biological soil crusts following a simulated rain event. Environ Microbiol 15:2799-2815. 41. Yang Y, McCarty PL. 1998. Competition for hydrogen within a chlorinated solvent dehalogenating anaerobic mixed culture. Environmental Science & Technology 32:35913597. 42. Caporaso JG, Lauber CL, Walters WA, Berg-Lyons D, Lozupone CA, Turnbaugh PJ, Fierer N, Knight R. 2011. Global patterns of 16S rRNA diversity at a depth of millions of sequences per sample. Proceedings of the National Academy of Sciences of the United States of America 108:4516-4522. 43. Caporaso JG, Lauber CL, Walters WA, Berg-Lyons D, Huntley J, Fierer N, Owens SM, Betley J, Fraser L, Bauer M, Gormley N, Gilbert JA, Smith G, Knight R. 2012. Ultra-high-throughput microbial community analysis on the Illumina HiSeq and MiSeq platforms. Isme Journal 6:1621-1624. 44. Schloss PD. 2009. A High-Throughput DNA Sequence Aligner for Microbial Ecology Studies. Plos One 4. 45. Schloss PD 6 August 2013, at 7.30 2013, posting date. MiSeq SOP. [Online.] 46. Kozich JJ, Westcott SL, Baxter NT, Highlander SK, Schloss PD. 2013. Development of a dual-index sequencing strategy and curation pipeline for analyzing amplicon sequence data on the MiSeq Illumina sequencing platform. Applied and Environmental Microbiology 79:5112-5120. 65 47. Pruesse E, Quast C, Knittel K, Fuchs BM, Ludwig W, Peplies J, Gloeckner FO. 2007. SILVA: a comprehensive online resource for quality checked and aligned ribosomal RNA sequence data compatible with ARB. Nucleic Acids Research 35:71887196. 48. Wang Q, Garrity GM, Tiedje JM, Cole JR. 2007. Naive Bayesian classifier for rapid assignment of rRNA sequences into the new bacterial taxonomy. Applied and Environmental Microbiology 73:5261-5267. 49. Edgar RC, Haas BJ, Clemente JC, Quince C, Knight R. 2011. UCHIME improves sensitivity and speed of chimera detection. Bioinformatics 27:2194-2200. 50. Kunapuli U, Lueders T, Meckenstock RU. 2007. The use of stable isotope probing to identify key iron-reducing microorganisms involved in anaerobic benzene degradation. Isme Journal 1:643-653. 51. Ronen Z, Yanovich Y, Goldin R, Adar E. 2008. Metabolism of the explosive hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) in a contaminated vadose zone. Chemosphere 73:1492-1498. 52. Klepárnik K, Foret F, Berka J, Goetzinger W, Miller AW, Karger BL. 1996. The use of elevated column temperature to extend DNA sequencing read lengths in capillary electrophoresis with replaceable polymer matrices. Electrophoresis 17:1860-1866. 53. Shendure J, Ji HL. 2008. Next-generation DNA sequencing. Nature Biotechnology 26:1135-1145. 54. Zhang H, Sekiguchi Y, Hanada S, Hugenholtz P, Kim H, Kamagata Y, Nakamura K. 2003. Gemmatimonas aurantiaca gen. nov., sp nov., a gram-negative, aerobic, polyphosphate-accumulating micro-organism, the first cultured representative of the new bacterial phylum Gemmatimonadetes phyl. nov. Int J Syst Evol Micr 53:1155-1163. 55. Corgié S, Beguiristain T, Leyval C. 2004. Spatial distribution of bacterial communities and phenanthrene degradation in the rhizosphere of Lolium perenne L. Applied and Environmental Microbiology 70:3552-3557. 56. Regonne RK, Martin F, Mbawala A, Ngassoum MB, Jouanneau Y. 2013. Identification of soil bacteria able to degrade phenanthrene bound to a hydrophobic sorbent in situ. Environmental Pollution 180:145-151. 57. Martin F, Torelli S, Le Paslier D, Barbance A, Martin-Laurent F, Bru D, Geremia R, Blake G, Jouanneau Y. 2012. Betaproteobacteria dominance and diversity shifts in the bacterial community of a PAH-contaminated soil exposed to phenanthrene. Environmental Pollution 162:345-353. 58. Singleton DR, Jones MD, Richardson SD, Aitken MD. 2013. Pyrosequence analyses of bacterial communities during simulated in situ bioremediation of polycyclic aromatic hydrocarbon-contaminated soil. Applied Microbiology and Biotechnology 97:8381-8391. 66 59. Sun W, Sierra-Alvarez R, Milner L, Field JA. 2010. Anaerobic oxidation of arsenite linked to chlorate reduction. Applied and Environmental Microbiology 76:6804-6811. 60. Martins M, Faleiro ML, Chaves S, Tenreiro R, Costa MC. 2010. Effect of uranium (VI) on two sulphate-reducing bacteria cultures from a uranium mine site. Science of the Total Environment 408:2621-2628. 61. Martins M, Faleiro ML, Chaves S, Tenreiro R, Santos E, Costa MC. 2010. Anaerobic bio-removal of uranium (VI) and chromium (VI): comparison of microbial community structure. Journal of Hazardous Materials 176:1065-1072. 62. Mbadinga SM, Wang L-Y, Zhou L, Liu J-F, Gu J-D, Mu B-Z. 2011. Microbial communities involved in anaerobic degradation of alkanes. International Biodeterioration & Biodegradation 65:1-13. 63. van der Zaan BM, Saia FT, Stams AJ, Plugge CM, de Vos WM, Smidt H, Langenhoff AA, Gerritse J. 2012. Anaerobic benzene degradation under denitrifying conditions: Peptococcaceae as dominant benzene degraders and evidence for a syntrophic process. Environmental Microbiology 14:1171-1181. 64. Rotaru AE, Probian C, Wilkes H, Harder J. 2010. Highly enriched Betaproteobacteria growing anaerobically with p‐xylene and nitrate. Fems Microbiology Ecology 71:460468. 67 CHAPTER 4 COMPARATIVE STUDY OF RDX DEGRADING MICROBIAL COMMUNITIES IN FOUR SOILS ABSTRACT The explosive hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) has contaminated many military sites. Recently, attempts to remediate these sites have focused on biostimulation to promote in situ RDX biodegradation. Although many RDX degrading isolates have been obtained in the laboratory, little is known about the potential of microorganisms to degrade this chemical in mixed microbial communities. The current study compared four soils to elucidate the microorganisms linked to RDX degradation. These soils were selected as they had no previous exposure to RDX, therefore their microbial communities offered an excellent baseline to determine changes following exposure to RDX. The work involved the collection of two sets of molecular data. First, high throughput sequencing was used to determine which phylotypes experienced an increase in relative abundance following RDX degradation. For this, DNA was sequenced from 1) the initial soil, 2) microcosms following RDX degradation and 3) control microcosms without RDX. Second, to ascertain which phylotypes were responsible for label (ring 15N or 13C) uptake from RDX, stable isotope probing (SIP) was performed on two soils. The sequencing data provided valuable information on which phylotypes increased in abundance in RDX degrading microcosms compared to control microcosms. The most notable trend was the increase in abundance of Brevundimonas and/or unclassified Bacillaceae 1 in the four soils 68 studied. These data indicate these two phylotypes are benefiting from RDX degradation under these conditions. Although members of the family Bacillaceae 1 have previously been linked to RDX degradation, the genus Brevundimonas has not been previously associated with RDX degradation. SIP data indicated that phylotypes classifying within the Firmicutes and Actinobacteria (Rhodococcus) were involved in label uptake from RDX. Overall, these data suggest both novel RDX degraders and previously reported RDX degraders were associated with RDX removal in these soils. INTRODUCTION The manufacturing, transport and use of the nitramine explosive hexahydro-1,3,5-trinitro-1,3,5triazine (RDX) has resulted in groundwater and sediment contamination at many military sites. RDX has moderate solubility, low sorption and low vapor pressure resulting in significant mobility in groundwater and sediments. Due to neurotoxic effects, the potential carcinogenic nature of RDX and widespread RDX contamination, the US EPA has issued a life-time drinking water health advisory level of 2 µg/L RDX. Although RDX was initially thought to be recalcitrant, the chemical has been shown amenable to biodegradation under both aerobic and anaerobic conditions. Thus, bioremediation has been considered a viable option to treat RDX contaminated sites. As many of these sites are oxygen depleted, it is more suitable to consider RDX biodegradation under anaerobic conditions. Several bacteria have been isolated with the ability to degrade RDX under anaerobic conditions. These microorganisms classify within the phyla Firmicutes, Fusobacteria and Proteobacteria. The first anaerobic RDX degrading isolate obtained, Clostridia bifermentans, classified with the 69 Firmicutes (1). Other Clostridia isolates have also been reported with RDX degrading abilities, including, for example, Clostridia sp. EDB2 and Clostridia acetobutylicum (2-5). Other RDX degrading bacteria belonging to the Firmicutes include two Acetobacterium spp. (6, 7), two Bacillus spp. (8) and a Desulfitobacterium strain (9). Fusobacteria sp. HAQ-EB21 (10, 11) is the only anaerobic RDX degrading isolate from the phylum Fusobacteria. The Proteobacteria, particularly the classes Gamma and Deltaproteobacteria, contain diverse anaerobic RDX degrading isolates. In the Gammaproteobacteria, RDX degrading bacteria have been isolated from the families Enterobacteriaceae (12-16), Pseudomonadaceae (8, 17), Shewanellaceae (10, 18, 19) and Xanthomonadaceae (20). Isolates belonging to Cystobacteraceae (9), Geobacteraceae (21) or Desulfovibrionaceae (11, 22, 23) within the Deltaproteobacteria are also able to degrade RDX. Although much is known about pure cultures able to degrade RDX in the laboratory, very little is known about the microorganisms responsible for RDX degradation in mixed communities or at contaminated sites. Researchers have reported the microorganisms detected in field samples actively degrading RDX were generally not closely related to previously reported RDX degrading isolates (24). It is now widely recognized that only a small fraction of microorganisms can be isolated and cultivated in the laboratory (25), therefore, it is likely that many RDX degraders have yet to be identified. This information is important because if several key species were consistently linked to effective RDX degradation in mixed communities, their presence could provide strong evidence of RDX natural attenuation. 70 Recently, a small number of studies have used stable isotope probing (SIP) to determine which microorganisms were involved in carbon and/or nitrogen uptake from RDX (26-29). One report revealed the important of Sphingobacteriales (Bacteroidetes) in RDX degradation (28). A SIP (15N labeled RDX) study by another group using samples from the Picatinny Arsenal Site, identified five phylotypes similar to known RDX degraders and ten phylotypes not previously linked to RDX degradation (29). A 13C labeled RDX SIP study again using samples from one site, also revealed that novel phylotypes and phylotypes similar to known RDX degraders were responsible for label uptake (27). A recent SIP investigation (15N labeled RDX) using material from the Eglin Air Force Base bombing range, reported that RDX degradation was spatially heterogeneous and dependent on carbon source addition (26). These researchers reported that Rhodococcus (a known RDX degrader) as well as several Proteobacteria were highly labeled with 15N. These studies have led to the conclusion that RDX degrading microorganisms are phylogenetically more diverse than was previously thought. In the past few years, SIP has been combined with high throughput sequencing. For example, researchers have examined the biodegradation of toluene (30), the thermophilic anaerobic conversion of microcrystalline cellulose (31), nitrogen incorporation from monoammonium phosphate in arctic soils (32) and carbon uptake in microbial fuel cells (33) using this technique. In the current study, SIP was combined with high throughput sequencing to compare the RDX degrading communities in four soils without prior exposure to RDX. This is the first investigation of RDX degrading communities in uncontaminated soils using the combined approaches. Further, is it is first study to use high throughput sequencing to compare the microbial communities between RDX degrading soils. 71 METHODS Chemicals Unlabeled and ring-labeled RDX (13C3, 99%; 15N3, 98%) dissolved in acetonitrile were purchased from Cambridge Isotope Laboratories, Inc. (Tewksbury, MA). HPLC grade acetonitrile (≥99.8 % purity) was purchased from EMD Chemicals Inc. (New Jersey, USA). Other reagents were purchased from Sigma-Aldrich (St. Louis, MO, USA), Fisher BioReagent (New Jersey, USA), or Invitrogen (Carlsbad, CA, USA) unless otherwise stated. Experimental Design Four agricultural soils were utilized in these experiments and, to our knowledge, these soils have not previously been exposed to RDX (Table 1). The soil microcosms for screening RDX degradation were established as previously described, with only slight modifications (28, 34). Briefly, for each soil, 60 mL serum bottles contained 1 g soil (wet weight), 20 mg L-1 unlabeled RDX dissolved in acetonitrile (as the sole nitrogen source), 201.6 mg L-1 glucose, and 4 mL of a minimal salts media (MSM) (28, 34). For the high throughput sequencing and SIP studies, the microcosms were prepared with 1 g soil (wet weight), 1 mL of inocula from the corresponding RDX screening microcosm, 20 mg L-1 unlabeled or ring-labeled RDX dissolved in acetonitrile (as the sole nitrogen source), 201.6 mg L-1 glucose, and 4 mL of a minimal salts media (MSM). Triplicate live sample microcosms were amended with ring-labeled RDX (13C3, 15N3), triplicate live and killed abiotic controls (autoclaved) microcosms were amended with unlabeled RDX (12C3, 14N3). Two successive amendments of 20 mg L-1 of RDX were added to the live labeled and unlabeled RDX 72 microcosms. The triplicate abiotic killed controls were prepared in a similar manner to the live microcosms except that they received only one amendment of unlabeled RDX. All microcosms were sealed with rubber stoppers and aluminum seals and were covered with aluminum foil to prevent photodegradation. The microcosms were incubated between 3 to 7 weeks in the dark at room temperature (~20 °C) without shaking while being monitored for RDX degradation. Due to time constraints, only DNA extracted from soils 9 and 10 were ultracentrifuged for fraction analysis. However, total DNA extracts from all four soils, extracted from both replicates of the labeled RDX (hereafter called Labeled 1, Time t and Labeled 2, Time t) and unlabeled RDX microcosms (hereafter called Unlabeled 1, Time t and Unlabeled 2, Time t) were sequenced. Additional live control microcosms, without RDX addition, were also prepared with all four soils and these were maintained under the same conditions as the live sample microcosms. Nucleic acids were extracted from these control microcosms (hereafter called No RDX, Time t) at the same time as those extracted from the RDX amended microcosms. Further, DNA was extracted from each soil prior to any manipulation (hereafter called No RDX, Time 0). These two sets of DNA extracts were sequenced to determine how the communities changed followed RDX degradation compared to the initial soil and the no RDX controls. RDX Extraction and Analysis RDX concentrations were determined using high performance liquid chromatography (HPLC) as previously described (34) with modifications. For this, 0.5 mL (extracted using a 1 mL BD syringe with a 21 gauge needle) was placed in a 1.7 mL sterile microcentrifuge tube. An equal 73 volume of acetonitrile was added and the microcentrifuge tubes were shaken for 2 hours at room temperature to extract RDX. The samples were then centrifuged for 5 minutes at 10,000 x g and the supernatant was filtered using acetonitrile wetted filters (PVDF, 0.22 µm, Whatman) into HPLC amber vials (Sigma-Aldrich, St. Louis, MO, USA). External standards for the calibration curve were prepared with a dilution factor of 2 to account for the sample dilution at the liquidliquid extraction step. HPLC analysis involved a Perkin Elmer (PE) series 200 autosampler; PE binary LC Pump 250; PE diode array detector 235C, at wavelength 255nm; Supelco C18 (25 cm X 4.6 mm, 5 µm) column; and isocratic conditions (40% acetonitrile and 60% 0.1% H3PO4 acidified deionized water) at a flow rate of 1 mL min-1. DNA Extraction and Isopycnic Centrifugation As stated above, total genomic DNA from all four soils was extracted to establish the baseline microbial community (No RDX, Time 0). Total genomic DNA was also extracted from all sixteen microcosms (four soils, Labeled 1 & 2, Unlabeled 1 & 2, Time t) after the complete degradation of the second amendment of RDX. In addition, DNA was extracted from the corresponding live control microcosms at the same time (No RDX, Time t). All DNA extractions used the Power Soil DNA extraction kit (MO BIO Laboratories, Inc. Carlsbad, CA). Isopycnic centrifugation was conducted for soils 9 and 10, following previously described procedures (35-40). DNA extracted from one labeled and one unlabeled microcosm for these soils was ultracentrifuged separately in Tris EDTA - Cesium Chloride (CsCl) gradients to obtain density gradient fractions. Approximately 3 to 5 µg of the total genomic DNA was added to a CsCl and TrisEDTA solution with an initial buoyant density (BD) of ~1.72 g mL-1 in a 5.1 mL 74 Beckmann Coulter Quick-Seal polyallomer tubes. Tubes were then sealed (Cordless quick-seal tube topper, Beckman) and centrifuged at 178,000 x g for 46 hours at 20 oC in a Wx Sorvall Ultra 80 ultracentrifuge fitted with a Stepsaver 70 V6 Vertical Titanium Rotor (Thermo Scientific). Each sample was then divided into 20 fractions (each 250 µL) by displacement with molecular grade water using a syringe pump system. The BD of all samples was measured using a model AR200 digital refractometer (Leica Microsystems Inc.) and the DNA from each fraction was precipitated using ethanol and glycogen. Precipitated DNA was then re-suspended in 30 µL PCR grade water and stored at -20 °C for further analysis. Amplicon Sequencing and Data Analysis Total genomic DNA extracted from four microcosms (Labeled 1, Labeled 2, Unlabeled 1, Unlabeled 2, Time t) for all four soils (soil 9, 10 , 11 and 12) and from the microcosms without RDX (No RDX, Time t) for all four soils were submitted for amplicon sequencing at Michigan State University’s Research Technology and Support Facility (RTSF). Total genomic DNA extracted from all four of the initial soils (no RDX, Time 0) were also amplicon sequenced. Additionally, the first ultracentrifugation eight fractions with high BD values from soils 9 and 10 were also submitted for amplicon sequencing, in triplicate. Amplification of the V4 region and paired-end high throughput amplicon sequencing on the Illumina MiSeq platform at RTSF was conducted using a procedure previously described (41, 42). Sequencing data obtained from the MiSeq platform Laboratory Information Management System (LIMS) was analyzed using Mothur v.1.33.2 (43) using the MiSeq standard operating procedure (44). The sequences data in the fastq format were processed using Mothur to remove the 75 barcodes and these were then aligned to form contiguous sequences. The data was checked for sequencing errors and read length. The sequences were then aligned and checked for chimeras using UCHIME in Mothur (45). Following which, the sequences were classified into OTU’s and phylotypes using the Ribosomal Database Project dataset within Mothur. The sequence data was then imported into Microsoft Excel for further analysis. High throughput sequences were also processed using the default pipeline option on MG-RAST (46) to create the MLTreeMap and the Heatmap with the following parameters, annotation source: RDP; maximum e-value cutoff :1e-5; minimum % identity cutoff: 60%; minimum alignment length cutoff: 15. Illumina sequencing data were deposited to the NCBI Sequence Read Archive under Bioproject Number PRJNA263419. RESULTS AND DISCUSSION RDX Degradation All four agricultural soils illustrated RDX degradation (Figure 4.1). The first amendment of RDX was degraded in 15 to 31 days and the second amendment was degraded in 7 to 12 days. No significant degradation of RDX was observed in the killed control microcosms. Among the four soils, RDX degradation was the fastest in the microcosms with soil 10 and was the slowest in the microcosms with soil 9. Comparison at Phyla Level Total genomic DNA from all treatments and live controls (all four soils) as well as DNA from the ultracentrifugation fractions (soils 9 and 10) were submitted for amplicon sequencing (116 samples in total). An average of >120,000 reads was obtained per sample (Table 2). On average, 76 27% of the sequences did not align properly into contiguous sequences resulting in read lengths >275 bp and these were dropped from the analysis. Another 2.3% (average) of the sequences were identified as chimeric or as non-bacterial (mitochondrial/chloroplast) by the Bayesian classifier within Mothur. Phylotype classification and sequence abundance data were used to assess the microbial communities in each of the four soils (Figure 4.2). The relative abundance of each phyla for all four initial soils (No RDX, Time 0) were remarkably similar. That is, the phylotypes classified into Firmicutes, Proteobacteria, Actinobacteria (the three most dominant), Bacteroidetes, Acidobacteria, Verrucomicrobia, Chloroflexi and Gemmatimonadetes. In contrast, for all four soils, in both the RDX amended microcosms (Labeled, Unlabeled 1 & 2, Time t) and in the microcosms not amended with RDX (No RDX, Time t), the communities were primarily composed of Proteobacteria (particularly soils 10 and 11) and Firmicutes (particularly soils 9 and 12). This increase in Firmicutes is likely a result of oxygen depletion in these samples (the microcosms were sealed for more than 20 days). For the RDX amended samples, these results are somewhat expected, as the majority of anaerobic RDX degrading isolates belong to the Firmicutes and Proteobacteria (Table 1). At the phyla level, for soils 9 and 12, the classifications were similar between the microcosms amended with RDX and those not amended with RDX. For soils 10 and 11, the abundance of Proteobacteria increased when RDX was amended. Although phyla level classifications provide a useful baseline to compare these communities, comparisons at the phylotype level are more informative and will be discussed below. 77 Microbial Community Analysis The microbial communities were also compared using MLTreeMap (Figure 4.3a). Again, the phyla with the greatest representation and the most diversity included the Proteobacteria and Firmicutes. However, no clear trends between the treatments could be ascertained from this analysis. A heatmap was also generated with the sequencing data for all four soils and for the microcosms with and without RDX (Figure 4.3b). The image illustrates distinct trends between the four initial soils and the other treatments. As expected, the microbial communities from the four initial soils clustered together. Clear differences can be seen between the initial soils and the other treatments for the Clostridia, Cloroflexi, Gammatimonadetes, Spirochaetia and unclassified Acidobacteria. The heatmap did not provide any strong trends for the differences in community structure between the RDX amended microcosms and those not amended with RDX. Rarefaction curves were also generated for the initial soils (No RDX, Time 0), for the microcosms amended with RDX (Labeled, Unlabeled 1 & 2, Time t) and for the microcosms without RDX (No RDX, Time t) (Figure 4.4a). The curves are all similar except for three (soils 10, 11 and 12) of the four initial soils, which illustrated a much steeper gradient, indicating incomplete sequence coverage for these samples. The curves for the other samples started to plateau, suggesting more complete sequence coverage. Principal component analysis was also performed on the sequencing data (Figure 4.4b). Clustering occurred for each soil in the four samples amended with RDX (Labeled 1 & 2, Unlabeled 1 & 2). The clusters for soils 10 and 11 were close together, as were the clusters for 78 soils 9 and 12. Separate clusters were also formed for each of the four initial soils (No RDX, Time 0) and for the four soil microcosms without RDX addition (No RDX, Time t). To more precisely compare the relative abundance of phylotypes across treatments, the most abundant phylotypes (>1% relative abundance) for each microbial community were determined. The most abundant phylotypes were similar across all four initial soils (No RDX, Time 0) (Figure 4.5a). The majority of the abundant phylotypes classified within the Firmicutes, Proteobacteria, Actinobacteria and Acidobacteria, with only one in each of the Chloroflexi, Verrucomicrobia, Gemmatimonadetes and Bacteroidetes. Each of the four soils contained a large relative abundance of unclassified Bacteria. In general, the most abundant phylotypes in all four soils, included unclassified Bacillales, unclassified Sphingomonadaceae, Arthrobacter, unclassified Actinomycetales, unclassified Actinobacteria, Acidobacteria Gp1 and Gp 3 and unclassified Chitinophagaceae. The most abundant phylotypes in the microcosms not amended with RDX (No RDX, Time t) were different (Figure 4.5b). These phylotypes only classified within the Firmicutes, Proteobacteria, Actinobacteria and Bacteroidetes. The most abundant phylotypes were primarily Firmicutes, including Clostridium XIVa, Tissierella, Sedimentibacter and unclassified Clostridiaceae 1. In addition, the phylotypes Rhodococcus and unclassified Chitinophagaceae were abundant in three of the four soils. Not surprisingly, these data clearly indicate that the incubation conditions (sealed bottles with glucose and a minimal salts media) completely changed the microbial community structure. Interestingly, the microbial community that developed following incubation was similar between soils. 79 The most abundant phylotypes were also determined for the microcosms amended with RDX (Figure 4.6). Similar to the data discussed above (No RDX, Time t), the abundant phylotypes classified only within the phyla Firmicutes, Proteobacteria, Actinobacteria and Bacteroidetes (only for soils 10 and 11). Again, the majority of abundant phylotypes classified within the Firmicutes. Similar phylotypes were enriched across all four soils. For the Firmicutes this included Sedimentibacter, Tissierella, unclassified Clostridiaceae 1, unclassified Bacillaceae 1. Other common phylotypes included Brevundimonas (soils 10, 11 and 12), Rhodococcus and unclassified Bacteria (all four soils). In general, the relative abundance values were similar between replicates and between the labeled and unlabeled RDX amended samples. Phylotypes Increasing in Abundance Following RDX Degradation The communities were further examined to ascertain which microorganisms increased in abundance following RDX degradation compared to the initial soils and the microcosms without RDX. These data indicate which microorganisms are gaining a benefit from RDX degradation (perhaps as an energy, carbon or nitrogen source). For each soil, this involved the selection of the eight most abundant phylotypes in the microbial communities following RDX degradation. Following this, for each soil, the relative abundance of these eight phylotypes was determined in the initial soils and in the microcosms not amended with RDX. The comparison of these data for each soil has been summarized (Figure 4.7). The data analysis produced two interesting common trends between the four soils. The phylotypes Brevundimonas (Proteobacteria, Alphaproteobacateria, Caulobacterales, Caulobateraceae) and unclassified Bacillaceae 1 (Firmicutes, Bacilli, Bacillales) were both 80 dominant following RDX degradation compared to the controls (initial soil and no RDX microcosms) in at least three of the four soils. For example, Brevundimonas illustrated a clear increase in relative abundance following RDX degradation in all four replicates for soils 10, 11 and 12, compared to the controls. The same trend can be seen for unclassified Bacillaceae 1 for soils 9, 11 and 12. The family Bacillaceae 1 contains 36 genera, any of which could have been involved in RDX degradation in the current study. Bacillus is the only genus within the family previously linked to RDX degradation (8). Members of this family were also dominant among the RDX degrading microbial community from a saturated contaminated surface soil (47) and have previously been reported to be involved in PCB (48) and PAH (49) degradation. The phylotype Brevundimonas classifies within the Alphaproteobacteria. Although many known anaerobic RDX degraders belong to the phylum Proteobacteria, only one RDX isolate (Methylobacterium sp.) belongs to the Alphaproteobacteria (50). Therefore, the current study has provided new insights into the importance of this phylotype for RDX degradation. In other research, Brevundimonas was involved in degradation of an organophosphorus insecticide (51) and cellulose (52). They have also been associated with direct petroleum hydrocarbons degradation (53, 54) and with the root-associated bacterial communities of plants involved in phytoremediation (55). The phylotypes Rhodococcus, Tisserella (soil 9), unclassified Erysipelotrichaceae (soil 10), unclassified Alcaligenaceae (soil 11) and Flavobacterium (soil 12) were also more abundant in 81 the RDX treated samples in comparison to the no RDX controls, however, their relative abundance values were lower than Brevundimonas or unclassified Bacillaceae 1 (Figure 4.7). Phylotypes Indicated In Label Uptake Eight heavy fractions generated from the labeled and unlabeled RDX amended microcosms were sequenced for soils 9 and 10. The most abundant phylotypes in the heavy fractions from the labeled RDX amended samples were the focus of additional analysis, as these have previously been linked to label uptake (35-38). The relative abundance of these phylotypes were compared to their relative abundance values in the unlabeled fractions (Figure 4.8). In soil 9, no clear enrichment of phylotypes was observed in the fractions from the labeled RDX amended microcosm compared to the unlabeled RDX amended microcosm (Figure 4.8a). Although the average relative abundance values were higher for unclassified Bacillaceae 1, unclassified Clostridiales and Tissierella from the labeled RDX amended microcosm, the standard deviations for each point were large. This trend could be attributed to any of three factors. One, the data was not normalized to DNA concentration or the number of 16S rRNA genes in each fraction (based on current studies in our laboratory, it is likely that the DNA concentration was considerably less in the unlabeled RDX amended fractions). Unfortunately, not enough fraction samples remained for this analysis. The second factor could be a dilution of the label between several phylotypes, resulting in no clear trend for any particular phylotype. Third, only the ring carbon and nitrogen was labeled, therefore uptake of any other nitrogen would not have been noted with this experimental design. As discussed above, for soil 9, unclassified Bacilliaceae 1 was enriched following RDX degradation. Unfortunately, the SIP 82 results (Figure 4.8a) are not strong enough to indicate this phylotypes was also responsible for ring label uptake. The standard deviations were smaller for the fractions generated from soil 10 (Figure 4.8b). In this soil, phylotypes belonging to Pusillimonas, Rhodococcus and Sedimentibacter were more abundant in the fractions from the labeled amended samples compared to the fractions from the unlabeled RDX amended samples. These data indicate these phylotypes were likely involved in carbon or nitrogen uptake from RDX. All three phylotypes were among the most abundant microorganisms in the RDX degrading microbial communities. Unexpectedly, Brevundimonas was not enriched in the heavy fractions of soil 10 and this may be attributed by the inability of this microorganism to uptake the ring nitrogen or carbon. Consistent with the fracion data for soil 10, many RDX degrading Rhodococcus isolates have been identified, including Rhodococcus rhodochrous 11Y (56), Rhodococcus sp. strain DN22 (57-60), Rhodococcus strain YH1 (61), and others (62-64). A recent SIP study also found Rhodococcus sp. to be associated with RDX degradation in explosive contaminated soils (26). Although previous research has reported RDX degradation by Rhodococcus under aerobic conditions, one report indicated RDX degradation under oxygen depleted or anaerobic conditions (63). The other genera associated with label uptake in soil 10, Sedimentibacter and Pusillimonas, have not previously been associated with RDX biodegradation. Members of the genus Sedimentibacter were found to play a role in DDT transformation under iron reducing conditions 83 (65) and were one of the dominant microorganisms in a hydrocarbon contaminated aquifer (66). Members of the genus Pusillimonas, (Proteobacteria, Betaproteobacteria, Burkholderiales, Alcaligenaceae) were dominant members of microbial community capable of concurrently degrading both phenanthrene and arsenite (67). They were also associated to degradation of synthetic and natural estrogen (68) and high molecular weight PAHs (69). In summary, the high throughput sequencing data provided valuable information on which phylotypes increased in abundance following RDX degradation compared to the initial soil and microcosms not amended with RDX. The most notable trend was the common increase in abundance of Brevundimonas and/or unclassified Bacillaceae 1 in the four soils studied. The data indicate these two phylotypes are benefiting from RDX degradation under these conditions. Future research is needed to determine the important of these phylotypes at RDX contaminated sites. 84 APPENDIX 85 Table 4.1. The properties of the four agricultural soils used in this study. Soil 9 10 11 12 Crop Corn Corn Alfalfa Corn Organic Matter % 1.2 1 1.5 2.2 Sand % 70 76 80 86 Silt % 24 22 20 14 Clay % 6 2 0 0 Textural Classification Sandy Loam Loamy Sand Loamy Sand Sand Soil pH 4.9 5.5 6.4 6.1 Calcium mg/kg 489 787 811 1312 Magnesium mg/kg 68 42 74 242 Potassium mg/kg 151 108 173 139 Phosphorus mg/kg 46 79 106 268 CEC meq/10 0g 5 5 4.6 8.6 Table 4.2. A summary of MiSEQ Illumina data generated from all genomic total DNA samples and fractions for Soil 9 and Soil 10. Total DNA samples for four soils (No RDX Time 0 and t, Labeled 1 & 2, Unlabeled 1 &2 ) (24 samples) Triplicates, 8 fractions, Soil 9, Labeled 1 (23 samples) Triplicates, 8 fractions, Soil 9, Unlabeled 1 (23 samples) Triplicates, 8 fractions, Soil 10, Labeled 1 (23 samples) Triplicates, 8 fractions, Soil 10, Unlabeled 1 (23 samples) # of Final # of Final # of % Sequences Unique Sequences Chimeric Following Sequences Make Contigs Command 2,845,956 75,964 1,570,985 10.00 2,304,194 27,609 1,640,354 1.20 2,275,419 27,118 1,646,062 1.20 4,098,173 35,706 3,123,322 0.63 3,286,585 25,614 2,516,050 0.33 86 OM Level High Low Low Low Table 4.3. The classification of anaerobic RDX degrading isolates from previous studies. Class Phylum Firmicutes Order Family Full name Reference(s) Clostridia Clostridiales Clostridiaceae Clostridium bifermentans (1) Clostridia Clostridiales Clostridiaceae (4, 5) Clostridia Clostridia Clostridia Clostridiales Clostridiales Clostridiales Clostridiaceae Clostridiaceae Clostridiaceae Clostridia Clostridia Bacilli Phylum Fusobacteria Clostridiales Clostridiales Bacillales Clostridiaceae Peptococcaceae Bacillaceae Clostridium bifermentans (HAW-1, HAW-G3, HAW-G4, HAW- E3 and, HAW-EC1) Clostrdium sp. HAW-EB17 Clostridium sp. EDB2 Clostridium acetobutylicum Acetobacterium malicum sp. HAAP-1 Acetobacterium paludosum Desulfitobacterium sp. Bacillus sp. HPB2 and HPB3 Fusobacteria Fusobacteriales Fusobacteriaceae Fusobacteria HAW-21 (11) Enteriobacteriales Enteriobacteriales Enteriobacteriales Enteriobacteriales Enteriobacteriales Enteriobacteriales Pseudomonadales Alteromonadales Xanthomonadales Desulfovibrionales Desulfuromonadales Myxococcales Enteriobacteriaceae Enteriobacteriaceae Enteriobacteriaceae Enteriobacteriaceae Enteriobacteriaceae Enteriobacteriaceae Pseudomonadaceae Shewanellaaceae Xanthomonadaceae Desulfovibrionaceae Geobacteraceae Cystobacteraceae Serratia sp. Enterobacter sp. Citrobacter sp. Morganella sp. Klebsiella sp. Providencia sp. Pseudomonas sp. Shewenella sp. Strenophomonas sp. Desulfovibrio sp. Geobacter sp. Anaeromxyobacter sp. (15) (12, 14) (13) (13, 14) (16) (13) (8) (11, 18, 19) (20) (11, 22, 23) (21) (9) (11) (2) (3) (3, 7) (6) (9) (8) Phylum Proteobacteria Gammaproteobacteria Gammaproteobacteria Gammaproteobacteria Gammaproteobacteria Gammaproteobacteria Gammaproteobacteria Gammaproteobacteria Gammaproteobacteria Gammaproteobacteria Deltaproteobacteria Deltaproteobacteria Deltaproteobacteria 87 25 25 RDX Concentration (mg/L) Aa 20 20 15 15 10 10 5 5 bB CONTROL LABELED UNLABELED 0 0 1 31 32 1 43 25 c C 20 20 15 15 10 10 5 5 0 15 16 23 25 dD 0 1 26 27 39 1 19 20 28 Time since RDX was ammended (days) Figure 4.1. RDX in microcosms amended with labeled RDX (13C3, 15N3), unlabeled RDX (12C3, 14 N3), and killed abiotic controls for Soil 9 (A), Soil 10 (B), Soil 11 (C), Soil 12 (D). Error bars denote standard deviation from triplicate micrososms. Arrows indicate hypothetical RDX concentrations on the day of respike. 88 100% 90% 80% 70% 60% 50% 40% 30% 20% 10% Soil 9 Soil 10 Soil 11 Unlabeled 2 Time(t) Labeled 2 Time(t) Unlabeled 1 Time(t) Labeled 1 Time(t) No RDX Time(t) No RDX Time(0) Unlabeled 2 Time(t) Labeled 2 Time(t) Unlabeled 1 Time(t) Labeled 1 Time(t) No RDX Time(t) No RDX Time(0) Unlabeled 2 Time(t) Labeled 2 Time(t) Unlabeled 1 Time(t) Labeled 1 Time(t) No RDX Time(t) No RDX Time(0) Unlabeled 2 Time(t) Labeled 2 Time(t) Unlabeled 1 Time(t) Labeled 1 Time(t) No RDX Time(t) No RDX Time(0) 0% Soil 12 Firmicutes Proteobacteria Actinobacteria unclassified Bacteroidetes Acidobacteria Verrucomicrobia Chloroflexi Gemmatimonadetes Others Figure 4.2. Phyla level comparison between the four soils of the communities in 1) the initial soils (No RDX, Time 0), 2) replicates of the labeled and unlabeled RDX amended microcosms (Labeled 1 & 2, Unlabeled 1 & 2, Time t) and 3) microcosms without RDX addition (No RDX, Time t). 89 A Figure 4.3. MLTree Map (A) and heatmap analysis (B) of the phylotypes within the initial soils and in samples with and without RDX. To simplify the presentation, only one replicate was used for the RDX amended samples. 90 Figure 4.3.(cont’d) B 91 A 6000 Soil 9 5000 Number of OTUs 4000 Soil 10 3000 Soil 11 2000 1000 Soil 12 No RDX Time(0) No RDX Time(t) Labeled 1 Time(t) Unlabeled 1 Time(t) Labeled 2 Time(t) Unlabeled 2 Time(t) No RDX Time(0) No RDX Time(t) Labeled 1 Time(t) Unlabeled 1 Time(t) Labeled 2 Time(t) Unlabeled 2 Time(t) No RDX Time(0) No RDX Time(t) Labeled 1 Time(t) Unlabeled 1 Time(t) Labeled 2 Time(t) Unlabeled 2 Time(t) No RDX Time(0) No RDX Time(t) Labeled 1 Time(t) Unlabeled 1 Time(t) Labeled 2 Time(t) Unlabeled 2 Time(t) 0 0 50000 100000 150000 200000 Number of sequences B 0.4 S9RDXTDNA PC2 percent variation explained 20.9 % 0.3 Soil 9 S12RDXTDNA 0.2 NoRDXt 0.1 Soil 10 0 -0.8 -0.6 -0.4 -0.2 0 0.2 -0.1 0.4 S11RDXTD NoRDXt0 Soil 11 NA -0.2 S10RDXTD NA No RDX Time(0) No RDX Time(t) Labeled 1 Time(t) Unlabeled 1 Time(t) Labeled 2 Time(t) Unlabeled 2 Time(t) No RDX Time(0) No RDX Time(t) Labeled 1 Time(t) Unlabeled 1 Time(t) Labeled 2 Time(t) Unlabeled 2 Time(t) No RDX Time(0) No RDX Time(t) Labeled 1 Time(t) Unlabeled 1 Time(t) Labeled 2 Time(t) Unlabeled 2 Time(t) No RDX Time(0) No RDX Time(t) Labeled 1 Time(t) Unlabeled 1 Time(t) Labeled 2 Time(t) Unlabeled 2 Time(t) -0.3 Soil 12 -0.4 PC1 percent variation explained 30.3 % Figure 4.4. Rarefaction curves (A) and Principal Component Analysis (B) of sequencing data generated from all four soils, including total DNA extracts of the labeled RDX (Labeled 1 & 2, Time t), unlabeled RDX (Unlabeled 1 & 2, Time t), no RDX (No RDX, Time t) microcosms and the initial soils (No RDX, Time 0). 92 0 93 unclassified Bacteria Gemmatimonas unclassified Bacteria unclassified Chitinophagaceae 5 unclassified Chitinophagaceae 10 Chloroflexi Verrucomicrobia Gemmatimonadetes Bacteroidetes Acidobacteria Spartobacteria Soil 12 Ktedonobacter Acidobacteria Gp16 Acidobacteria Gp6 Acidobacteria Gp4 Acidobacteria Gp3 Acidobacteria Gp1 unclassified Actinobacteria 20 Pedobacter unclassified Actinomycetales unclassified Solirubrobacterales A. No RDX, Time 0 (initial soil) Chryseobacterium Arthrobacter Rhodococcus Proteobacteria unclassified Xanthomonadaceae unclassified Microbacteriaceae unclassified Intrasporangiaceae Actinobacteria Rhodanobacter unclassified Cystobacteraceae unclassified Alcaligenaceae unclassified Burkholderiales unclassified Nocardioidaceae unclassified Micromonosporaceae Soil 11 Methylophilus Firmicute Pusillimonas Arthrobacter Pseudonocardia Soil 10 Massilia unclassified Sphingomonadaceae Aeromicrobium Solirubrobacter Conexibacter unclassified Proteobacteria Proteobacteria Brevundimonas Methylophilus unclassified Betaproteobacteria Soil 9 unclassified Firmicutes unclassified Bacillales unclassified Rhizobiales unclassified Alphaproteobacteria 25 unclassified Lachnospiraceae unclassified Bacillaceae 1 unclassified Erysipelotrichaceae Soil 11 unclassified Clostridiaceae 1 unclassified Sphingomonadaceae Sphingomonas Hyphomicrobium unclassified Bacillales unclassified Bacillaceae 1 Paenibacillus Lysinibacillus Firmicute unclassified Ruminococcaceae Soil 10 unclassified Lachnospiraceae 30 Sporomusa Proteiniborus Soil 9 Clostridium III Lysinibacillus 15 Bacillus Tumebacillus 15 Sedimentibacter 20 Tissierella Clostridium sensu stricto 0 Clostridium XlVa Relative Abundance (%) 30 B. No RDX, Time T 25 Actinobacteria Bacteroidetes Soil 12 10 5 Figure 4 5. A comparison of the relative abundance (%) of the most common phylotypes (>1% , at the lowest classification level) in all four soils in the no RDX samples, including the initial soils (No RDX, Time 0) (A) and the microcosms without RDX (No RDX, Time t) (B). 0 94 Unlabeled Rep 1 20 Proteobacteria Labeled Rep 2 unclassified Bacteria Proteobacteria unclassified Bacteria Arthrobacter Rhodococcus Labeled Rep 2 unclassified Bacteria Proteobacteria unclassified Bacteria Firmicutes Chryseobacterium Proteobacteria Pedobacter Rhodococcus unclassified Oxalobacteraceae unclassified Cystobacteraceae Dokdonella unclassified Firmicutes unclassified Bacillales unclassified Lachnospiraceae unclassified Ruminococcaceae unclassified Erysipelotrichaceae unclassified Bacillaceae 1 unclassified Clostridiaceae 1 Unlabeled Rep 1 Flavobacterium Labeled Rep 1 Rhodococcus Rhodanobacter Brevundimonas Firmicutes Rhodococcus unclassified Alcaligenaceae Brevundimonas unclassified Firmicutes unclassified Bacillales unclassified Lachnospiraceae unclassified Erysipelotrichaceae unclassified Bacillaceae 1 Clostridium III Clostridium XlVa Tissierella Lysinibacillus Labeled Rep 1 Pseudomonas 40 unclassified Sphingobacteriaceae Firmicutes Brevundimonas unclassified Firmicutes unclassified Bacillales unclassified Firmicutes unclassified Planococcaceae C. Soil 11, RDX Time T unclassified Planococcaceae unclassified Clostridiaceae 1 D. Soil 12, RDX Time T unclassified Lachnospiraceae Labeled Rep 2 unclassified Bacillaceae 1 40 Labeled Rep 2 unclassified Clostridiaceae 1 Unlabeled Rep 1 unclassified Clostridiaceae 1 Unlabeled Rep 1 unclassified Bacillaceae 1 unclassified Clostridiaceae 1 B. Soil 10, RDX Time T unclassified Erysipelotrichaceae Tissierella Pusillimonas 60 Sedimentibacter A. Soil 9, RDX Time T Lysinibacillus 0 Tissierella Labeled Rep 1 Pusillimonas 60 Labeled Rep 1 Clostridium XlVa 60 Tissierella 80 Sedimentibacter 0 Sedimentibacter 40 Sedimentibacter Relative Abundance (%) 40 30 20 10 0 Firmicutes Unlabeled Rep 2 Actinobacteria Unlabeled Rep 2 20 Actinobacteria Bacteroidetes Unlabeled Rep 2 20 Actinobacteria 80 Actinobacteria Bacteroidetes Unlabeled Rep 2 Figure 4.6. A comparison of the relative abundance (%) of the most common phylotypes (>1% , at the lowest classification level) in all four soils (A-D) following RDX degradation. 30 Soil 9 Initial No RDX Labeled Rep 1 Unlabeled Rep 1 Labeled Rep 2 Unlabeled Rep 2 20 10 0 unclassified Bacillaceae 1 60 unclassified Clostridiaceae 1 unclassified Cystobacteraceae Rhodococcus Lysinibacillus Tissierella unclassified Erysipelotrichaceae Soil 10 Initial 40 Relative Abundance (%) Sedimentibacter No RDX Labeled Rep 1 Unlabeled Rep 1 Labeled Rep 2 Unlabeled Rep 2 20 0 Brevundimonas unclassified Erysipelotrichaceae Tissierella Pusillimonas Sedimentibacter Rhodococcus unclassified Clostridiaceae 1 unclassified Lachnospiraceae 80 60 Soil 11 Initial No RDX Labeled Rep 1 Unlabeled Rep 1 Labeled Rep 2 Unlabeled Rep 2 40 20 0 Brevundimonas unclassified Bacillaceae 1 unclassified Alcaligenaceae Tissierella unclassified Clostridiaceae 1 Rhodococcus unclassified Planococcaceae Sedimentibacter 60 40 Initial No RDX Labeled Rep 1 Unlabeled Rep 1 Labeled Rep 2 Unlabeled Rep 2 20 0 unclassified Bacillaceae 1 Brevundimonas unclassified Sedimentibacter Erysipelotrichaceae unclassified Firmicutes unclassified Clostridiaceae 1 Flavobacterium Rhodococcus Figure 4.7. The relative abundance (%) of the eight most abundant phylotypes in the four RDX amended agricultural soils (Labeled 1 &2, Unlabeled 1 &2, Time t) compared to the RDX controls (initial soil and no RDX microcosms). Boxed phylotypes illustrate a greater abundance in RDX amended samples compared to the RDX controls for at least three soils. The phylotypes are at genus level, unless unclassified. 95 A. Soil 9 20 B. Soil 10 unclassified Bacillaceae 1 15 Pusillimonas 15 10 Unlabeled 10 Labeled 5 5 Relative Abundance (%) 0 1.73 12 1.74 1.75 1.76 1.77 0 1.73 1.74 1.75 1.76 1.77 1.78 1.79 1.78 unclassified Clostridiales 15 8 10 4 5 0 1.73 10 1.74 1.75 1.76 1.77 Rhodococcus 0 1.73 1.74 1.75 1.76 1.77 1.78 1.79 1.78 Tissierella 10 Sedimentibacter 8 6 5 4 2 0 1.73 1.74 1.75 1.76 1.77 0 1.73 1.74 1.75 1.76 1.77 1.78 1.79 1.78 Buoyant Density (g mL-1) Figure 4.8. Relative abundance of the most common phylotypes in heavy fractions from the labeled RDX amended microcosms compared to those from the unlabeled RDX amended microcosms in Soil 9 (A) and Soil 10 (B). Bars illustrate standard deviations from triplicate sequencing data. 96 REFERENCES 97 REFERENCES 1. Regan KM, Crawford RL. 1994. Characterization of Clostridium bifermentans and its biotransformation of 2,4,6-trinitrotoluene (TNT) and 1,3,5-triaza-1,3,5trinitrocyclohexane (RDX). Biotechnology Letters 16:1081-1086. 2. Bhushan B, Halasz A, Thiboutot S, Ampleman G, Hawari J. 2004. Chemotaxismediated biodegradation of cyclic nitramine explosives RDX, HMX, and CL-20 by Clostridium sp. EDB2. Biochemical Biophysical Research Communications 316:816821. 3. Zhang CL, Hughes JB. 2003. Biodegradation pathways of hexahydro-1,3,5-trinitro1,3,5-triazine (RDX) by Clostridium acetobutylicum cell-free extract. Chemosphere 50:665-671. 4. Zhao JS, Paquet L, Halasz A, Hawari J. 2003. Metabolism of hexahydro-1,3,5-trinitro1,3,5-triazine through initial reduction to hexahydro-1-nitroso-3,5-dinitro-1,3,5-triazine followed by denitration in Clostridium bifermentans HAW-1. Applied Microbiology and Biotechnology 63:187-193. 5. Zhao JS, Spain J, Hawari M. 2003. Phylogenetic and metabolic diversity of hexahydro1,3,5-trinitro-1,3,5-triazine (RDX)-transforming bacteria in strictly anaerobic mixed cultures enriched on RDX as nitrogen source. Fems Microbiology Ecology 46:189-196. 6. Sherburne LA, Shrout JD, Alvarez PJJ. 2005. Hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) degradation by Acetobacterium paludosum. Biodegradation 16:539-547. 7. Adrian NR, Arnett CM. 2004. Anaerobic biodegradation of hexahydro-1,3,5-trinitro1,3,5-triazine (RDX) by Acetobacterium malicum strain HAAP-1 isolated from a methanogenic mixed culture. Current Microbiology 48:332-340. 8. Singh R, Soni P, Kumar P, Purohit S, Singh A. 2009. Biodegradation of high explosive production effluent containing RDX and HMX by denitrifying bacteria. World Journal of Microbiology & Biotechnology 25:269-275. 9. Kwon MJ, Finneran KT. 2008. Hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) and octahydro-1,3,5,7-tetranitro-1,3,5,7-tetrazocine (HMX) biodegradation kinetics amongst several Fe(III)-reducing genera. Soil & Sediment Contamination 17:189-203. 10. Zhao JS, Paquet L, Halasz A, Manno D, Hawari M. 2004. Metabolism of octahydro1,3,5,7-tetranitro-1,3,5,7-tetrazocine by Clostridium bifermentans strain HAW-1 and several other H2 producing fermentative anaerobic bacteria. FEMS Microbiology Letters 237:65-72. 98 11. Zhao JS, Spain J, Thiboutot S, Ampleman G, Greer C, Hawari J. 2004. Phylogeny of cyclic nitramine-degrading psychrophilic bacteria in marine sediment and their potential role in the natural attenuation of explosives. Fems Microbiology Ecology 49:349-357. 12. Pudge IB, Daugulis AJ, Dubois C. 2003. The use of Enterobacter cloacae ATCC 43560 in the development of a two-phase partitioning bioreactor for the destruction of hexahydro-1,3,5-trinitro-1,3,5-s-triazine (RDX). Journal of Biotechnology 100:65-75. 13. Kitts CL, Cunningham DP, Unkefer PJ. 1994. Isolation of three hexahydro-1,3,5trinitro-1,3,5-triazine-degrading species of the family Enterobacteriaceae from nitramine explosive-contaminated soil. Applied and Environmental Microbiology 60:4608-4611. 14. Kitts CL, Green CE, Otley RA, Alvarez MA, Unkefer PJ. 2000. Type I nitroreductases in soil enterobacteria reduce TNT (2,4,6-trinitrotoluene) and RDX (hexahydro-1,3,5-trinitro-1,3,5-triazine). Canadian Journal of Microbiology 46:278-282. 15. Young DM, Unkefer PJ, Ogden KL. 1997. Biotransformation of hexahydro-1,3,5trinitro-1,3,5-triazine (RDX) by a prospective consortium and its most effective isolate Serratia marcescens. Biotechnology and Bioengineering 53:515-522. 16. Zhao JS, Halasz A, Paquet L, Beaulieu C, Hawari J. 2002. Biodegradation of hexahydro-1,3,5-trinitro-1,3,5-triazine and its mononitroso derivative hexahydro-1nitroso-3,5-dinitro-1,3,5-triazine by Klebsiella pneumoniae strain SCZ-1 isolated from an anaerobic sludge. Applied and Environmental Microbiology 68:5336-5341. 17. Cho YS, Lee BU, Oh KH. 2008. Simultaneous degradation of nitroaromatic compounds TNT, RDX, atrazine, and simazine by Pseudomonas putida HK-6 in bench-scale bioreactors. Journal of Chemical Technology and Biotechnology 83:1211-1217. 18. Zhao JS, Manno D, Beaulieu C, Paquet L, Hawari J. 2005. Shewanella sediminis sp nov., a novel Na+-requiring and hexahydro-1,3,5-trinitro-1,3,5-trinitro-degrading bacterium from marine sediment. International Journal of Systematic and Evolutionary Microbiology 55:1511-1520. 19. Zhao JS, Manno D, Leggiadro C, O'Neill D, Hawari J. 2006. Shewanella halifaxensis sp nov., a novel obligately respiratory and denitrifying psychrophile. International Journal of Systematic and Evolutionary Microbiology 56:205-212. 20. Binks PR, Nicklin S, Bruce NC. 1995. Degradation of hexahydro-1,3,5-trinitro-1,3,5triazine (RDX) by Stenotrophomonas maltophilia Pb1. Applied and Environmental Microbiology 61:1318-1322. 21. Kwon MJ, Finneran KT. 2008. Biotransformation products and mineralization potential for hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) in abiotic versus biological degradation pathways with anthraquinone-2,6-disulfonate (AQDS) and Geobacter metallireducens. Biodegradation 19:705-715. 99 22. Arnett CM, Adrian NR. 2009. Cosubstrate independent mineralization of hexahydro1,3,5-trinitro-1,3,5-triazine (RDX) by a Desulfovibrio species under anaerobic conditions. Biodegradation 20:15-26. 23. Boopathy R, Gurgas M, Ullian J, Manning JF. 1998. Metabolism of explosive compounds by sulfate-reducing bacteria. Current Microbiology 37:127-131. 24. Fuller ME, Steffan RJ. 2008. Groundwater chemistry and microbial ecology effects on explosives biodegradation. SERDP Final Report Project ER-1378. 25. Amann RI, Ludwig W, Schleifer KH. 1995. Phylogenetic identification and in-situ detection of individual microbial cells without cultivation. Microbiol Rev 59:143-169. 26. Andeer P, Stahl DA, Lillis L, Strand SE. 2013. Identification of microbial populations assimilating nitrogen from RDX in munitions contaminated military training range soils by high sensitivity stable isotope probing. Environmental Science & Technology 47:10356-10363. 27. Cho KC, Lee DG, Roh H, Fuller ME, Hatzinger PB, Chu K-H. 2013. Application of 13 C-stable isotope probing to identify RDX-degrading microorganisms in groundwater. Environmental Pollution 178:350-360. 28. Jayamani I, Manzella MP, Cupples AM. 2013. RDX degradation potential in soils previously unexposed to RDX and the identification of RDX-degrading species in one agricultural soil using stable isotope probing. Water Air and Soil Pollution 224. 29. Roh H, Yu CP, Fuller ME, Chu KH. 2009. Identification of hexahydro-1,3,5-trinitro1,3,5-triazine-degrading microorganisms via 15N-stable isotope probing. Environmental Science & Technology 43:2505-2511. 30. Pilloni G, von Netzer F, Engel M, Lueders T. 2011. Electron acceptor-dependent identification of key anaerobic toluene degraders at a tar-oil-contaminated aquifer by Pyro-SIP. Fems Microbiology Ecology 78:165-175. 31. Xia Y, Cai L, Zhang T, Fang HHP. 2012. Effects of substrate loading and co-substrates on thermophilic anaerobic conversion of microcrystalline cellulose and microbial communities revealed using high-throughput sequencing. International Journal of Hydrogen Energy 37:13652-13659. 32. Bell TH, Yergeau E, Martineau C, Juck D, Whyte LG, Greer CW. 2011. Identification of nitrogen-incorporating Bacteria in petroleum-contaminated Arctic soils by using [N-15] DNA-based stable isotope probing and pyrosequencing. Applied and Environmental Microbiology 77:4163-4171. 33. Song Y, Xiao L, Jayamani I, He Z, Cupples AM. In Press. A novel method to characterize bacterial communities affected by carbon source and electricity generation in microbial fuel cells using stable isotope probing and Illumina sequencing. . Journal of Microbiological Methods. 100 34. Thompson KT, Crocker FH, Fredrickson HL. 2005. Mineralization of the cyclic nitramine explosive hexahydro-1,3,5-trinitro-1,3,5-triazine by Gordonia and Williamsia spp. Applied and Environmental Microbiology 71:8265-8272. 35. Sun WM, Cupples AM. 2012. Diversity of five anaerobic toluene-degrading microbial communities investigated using stable isotope probing. Applied and Environmental Microbiology 78:972-980. 36. Sun WM, Sun XX, Cupples AM. 2012. Anaerobic methyl tert-butyl ether-degrading microorganisms identified in wastewater treatment plant samples by stable isotope probing. Applied and Environmental Microbiology 78:2973-2980. 37. Sun WM, Sun XX, Cupples AM. 2014. Identification of Desulfosporosinus as tolueneassimilating microorganisms from a methanogenic consortium. Int Biodeter Biodegr 88:13-19. 38. Sun WM, Xie SG, Luo CL, Cupples AM. 2010. Direct Link between Toluene Degradation in Contaminated-Site Microcosms and a Polaromonas Strain. Applied and Environmental Microbiology 76:956-959. 39. Xie SG, Sun WM, Luo CL, Cupples AM. 2010. Stable Isotope Probing Identifies Novel m-Xylene Degraders in Soil Microcosms from Contaminated and Uncontaminated Sites. Water Air and Soil Pollution 212:113-122. 40. Xie SG, Sun WM, Luo CL, Cupples AM. 2011. Novel aerobic benzene degrading microorganisms identified in three soils by stable isotope probing. Biodegradation 22:7181. 41. Caporaso JG, Lauber CL, Walters WA, Berg-Lyons D, Huntley J, Fierer N, Owens SM, Betley J, Fraser L, Bauer M, Gormley N, Gilbert JA, Smith G, Knight R. 2012. Ultra-high-throughput microbial community analysis on the Illumina HiSeq and MiSeq platforms. ISME Journal 6:1621-1624. 42. Caporaso JG, Lauber CL, Walters WA, Berg-Lyons D, Lozupone CA, Turnbaugh PJ, Fierer N, Knight R. 2011. Global patterns of 16S rRNA diversity at a depth of millions of sequences per sample. P Natl Acad Sci USA 108:4516-4522. 43. Schloss PD. 2009. A high-throughput DNA sequence aligner for microbial ecology studies. Plos One 4. 44. Kozich JJ, Westcott SL, Baxter NT, Highlander SK, Schloss PD. 2013. Development of a dual-index sequencing strategy and curation pipeline for analyzing amplicon sequence data on the MiSeq Illumina Sequencing Platform. Applied and Environmental Microbiology 79:5112-5120. 45. Edgar RC, Haas BJ, Clemente JC, Quince C, Knight R. 2011. UCHIME improves sensitivity and speed of chimera detection. Bioinformatics 27:2194-2200. 101 46. Meyer F, Paarmann D, D'Souza M, Olson R, Glass EM, Kubal M, Paczian T, Rodriguez A, Stevens R, Wilke A, Wilkening J, Edwards RA. 2008. The metagenomics RAST server - a public resource for the automatic phylogenetic and functional analysis of metagenomes. Bmc Bioinformatics 9. 47. Ringelberg D, Richmond M, Foley K, Reynolds C. 2008. Utility of lipid biomarkers in support of bioremediation efforts at army sites. Journal of Microbiological Methods 74:17-25. 48. Bertin L, Capodicasa S, Fedi S, Zannoni D, Marchetti L, Fava F. 2011. Biotransformation of a highly chlorinated PCB mixture in an activated sludge collected from a Membrane Biological Reactor (MBR) subjected to anaerobic digestion. J Hazard Mater 186:2060-2067. 49. Wu YC, Luo YM, Zou DX, Ni JZ, Liu WX, Teng Y, Li ZG. 2008. Bioremediation of polycyclic aromatic hydrocarbons contaminated soil with Monilinia sp.: degradation and microbial community analysis. Biodegradation 19:247-257. 50. Van Aken B, Yoon JM, Schnoor JL. 2004. Biodegradation of nitro-substituted explosives 2,4,6-trinitrotoluene, hexahydro-1,3,5-trinitro-1,3,5-triazine, an octahydro1,3,5,7-tetranitro-1,3,5-tetrazocine by a phytosymbiotic Methylobacterium sp associated with poplar tissues (Populus deltoides x nigra DN34). Applied and Environmental Microbiology 70:508-517. 51. Deshpande NM, Sarnaik SS, Paranjpe SA, Kanekar PP. 2004. Optimization of dimethoate degradation by Brevundimonas sp MCM B-427 using factorial design: studies on interactive effects of environmental factors. World Journal of Microbiology & Biotechnology 20:455-462. 52. Konno N, Habu N, Maeda I, Azuma N, Isogai A. 2006. Cellouronate (beta-1,4-linked polyglucuronate) lyase from Brevundimonas sp SH203: Purification and characterization. Carbohyd Polym 64:589-596. 53. Chaineau CH, Morel J, Dupont J, Bury E, Oudot J. 1999. Comparison of the fuel oil biodegradation potential of hydrocarbon-assimilating microorganisms isolated from a temperate agricultural soil. Sci Total Environ 227:237-247. 54. Chang YT, Lee JF, Chao HP. 2007. Variability of communities and physiological characteristics between free-living bacteria and attached bacteria during the PAH biodegradation in a soil/water system. Eur J Soil Biol 43:283-296. 55. Phillips LA, Germida JJ, Farrell RE, Greer CW. 2008. Hydrocarbon degradation potential and activity of endophytic bacteria associated with prairie plants. Soil Biology & Biochemistry 40:3054-3064. 56. Seth-Smith HMB, Rosser SJ, Basran A, Travis ER, Dabbs ER, Nicklin S, Bruce NC. 2002. Cloning, sequencing, and characterization of the hexahydro-1,3,5-trinitro-1,3,5- 102 triazine degradation gene cluster from Rhodococcus rhodochrous. Applied and Environmental Microbiology 68:4764-4771. 57. Bhushan B, Trott S, Spain JC, Halasz A, Paquet L, Hawari M. 2003. Biotransformation of hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) by a rabbit liver cytochrome p450: insight into the mechanism of RDX biodegradation by Rhodococcus sp strain DN22. Applied and Environmental Microbiology 69:1347-1351. 58. Coleman NV, Nelson DR, Duxbury T. 1998. Aerobic biodegradation of hexahydro1,3,5-trinitro-1,3,5-triazine (RDX) as a nitrogen source by a Rhodococcus sp., strain DN22. Soil Biology & Biochemistry 30:1159-1167. 59. Coleman NV, Spain JC, Duxbury T. 2002. Evidence that RDX biodegradation by Rhodococcus strain DN22 is plasmid-borne and involves a cytochrome p-450. Journal of Applied Microbiology 93:463-472. 60. Fournier D, Halasz A, Spain J, Fiurasek P, Hawari J. 2002. Determination of key metabolites during biodegradation of hexahydro-1,3,5-trinitro-1,3,5-triazine with Rhodococcus sp strain DN22. Applied and Environmental Microbiology 68:166-172. 61. Nejidat A, Kafka L, Tekoah Y, Ronen Z. 2008. Effect of organic and inorganic nitrogenous compounds on RDX degradation and cytochrome P-450 expression in Rhodococcus strain YH1. Biodegradation 19:313-320. 62. Bernstein A, Adar E, Nejidat A, Ronen Z. 2011. Isolation and characterization of RDX-degrading Rhodococcus species from a contaminated aquifer. Biodegradation 22:997-1005. 63. Fuller ME, Perreault N, Hawari J. 2010. Microaerophilic degradation of hexahydro1,3,5-trinitro-1,3,5-triazine (RDX) by three Rhodococcus strains. Letters in Applied Microbiology 51:313-318. 64. Seth-Smith HMB, Edwards J, Rosser SJ, Rathbone DA, Bruce NC. 2008. The explosive-degrading cytochrome P450 system is highly conserved among strains of Rhodococcus spp. Applied and Environmental Microbiology 74:4550-4552. 65. Chen MJ, Cao F, Li FB, Liu CS, Tong H, Wu WJ, Hu M. 2013. Anaerobic transformation of DDT related to iron(III) reduction and microbial community structure in paddy soils. J Agr Food Chem 61:2224-2233. 66. Tischer K, Kleinsteuber S, Schleinitz KM, Fetzer I, Spott O, Stange F, Lohse U, Franz J, Neumann F, Gerling S, Schmidt C, Hasselwander E, Harms H, Wendeberg A. 2013. Microbial communities along biogeochemical gradients in a hydrocarboncontaminated aquifer. Environ Microbiol 15:2603-2615. 67. Tang JZ, Feng TC, Cui CZ, Feng YY. 2013. Simultaneous biodegradation of phenanthrene and oxidation of arsenite by a dual-functional bacterial consortium. Int Biodeter Biodegr 82:173-179. 103 68. Muller M, Patureau D, Godon JJ, Delgenes JP, Hernandez-Raquet G. 2010. Molecular and kinetic characterization of mixed cultures degrading natural and synthetic estrogens. Applied Microbiology and Biotechnology 85:691-701. 69. Llado S, Gracia E, Solanas AM, Vinas M. 2013. Fungal and bacterial microbial community assessment during bioremediation assays in an aged creosote-polluted soil. Soil Biology & Biochemistry 67:114-123. 104 CHAPTER 5 STABLE ISOTOPE PROBING REVEALS THE IMPORTANCE OF COMAMONAS AND PSEUDOMONADACEAE IN RDX DEGRADATION IN SAMPLES FROM A NAVY DETONATION SITE ABSTRACT This study investigated the microorganisms involved in RDX degradation from a detonation area at a Navy Base. Using Illumina sequencing, microbial communities were compared between the initial sample, samples following RDX degradation and controls not amended with RDX to determine which phylotypes increased in abundance following RDX degradation. The effect of glucose on these communities was also examined. In addition, stable isotope probing (SIP) using labeled (13C3, 15N3) RDX was performed. Illumina sequencing revealed several phylotypes were more abundant following RDX degradation compared to the initial soil and the no RDX controls. For the glucose amended samples, this trend was strong for an unclassified Pseudomonadeae phylotype and for Comamonas. Without glucose, Acinetobacter illustrated the greatest increase following RDX degradation compared to the initial soil and no RDX controls. Rhodococcus, a known RDX degrader, also increased in abundance following RDX degradation. For the SIP study, unclassified Pseudomonadaceae was the most abundant phylotype in the heavy fractions in both the presence and absence of glucose. In the glucose amended heavy 105 fractions, Comamonas and Anaeromxyobacter were also present. Without glucose, the heavy fractions also contained Azohydromonas and Rhodococcus. However, all four phylotypes were present at a much lower level compared to unclassified Pseudomonadaceae. Overall, these data indicate unclassified Pseudomonadaceae were primarily responsible for label uptake in both treatments. When glucose was present, Comamonas increased in abundance following RDX degradation and was enriched in the heavy fractions, suggesting, for the first time, that this phylotype is also important for RDX removal. INTRODUCTION RDX (hexahydro-1,3,5-trinitro-1,3,5-triazine) is an explosive that has caused widespread soil and water contamination at many military sites in the US and worldwide. Such contamination has been associated with manufacturing and load-assemble-package processes performed during or after World War II and the Korean Conflict. Remediation of these sites has been initiated since the early 1980s, however many still have groundwater contaminated with nitroaromatics. The U.S. EPA has established a health advisory level in drinking water for RDX (2 μg/L), indicating the potential threat to humans. RDX is susceptible to aerobic or anaerobic biodegradation (1-10), therefore bioremediation offers a plausible approach for site clean-up. Anaerobic RDX degradation can occur via hexahydro-1-nitroso-3,5-dinitro-1,3,5-triazine (MNX), hexahydro-1,3-dinitroso-5-nitro-1,3,5-triazine (DNX), and hexahydro-1,3,5-trinitroso1,3,5-triazine (TNX) (11), forming methanol and formaldehyde. Denitration has also been reported under anaerobic conditions, involving ring cleavage and methylene denitramine (MEDINA) formation (5, 9). Under aerobic conditions, denitration appears to be a common pathway (12). RDX denitration has been coupled with ring cleavage generating 4-nitro-2,4- 106 diazabutanal (NDAB) as a ring cleavage product (13). MEDINA has also been observed as a ring cleavage product (14, 15). In all, known RDX degrading aerobic or anaerobic isolates classify within the phyla Firmicutes, Actinobacteria, Proteobacteria (Alpha-, Gamma- and Deltaproteobacteria) and Fusobacteria (12). However, researchers have reported the microorganisms detected in field samples actively degrading RDX were generally not closely related to previously reported RDX degrading bacterial strains (16). The identification of RDX degraders at these sites is important for 1) determining if natural attenuation is a feasible approach across sites and 2) monitoring the populations of the microorganisms involved. Although much is known about the pure cultures able to degrade RDX in the laboratory, very little is known about the microorganisms responsible for RDX degradation at contaminated sites. One group concluded that previously reported RDX degrading bacteria did not capture the microbial diversity associated with RDX bioremediation in groundwater, especially under typical biostimulation approaches (17). Indeed, it is now widely recognized that only a small percentage of microorganisms can be isolated and cultivated in the laboratory (18), therefore, it is likely that in situ RDX degraders have yet to be identified. However, recent advances in molecular biology have enabled a wealth of information concerning a range of microbial processes. A number of molecular approaches have been used to investigate RDX degraders in mixed communities, such as terminal restriction fragment length polymorphism (TRFLP), Sanger sequencing (17, 19-21) and stable isotope probing (SIP) (22-25). SIP is a culture independent method that targets only active organisms and involves sample exposure to a labeled substrate, incubation, nucleic acid extraction, ultracentrifugation to separate the labeled nucleic acid from the unlabeled background nucleic acid, and finally molecular analysis to identify the organism(s) responsible for label 107 uptake. The method is advantageous because there is no requirement to work with pure cultures and so SIP can identify the microorganisms involved in label uptake in mixed cultures, e.g. soil samples. The method has been particularly valuable for detecting microorganisms involved in contaminant degradation (26-30). To date, for RDX degradation studies, SIP has only been combined with TRFLP and/or Sanger sequencing. These approaches have their limitations because it is not always possible to sequence enough clones to identify the phylotypes involved in label uptake. In this study, we advance this methodology by combining SIP with high throughput sequencing to investigate the microorganisms responsible for label uptake from RDX. This is the first report of using both methods to examine RDX degraders. Here, the objective was to identify the microorganisms involved in RDX degradation in soils from a detonation area at a Navy Base. The research also explores the effect of glucose on RDX degradation and the microorganisms involved. RDX biodegradation and bioremediation often involves the addition of a carbon source(31). Glucose was added in this study, based on previous research indicating enhanced RDX degradation in other soil samples when this substrate was added (24). The microbial communities were investigated both in the presence and in the absence of glucose. The research combined SIP and high throughput sequencing to provide two layers of data. Firstly, phylotypes illustrating an increase in relative abundance following RDX degradation compared to controls (initial soil and samples with no RDX added) were determined with Illumina sequencing. Secondly, DNA based SIP (15N and 13C RDX) was used to identify which microorganisms were involved in label uptake from RDX. The wealth of information 108 provided by coupling high throughput sequencing with SIP has enabled the identification of novel RDX degraders and has provided data that can be used to design primers for the detection of these microorganisms at other RDX contaminated sites. METHODS Chemicals Unlabeled and ring- labeled RDX (13C3, 99%; 15N3, 98%) in acetonitrile (1,000 mg/L) were purchased from Cambridge Isotope Laboratories (Andover, MA). Other reagents were purchased from Sigma Aldrich® Corp. (St. Louis, MO) or Thermo Fisher Scientific Inc. (Waltham, MA), unless otherwise stated. HPLC grade acetonitrile was purchased from EMD Millipore (Chemicals), a division of Merck KGaA (Darmstadt, Germany). Experimental Setup and DNA Extraction The soil was collected from a Navy Installation in Virginia. Soil samples were first screened for their ability to degrade RDX. For this, duplicate killed controls (autoclaved) and live microcosms were prepared using 4 mL of a mineral salts media (MSM) (Thompson et al 2005), 1 mL of glucose (1 g L-1), 1.5 g of soil (dry weight) and 10 mg L-1 of unlabeled RDX in amber serum bottles. Microcosms were sealed with rubber stoppers and aluminum seals and stored in the dark (~20 °C) without shaking. For the SIP study, microcosms were prepared with or without glucose. The serum bottles (30 mL) included 2 g of soil with either 4 mL of mineral media and 1 mL of glucose (1 g L-1 stock solution) or 5 mL of MSM. Then, 10 mg L-1 of labeled (13C3, 15N3-RDX) or unlabeled RDX was added. A total of 2 killed controls (sterilized by autoclaving), 4 unlabeled controls (10 mg L-1 unlabeled RDX) and 2 labeled sample microcosms (10 mg L-1 of labeled 109 13 C3, 15N3-RDX) were prepared for each of the two treatments (with or without glucose). Two additional controls (no RDX controls) were also included, both were not amended with RDX and only one was amended with glucose. These controls were incubated under the same conditions as the above samples. Following the degradation of the first addition of RDX (30 days), eight microcosms (replicate microcosms with labeled RDX and glucose, unlabeled RDX and glucose, labeled RDX and no glucose and unlabeled RDX and no glucose) and the two no RDX controls (with and without glucose) were sacrificed and the total genomic DNA was extracted. DNA was extracted using the Power Soil DNA extraction kit (MO BIO Laboratories, Inc. Carlsbad, CA) as per manufacturer’s instructions. In the remaining replicate microcosms initially amended with unlabeled RDX (with and without glucose), a second amendment of 10 mg L-1 of RDX (unlabeled) was added. This was degraded under 20 days and the total genomic DNA was extracted from these samples. DNA was also extracted from the initial soil. In all, fifteen total genomic DNA extracts were submitted for Illumina sequencing (a list is provided below). Analytical Methods RDX extraction and analysis methods were modified from those previously described (Thompson et al 2005). Briefly, 200 µL from each microcosm was removed into 1.7 mL microcentrifuge tubes. RDX was extracted by adding equal volumes of acetonitrile and shaking the tubes at room temperature for 30 minutes. Following this, the tubes were centrifuged at 10,000 rpm for 5 minutes. The supernatant (350 µL) was filtered using 0.22 µm sterile filters into HPLC amber vials (Sigma). The HPLC parameters were as follows: injector volume, 20 µL; 110 isocratic conditions (40% acetonitrile and 60% 0.01% H3PO4-acidified water, 1 mL/min); Perkin Elmers (PE) Series 200 autosampler; PE binary LC Pump 250; Waters UV detector; wavelength 255 nm (detection limit was 500 µg/L). Isopycnic Centrifugation Total genomic DNA extracted from four microcosms with one amendment of RDX was subject to ultracentrifugation and fractioning. This involved two microcosms amended with glucose (labeled and unlabeled RDX amended) and two not amended with glucose (labeled and unlabeled RDX amended). The extracted DNA was ultracentrifuged with cesium chloride and tris-EDTA buffer solution. Specifically, 3 to 5 µg of DNA was added to a Tris-EDTA (pH 8.0) buffer and cesium chloride solution. The initial buoyant density (BD) of this mixture was adjusted to ~1.72 g mL-1 using a model AR200 digital refractometer (Leica Microsystems Inc.). This mixture was then loaded to a 5.1 mL Quick-Seal polyallomer tubes (1.3 x 5.1 cm, Beckman Coulter) and sealed using a tube topper (Cordless quick-seal tube topper, Beckman). The tubes were balanced in a Stepsaver 70 V6 Vertical Titanium Rotor (Thermo Scientific) and centrifuged at 178,000 x g for 46 hours at 20 °C in a Wx Sorvall Ultracentrifuge to obtain density gradients. The density gradients were separated into 20 fractions (each 250 µL) using gravity by displacing the samples by molecular grade water pumped by a syringe pump. Each fraction was mixed and sampled to measure their refractive index to calculate the buoyant density. The fractions were cleaned using a glycogen and ethanol precipitation to remove the cesium chloride and precipitate the DNA. They were then re-suspended in 30 µL PCR grade molecular water and stored at -20 °C until further analysis. The concentration of DNA in each fraction was quantified with the Quant-iT™ dsDNA High-Sensitivity Assay Kit using the Qubit® 2.0 Fluorometer. 111 Illumina MiSeq™ High Throughput Amplicon Sequencing Both total genomic DNA extracts and selected fractions were submitted for high throughput amplicon sequencing following the protocol described elsewhere (32, 33) at the Research Technology Support Facility (RTSF) at Michigan State University. Fractions were selected based on the comparison of DNA concentrations over the buoyant density range in the fractions obtained from the unlabeled and labeled RDX amended microcosms. Those fractions from the labeled RDX amended microcosms illustrating a higher DNA concentration at the higher buoyant density values were selected for sequencing. This resulted in six fractions being sequenced for each of the four samples ultracentrifuged. Further, each fraction was sequenced with three replicates. In addition, fifteen total DNA samples were submitted for sequencing, including DNA extracted from 1) the soil prior to any incubation (called initial soil), 2) the two no RDX controls (one with glucose and the other without glucose), 3) replicate microcosms amended with labeled RDX and no glucose, 4) replicate microcosms amended with unlabeled RDX and no glucose, 5) replicate microcosms amended with labeled RDX and glucose, 6) replicate microcosms amended with unlabeled RDX and glucose, 7) replicate microcosms amended with two additions of unlabeled RDX and no glucose and 8) replicate microcosms amended with two additions of unlabeled RDX and one addition of glucose. Illumina specific fusion primers were used to amplify the V4 region of the 16S rRNA (32) and to add unique barcodes to samples in each well to enable pooling and sequencing. After the amplicons were checked on 1% agarose gel, equimolar amounts of the sample were pooled to normalize results, purified and then sequenced on the Illumina MiSeq™ Personal Sequencing System. The amplicon sequencing data in the fastq file format was analyzed on Mothur version 112 1.33.0 from Patrick D. Schloss Laboratory (34) using the MiSeq standard operating procedure (35, 36). Barcode information was removed from the sequence data and contiguous sequences were created using the forward and reverse reads, were analyzed for errors and then classified. Samples were checked for the proper read length (<275 bp), ambiguous bases and homopolymer length greater than 8 to eliminate such sequences. These sequences were then aligned with the SILVA bacteria database (37) for the V4 region. Chimeras, mitochondrial and chloroplast lineage sequences were removed and then the sequences were classified into OTU’s. The OTUs were then grouped into taxonomical levels with corresponding confidence levels. Rarefaction curves, Cho1 and Shannon values were determined for all DNA extracts using Mothur. Illumina sequencing data was deposited in the NCBI Sequence Read Archive under BioProject Number PRJNA264536. Data Analysis The most abundant phylotypes (>1% relative abundance) in all fifteen DNA extracts were determined and compared to ascertain differences and similarities between the microbial communities. The relative abundance of phylotypes was compared between the no RDX controls (initial soil and no RDX amended microcosms) to determine which phylotypes experienced an increase in abundance following exposure to RDX. In addition, the relative abundance (%) of the most abundant phylotypes in the fractions of the labeled RDX amended samples were determined. These abundance values were then compared to values from the fractions obtained from the unlabeled RDX amended samples. The purpose of this comparison was to determine which phylotypes had incorporated the 15N or 13C label from RDX and would thus be found in the heavier fractions. 113 RESULTS RDX degradation occurred in all live microcosms within 30 days, whereas no removal of RDX was noted in the abiotic controls (data not shown). In the live microcosms, additional HPLC peaks appeared and when these were compared to analytical standards, they were identified as the mono-, di- and tri-nitroso derivatives of RDX. By day 30, all additional peaks had disappeared and DNA was extracted at this time. An extra amendment of RDX was added to four microcosms and this was removed (as were the metabolites) within 20 days. Again, DNA was extracted at this time. High throughput sequencing was conducted on fifteen total genomic DNA extracts to investigate the microbial communities present following each treatment. For this, the most abundant phylotypes (>1% relative abundance) were compared between the microbial communities of each sample (Figure 5.1). The most abundant phylotypes in the initial soil microbial community was clearly different from the most abundant phylotypes in the other samples. For example, GP4 was the most abundant phylotype in the initial soil (12.8%), however in the other communities, it was present only at a low level (<2.1%). In addition, the three most common phylotypes in the other samples (unclassified Pseudomonadaceae, Acinetobacter and Gracilibacter) were present only at low levels in the initial soil community (<0.07%). As expected, the microbial communities were similar between the microcosms amended with labeled and unlabeled RDX. The addition of glucose also did not greatly change the most abundant phylotypes present. The four most abundant phylotypes in the microcosms amended with RDX in the absence of glucose were unclassified Pseudomonadaceae, Acinetobacter, Gracilibacter and unclassified Bacteroidetes. Whereas, in the presence of glucose and RDX, the four most abundant phylotypes 114 were Pseudomonadaceae, Acinetobacter, Gracilibacter and Comamonas. Further, the communities that received two amendments of RDX were similar to those that received only one amendment. Rarefaction curves (Figure 5.2) indicated that the majority of the populations were represented as these started to plateau for all of the samples. In contrast, the curve for the initial soil microbial community was still increasing, suggesting additional data is needed to represent the complete diversity of this sample. The total number of OTUs or species richness was estimated using the Chao1 estimator (Table 5.1). In general, the RDX and glucose amended microcosms appeared to have greater species richness compared to the RDX amended microcosms that did not receive glucose. The Chao1 value for the initial soil was notably higher than all of the other samples. The Shannon diversity value for the initial soil was also higher than all other Shannon values (Table 5.1). Again, the glucose amended samples illustrated higher Shannon values compared to the samples that were not amended with glucose. To determine which microorganisms were enriched following RDX degradation, the relative abundance of each in RDX amended microcosms was compared to the initial soil and the no RDX control microcosms. Phylotypes with greater relative abundance in the RDX amended microcosms compared to the no RDX controls are shown (Figure 5.3). For the glucose amended samples, five bacteria illustrated an increase over the controls (Figure 5.3a). From these, three illustrated a larger increase in abundance compared to the controls, including unclassified Pseudomonadeae, Pseudomonas and Comamonas. For the microbial communities in the absence of glucose, Acinetobacter illustrated the greatest increase compared to the initial soil and no 115 RDX controls (Figure 5.3b). In the absence of glucose, Rhodococcus (a genus containing many known RDX degrading isolates) also increased in abundance, although it was at a much lower level than Acinetobacter. Interestingly, both treatments resulted in a greater relative abundance of Comamonas and Sedimentibacter in the RDX amended samples compared to the initial soil and no RDX controls. Total DNA samples from the labeled and unlabeled RDX amended microcosms for both the glucose and no glucose samples were subject to ultracentrifugation and fractioning. The DNA concentration in each fraction was then determined (Figure 5.4). The DNA concentration was higher in the heavier fractions for the labeled RDX amended samples compared to the unlabeled RDX amended samples, indicating an enrichment of nucleic acids with 13C and/or 15N. This pattern occurred in both the glucose (Figure 5.4a) and the no glucose samples (Figure 5.4b). The heavy fractions were submitted for Illumina sequencing (three replicates for each) to determine which phylotypes were responsible for label uptake. The three most abundant phylotypes in the heavy fractions from the labeled RDX amended samples were determined for both treatments (Figure 5.5). The sequencing data from the three replicates were similar, resulting in small error bars. From this analysis, only one phylotype (unclassified Pseudomonadaceae) was similar between treatments and, in both cases, it was the most dominant phylotype in the heavy fractions. In the glucose amended heavy fractions Comamonas and Anaeromyxobacter were also present, but at a much lower level compared to unclassified Pseudomonadaceae (Figure 5.5a). In the heavy fractions from the no glucose samples, Azohydromonas and Rhodococcus were present and again this occurred at a much lower 116 level compared to the unclassified Pseudomonadaceae phylotype (Figure 5.5b). These data indicate unclassified Pseudomonadaceae were primarily responsible for label uptake in the soil microcosms, both in the absence and presence of glucose. DISCUSSION The potential for any site to biodegrade RDX is particularly difficult to predict because of the large number of RDX degrading isolates and the lack of correlation between these isolates and the species found at contaminated sites. To address this, the current study combined SIP and high throughput sequencing to provide an in depth analysis of the microorganisms linked to RDX degradation in samples from a Navy Base. The study utilized two layers of information to link RDX degradation with microorganism identity. Firstly, phylotypes illustrating an increase in relative abundance following RDX degradation, compared to controls (initial soil and samples with no RDX added) were determined. This approach is similar to using quantitative PCR to document growth on a substrate for specific phylotypes within mixed communities. It cannot be definitely stated that these microorganisms were growing using RDX as a carbon, nitrogen or energy source. However, because their abundance increased following RDX degradation, they were clearly benefiting from this process. Further, in replicate samples not exposed to RDX, this increase did not occur. Secondly, SIP was used to identify which microorganisms were involved in label uptake from RDX. Combining the results from these two approaches has provided an enhanced data set for documenting the microorganisms involved in RDX degradation. To date, a range of molecular methods have been used to examine microbial communities at RDX contaminated sites, including terminal restriction fragment length polymorphism (TRFLP), 117 lipid biomarker analysis (38), denaturing gradient gel electrophoresis (17, 19, 39), amplified 16S rDNA restriction analysis (ARDRA)(21), cloning and Sanger sequencing (17, 19-21). Only recently (2013), has high throughput sequencing (pyrosequencing, not in combination with SIP) been applied to examine RDX degrading microbial communities (40). Four studies have applied SIP to investigate RDX degradation (22-25). SIP has the advantage over these methods in that the labeled carbon or nitrogen derived from RDX can be linked to microorganism identity, providing more robust information on the microorganisms involved in RDX degradation. The current study is the first to combine SIP with high throughput sequencing to investigate the microorganisms responsible for RDX degradation. Using high throughput sequencing, rather than Sanger sequencing, with SIP provides a much greater depth of information on the microorganisms involved. To our knowledge, SIP has been used in combination with traditional Sanger sequencing in four studies to examine RDX degradation (22-25). One study using 15N and 13C labeled RDX revealed the important of Sphingobacteriales in label uptake (24). Another SIP study (15N labeled RDX) using samples from the Picatinny Arsenal Site identified five phylotypes similar to known RDX degraders and ten phylotypes not previously linked to RDX degradation (25). The authors reported that several of these phylotypes classified within the genus Pseudomonas. The authors did not report any enrichment of phylotypes from the Betaproteobacteria, indicating they did not find enrichment of the genus Comamonas. A more recent SIP study (2013), using 13C labeled RDX, also reported the importance of Pseudomonas phylotypes in label uptake (23). The 2013 study also found that Rhodoferax, a genus within the same family (Comamonadaceae) as Comamonas, was responsible for label uptake. However, the research also reported label uptake over a range of phyla, including Proteobacteria, Spirochaetes, Bacteroidetes, Firmicutes, and Actinobacteria (23). The fourth SIP study (15N 118 labeled RDX), using material from the Eglin Air Force Base bombing range, found that Rhodococcus illustrated the greatest amount of label uptake (22). From the other seven phylotypes that were associated with label uptake, one classified (Variovorax) within the same family as Comamonas and another classified within the same order (Burkholderiales). The results obtained in the current study agree with previous SIP studies, indicating the importance of phylotypes within the families Pseudomonadaceae and Comamonadaceae in label uptake from RDX. In addition, similar to previous studies, the current research also found that Rhodococcus was associated with label uptake. Taken together, these data indicate that primers designed towards these phylotypes would be advantageous for investigating the feasibility of bioremediation across RDX contaminated sites. Known RDX degrading isolates classify within four phyla, the Firmicutes, Actinobacteria, Proteobacteria and Fusobacteria. The phylum Firmicutes contains many RDX degrading bacteria and these fall within the class Clostridia or Bacilli (9, 10, 41-47). In the current study, unclassified Clostridiaceae 1 and Sedimentibacter were enriched following RDX degradation, but were not linked with label uptake from RDX. The phylum Actinobacteria also contains RDX degrading isolates including many within the genus Rhodococcus (14, 48-55). Other Actinobacteria isolates include Williamsia sp. KTR4 (56), Gordonia sp. KTR9 (56, 57) and Gordonia sp. YY1 (19). The data from the current study indicate Rhodococcus phylotypes increased in abundance following RDX degradation and the SIP data suggest these phylotypes are responsible for lower levels of label uptake. 119 RDX degrading isolates in the phylum Proteobacteria are found within the Alpha-, Gamma- and Deltaproteobacteria classes (12). In the current study, Anaeromxyobacter was the only Deltaproteobacteria implicated in label uptake and previous research has indicated this genus is capable of RDX degradation (46). In the Betaproteobacteria, two phylotypes were associated with label uptake (Comamonas was particularly enriched and Azohydromonas was enriched to a lower level). Interestingly, no pure cultures of Betaproteobacteria have been shown to be capable of RDX degradation. The most dominant phylotype responsible for label uptake was a Gammaproteobacteria, unclassified Pseudomonadaceae. Any of the eight genera within this family (Azomonas, Azorhizophilus, Azotobacter , Cellvibrio, Pseudomonas. Rhizobacter, Rugamonas, Serpens) could have been responsible for label uptake. These results collaborate previous research, as many Gammaproteobacteria isolates are able to degrade RDX (1, 4, 5, 58, 59). When the unclassified Pseudomonadaceae partial 16S rRNA gene sequence was compared to those in Genbank it was found to be 100% similar to twenty uncultured bacterium sequences, seven Pseudomonas sequences, and four Azotobacter tropicalis sequences. In the current study Pseudomonas was also enriched following RDX degradation and again Pseudomonas isolates have been shown to be able to metabolize RDX (47, 60). In summary, SIP was combined with high throughput sequencing to investigate the microorganisms involved in RDX degradation. To our knowledge, this is the first application of this combined approach to study RDX biodegradation. The study generated two important data sets. Firstly, high throughput sequencing data indicated several phylotypes increased in abundance following RDX degradation compared to the initial soil and no RDX controls. These trends suggest a growth benefit caused by the degradation of RDX as the microorganisms were 120 directly benefiting from RDX or from RDX metabolites. Secondly, using SIP, two phylotypes were found to be primarily responsible for label uptake, indicating incorporation of the label into their DNA. It is not possible to conclude if they were assimilating the label from RDX or from RDX metabolites. The two phylotypes associated with label uptake included Comamonas and an unclassified Pseudomonadaceae. The high throughput sequencing data indicated these two phylotypes also increased in abundance following RDX degradation compared to the controls. Taken together, the two data sets implicate these phylotypes in RDX degradation, as their populations both increased following RDX degradation and the label was incorporated into their DNA. Other SIP and pure culture studies have also reported the importance of phylotypes within the Pseudomonadaceae for RDX degradation. In contrast to pure culture and SIP studies, the current research provided evidence of the importance of Comamonas in RDX degradation. These data suggest the presence of these microorganisms at contaminated sites should enhance natural attenuation efforts. 121 APPENDIX 122 Table 5.1. Chao1 and Shannon and lower and upper confidence intervals (CI) values for all microbial communities Initial Soil Glucose, no RDX Glucose, lab 1 Glucose, lab 2 Glucose, unlab 1 Glucose, unlab 2 Glucose, unlab 1, time 2 Glucose, unlab 2, time 2 No glucose, no RDX No glucose, lab 1 No glucose, lab 2 No glucose, unlab 1 No glucose, unlab 2 No glucose, unlab 1, time 2 No glucose, unlab 2, time 2 Chao1 6444 3469 4019 4404 4974 3183 4261 4718 4531 3435 3502 3628 3238 4029 3725 Lower 95% CI 6331 3379 3928 4277 4839 3086 4168 4612 4427 3347 3406 3512 3155 3929 3622 Upper 95% CI 6574 3575 4125 4552 5130 3296 4368 4840 4653 3538 3614 3764 3337 4145 3845 123 Shannon 6.6 3.4 4.3 4.3 4.2 4.1 4.4 4.5 4.3 3.8 3.3 3.3 3.6 4.3 4.1 Lower 95% CI 6.5 3.4 4.2 4.3 4.2 4.0 4.4 4.5 4.3 3.7 3.3 3.3 3.6 4.3 4.1 Upper 95% CI 6.6 3.4 4.3 4.3 4.2 4.1 4.4 4.5 4.3 3.8 3.3 3.3 3.6 4.3 4.2 100 90 80 70 60 50 Relative Abundance (%) 40 30 20 10 0 Initial No No No No No No No Glucose, Glucose Glucose Glucose Glucose Glucose Glucose Soil Glucose Glucose Glucose Glucose Glucose Glucose Glucose no RDX Lab 1 Lab 2 Unlab 1 Unlab 2 Unlab 1 Unlab 2 or RDX Lab 1 Lab 2 Unlab 1 Unlab 2 Unlab 1 Unlab 2 Time 2 Time 2 Time 2 Time 2 unclassified Pseudomonadaceae Acinetobacter Gracilibacter unclassified Bacteroidetes unclassified Bacteria Pseudomonas unclassified Clostridiaceae 1 Comamonas unclassified Chitinophagaceae unclassified Clostridiales Gp4 unclassified Comamonadaceae Rhodococcus unclassified Oxalobacteraceae Azospirillum 3 genus incertae sedis Sedimentibacter Flavisolibacter unclassified Ruminococcaceae Clostridium III Pedobacter Brevundimonas Anaeromyxobacter unclassified Rhodocyclaceae Figure 5.1. Relative abundance of the most common phylotypes in RDX amended microcosms (with and without glucose) compared to their abundance in the initial soil community and no RDX microcosms. 124 Number of OTUs No glucose, lab 1 No glucose, unlab 1 No glucose, unlab 1, time 2 Glucose, no RDX No glucose, lab 2 No glucose, unlab 2 No glucose, unlab 2, time 2 Initial Soil Glucose, lab 1 Glucose, unlab 1 Glucose, unlab 1, time 2 Glucose, no RDX 6000 6000 5000 5000 4000 4000 3000 3000 2000 2000 1000 1000 Glucose, lab 2 Glucose, unlab 2 Glucose, unlab 2, time 2 Initial Soil 0 0 0 20000 40000 60000 80000 100000 120000 140000 0 20000 40000 60000 80000 100000 120000 140000 Number of sequences Figure 5.2. Rarefaction curves for the total microbial communities in the initial soil, RDX amended microcosms and the no RDX controls for both the no glucose amended (A) and glucose amended (B) microcosms. 125 A. Glucose Amended 40 30 Initial Soil Glucose Lab 2 Glucose Unlab 1 Time 2 Glucose, no RDX Glucose Unlab 1 Glucose Unlab 2 Time 2 Glucose Lab 1 Glucose Unlab 2 20 Relative Abundance (%) 10 0 unclassified Pseudomonadaceae Pseudomonas Comamonas unclassified Rhodocyclaceae Sedimentibacter B. No Glucose Amended 40 30 Initial Soil No Glucose Lab 2 No Glucose Unlab 1 Time 2 No Glucose or RDX No Glucose Unlab 1 No Glucose Unlab 2 Time 2 No Glucose Lab 1 No Glucose Unlab 2 20 10 0 Acinetobacter Comamonas Sedimentibacter unclassified Clostridiaceae 1 Rhodococcus Figure 5.3. Relative abundance of phylotypes illustrating a difference in relative abundance between the RDX amended microcosms and the no RDX controls (initial soil and no RDX microcosm) in the glucose amended microcosms (A) and in the microcosms that were not amended with glucose (B). 126 A. Glucose Amended Unlabeled 1 Labeled 1 10 Unlabeled 2 Labeled 2 8 DNA concentration (ng/µL) 6 4 2 0 1.700 1.720 1.740 1.760 1.780 B. No Glucose Amended 10 Labeled 1 Unlabeled 1 Labeled 2 Unlabeled 2 8 6 4 2 0 1.700 1.720 1.740 1.760 1.780 Buoyant Density (g/mL) Figure 5.4. DNA concentration (ng/µL) in fractions across buoyant density gradients obtained from samples amended with labeled and unlabeled RDX, with (A) or without (B) the addition of glucose. Replicate lines represent duplicate DNA measurements. The shaded area represents the fractions analyzed with Illumina sequencing. 127 A. Glucose Amended unclassified Pseudomonadaceae Abundance normalized to DNA mass in fractions Anaeromyxobacter Comamonas 60 60 60 Labeled Labeled Unlabeled Unlabeled Labeled 40 40 40 20 20 20 0 1.740 1.750 1.760 1.770 1.780 0 1.740 1.750 1.760 1.770 1.780 Unlabeled 0 1.740 1.750 1.760 1.770 1.780 Buoyant Density (g/mL) B. No Glucose Amended Azohydromonas unclassified Pseudomonadaceae 80 Labeled 60 Unlabeled Rhodococcus 80 80 Labeled Unlabeled 60 Labeled 60 40 40 40 20 20 20 0 1.745 1.750 1.755 1.760 1.765 1.770 0 1.745 1.750 1.755 1.760 1.765 1.770 Unlabeled 0 1.745 1.750 1.755 1.760 1.765 1.770 Buoyant Density (g/mL) Figure 5.5. The three most abundant phylotypes in the heavy fractions from the labeled RDX amended microcosms compared to their abundance in fractions of similar buoyant density from the unlabeled RDX amended microcosms. Data are shown from those amended with glucose (A, top three graphs) and those not amended with glucose (B, bottom three graphs). The y-axis represents relative abundance in each fraction normalized by the DNA mass in each fraction (relative abundance times the DNA mass (ng)). Error bars represent standard deviations from three replicates and may be too small to be seen. 128 REFERENCES 129 REFERENCES 1. Kitts CL, Green CE, Otley RA, Alvarez MA, Unkefer PJ. 2000. Type I nitroreductases in soil enterobacteria reduce TNT (2,4,6-trinitrotoluene) and RDX (hexahydro-1,3,5-trinitro-1,3,5-triazine). Canadian Journal of Microbiology 46:278-282. 2. Beller HR. 2002. Anaerobic biotransformation of RDX (hexahydro-1,3,5-trinitro-1,3,5triazine) by aquifer bacteria using hydrogen as the sole electron donor. Water Research 36:2533-2540. 3. Kitts CL, Cunningham DP, Unkefer PJ. 1994. Isolation of 3 hexahydro-1,3,5-trinitro1,3,5-triazine-degrading species of the family Enterobacteriaceae from nitramine explosive-contaminated soil. Applied and Environmental Microbiology 60:4608-4611. 4. Young DM, Unkefer PJ, Ogden KL. 1997. Biotransformation of hexahydro-1,3,5trinitro-1,3,5-triazine (RDX) by a prospective consortium and its most effective isolate Serratia marcescens. Biotechnology and Bioengineering 53:515-522. 5. Zhao JS, Halasz A, Paquet L, Beaulieu C, Hawari J. 2002. Biodegradation of hexahydro-1,3,5-trinitro-1,3,5-triazine and its mononitroso derivative hexahydro-1nitroso-3,5-dinitro-1,3,5-triazine by Klebsiella pneumoniae strain SCZ-1 isolated from an anaerobic sludge. Applied and Environmental Microbiology 68:5336-5341. 6. Adrian NR, Chow T. 2001. Identification of hydroxylamino-dinitroso-1,3,5-triazine as a transient intermediate formed during the anaerobic biodegradation of hexahydro-1,3,5trinitro-1,3,5-triazine. Environmental Toxicology and Chemistry 20:1874-1877. 7. Zhao JS, Spain J, Thiboutot S, Ampleman G, Greer C, Hawari J. 2004. Phylogeny of cyclic nitramine-degrading psychrophilic bacteria in marine sediment and their potential role in the natural attenuation of explosives. Fems Microbiology Ecology 49:349-357. 8. Zhao JS, Greer CW, Thiboutot S, Ampleman G, Hawari J. 2004. Biodegradation of the nitramine explosives hexahydro-1,3,5-trinitro-1,3,5-triazine and octahydro-1,3,5,7tetranitro-1,3,5,7-tetrazocine in cold marine sediment under anaerobic and oligotrophic conditions. Canadian Journal of Microbiology 50:91-96. 9. Zhao JS, Paquet L, Halasz A, Hawari J. 2003. Metabolism of hexahydro-1,3,5-trinitro1,3,5-triazine through initial reduction to hexahydro-1-nitroso-3,5-dinitro-1,3,5-triazine followed by denitration in Clostridium bifermentans HAW-1. Applied Microbiology and Biotechnology 63:187-193. 10. Adrian NR, Arnett CM. 2004. Anaerobic biodegradation of hexahydro-1,3,5-trinitro1,3,5-triazine (RDX) by Acetobacterium malicum strain HAAP-1 isolated from a methanogenic mixed culture. Current Microbiology 48:332-340. 130 11. McCormick NG, Cornell JH, Kaplan AM. 1981. Biodegradation of hexahydro-1,3,5trinitro-1,3,5-triazine. Applied and Environmental Microbiology 42:817-823. 12. Cupples AM. 2013. RDX degrading microbial communities and the prediction of microorganisms responsible for RDX bioremediation. Int Biodeter Biodegr 85:260-270. 13. Bernstein A, Ronen Z. 2012. Biodegradation of the explosives TNT, RDX and HMX. Microbial Degradation of Xenobiotics:135-176. 14. Fuller ME, Perreault N, Hawari J. 2010. Microaerophilic degradation of hexahydro1,3,5-trinitro-1,3,5-triazine (RDX) by three Rhodococcus strains. Letters in Applied Microbiology 51:313-318. 15. Halasz A, Manno D, Strand SE, Bruce NC, Hawari J. 2010. Biodegradation of RDX and MNX with Rhodococcus sp. strain DN22: new insights into the degradation pathway. Environmental Science & Technology 44:9330-9336. 16. Fuller ME, Steffan RJ. 2008. Groundwater chemistry and microbial ecology effects on explosives biodegradation. SERDP Final Report Project ER-1378. 17. Fuller ME, McClay K, Higham M, Hatzinger PB, Steffan RJ. 2010. Hexahydro-1,3,5trinitro-1,3,5-triazine (RDX) bioremediation in groundwater: are known RDX-degrading bacteria the dominant players? Bioremediation Journal 14:121-134. 18. Amann RI, Ludwig W, Schleifer KH. 1995. Phylogenetic identification and in-situ detection of individual microbial cells without cultivation. Microbiol Rev 59:143-169. 19. Ronen Z, Yanovich Y, Goldin R, Adar E. 2008. Metabolism of the explosive hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) in a contaminated vadose zone. Chemosphere 73:1492-1498. 20. Kwon MJ, O'Loughlin EJ, Antonopoulos DA, Finneran KT. 2011. Geochemical and microbiological processes contributing to the transformation of hexahydro-1,3,5-trinitro1,3,5-triazine (RDX) in contaminated aquifer material. Chemosphere 84:1223-1230. 21. Kwon MJ, Finneran KT. 2010. Electron shuttle-stimulated RDX mineralization and biological production of 4-nitro-2,4-diazabutanal (NDAB) in RDX-contaminated aquifer material. Biodegradation 21:923-937. 22. Andeer P, Stahl DA, Lillis L, Strand SE. 2013. Identification of microbial populations assimilating nitrogen from RDX in munitions contaminated military training range soils by high sensitivity stable isotope probing. Environmental Science & Technology 47:10356-10363. 23. Cho KC, Lee DG, Roh H, Fuller ME, Hatzinger PB, Chu K-H. 2013. Application of 13 C-stable isotope probing to identify RDX-degrading microorganisms in groundwater. Environmental Pollution 178:350-360. 131 24. Jayamani I, Manzella MP, Cupples AM. 2013. RDX degradation potential in soils previously unexposed to RDX and the identification of RDX-degrading species in one agricultural soil using stable isotope probing. Water Air and Soil Pollution 224. 25. Roh H, Yu CP, Fuller ME, Chu KH. 2009. Identification of hexahydro-1,3,5-trinitro1,3,5-triazine-degrading microorganisms via 15N-stable isotope probing. Environmental Science & Technology 43:2505-2511. 26. Sun WM, Cupples AM. 2012. Diversity of five anaerobic toluene-degrading microbial communities investigated using stable isotope probing. Applied and Environmental Microbiology 78:972-980. 27. Sun WM, Sun XX, Cupples AM. 2012. Anaerobic methyl tert-butyl ether-degrading microorganisms identified in wastewater treatment plant samples by stable isotope probing. Applied and Environmental Microbiology 78:2973-2980. 28. Sun WM, Xie SG, Luo CL, Cupples AM. 2010. Direct Link between Toluene Degradation in Contaminated-Site Microcosms and a Polaromonas Strain. Applied and Environmental Microbiology 76:956-959. 29. Xie SG, Sun WM, Luo CL, Cupples AM. 2010. Stable Isotope Probing Identifies Novel m-Xylene Degraders in Soil Microcosms from Contaminated and Uncontaminated Sites. Water Air and Soil Pollution 212:113-122. 30. Xie SG, Sun WM, Luo CL, Cupples AM. 2011. Novel aerobic benzene degrading microorganisms identified in three soils by stable isotope probing. Biodegradation 22:7181. 31. Michalsen MM, Weiss R, King A, Gent D, Medina VF, Istok JD. 2013. Push-pull tests for estimating RDX and TNT degradation rates in groundwater. Groundwater Monitoring & Remediation 33:61-68. 32. Caporaso JG, Lauber CL, Walters WA, Berg-Lyons D, Lozupone CA, Turnbaugh PJ, Fierer N, Knight R. 2011. Global patterns of 16S rRNA diversity at a depth of millions of sequences per sample. Proceedings of the National Academy of Sciences of the United States of America 108:4516-4522. 33. Caporaso JG, Lauber CL, Walters WA, Berg-Lyons D, Huntley J, Fierer N, Owens SM, Betley J, Fraser L, Bauer M, Gormley N, Gilbert JA, Smith G, Knight R. 2012. Ultra-high-throughput microbial community analysis on the Illumina HiSeq and MiSeq platforms. Isme Journal 6:1621-1624. 34. Schloss PD. 2009. A High-Throughput DNA Sequence Aligner for Microbial Ecology Studies. Plos One 4. 35. Schloss PD 6 August 2013, at 7.30 2013, posting date. MiSeq SOP. [Online.] 132 36. Kozich JJ, Westcott SL, Baxter NT, Highlander SK, Schloss PD. 2013. Development of a dual-index sequencing strategy and curation pipeline for analyzing amplicon sequence data on the MiSeq Illumina sequencing platform. Applied and Environmental Microbiology 79:5112-5120. 37. Pruesse E, Quast C, Knittel K, Fuchs BM, Ludwig W, Peplies J, Gloeckner FO. 2007. SILVA: a comprehensive online resource for quality checked and aligned ribosomal RNA sequence data compatible with ARB. Nucleic Acids Research 35:71887196. 38. Ringelberg D, Richmond M, Foley K, Reynolds C. 2008. Utility of lipid biomarkers in support of bioremediation efforts at army sites. Journal of Microbiological Methods 74:17-25. 39. Moshe SSB, Ronen Z, Dahan O, Weisbrod N, Groisman L, Adar E, Nativ R. 2009. Sequential biodegradation of TNT, RDX and HMX in a mixture. Environmental Pollution 157:2231-2238. 40. Livermore J, Oh Jin Y, LePuil M, Arnseth R, Mattes TE. 2013. Microbial community dynamics during acetate biostimulation of RDX-contaminated groundwater. Environmental Science & Technology 47:7672-7678. 41. Regan KM, Crawford RL. 1994. Characterization of Clostridium bifermentans and its biotransformation of 2,4,6-trinitrotoluene (TNT) and 1,3,5-triaza-1,3,5trinitrocyclohexane (RDX). Biotechnology Letters 16:1081-1086. 42. Bhushan B, Halasz A, Thiboutot S, Ampleman G, Hawari J. 2004. Chemotaxismediated biodegradation of cyclic nitramine explosives RDX, HMX, and CL-20 by Clostridium sp. EDB2. Biochemical Biophysical Research Communications 316:816821. 43. Zhang CL, Hughes JB. 2003. Biodegradation pathways of hexahydro-1,3,5-trinitro1,3,5-triazine (RDX) by Clostridium acetobutylicum cell-free extract. Chemosphere 50:665-671. 44. Zhao JS, Spain J, Hawari M. 2003. Phylogenetic and metabolic diversity of hexahydro1,3,5-trinitro-1,3,5-triazine (RDX)-transforming bacteria in strictly anaerobic mixed cultures enriched on RDX as nitrogen source. Fems Microbiology Ecology 46:189-196. 45. Sherburne LA, Shrout JD, Alvarez PJJ. 2005. Hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) degradation by Acetobacterium paludosum. Biodegradation 16:539-547. 46. Kwon MJ, Finneran KT. 2008. Hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) and octahydro-1,3,5,7-tetranitro-1,3,5,7-tetrazocine (HMX) biodegradation kinetics amongst several Fe(III)-reducing genera. Soil & Sediment Contamination 17:189-203. 133 47. Singh R, Soni P, Kumar P, Purohit S, Singh A. 2009. Biodegradation of high explosive production effluent containing RDX and HMX by denitrifying bacteria. World Journal of Microbiology & Biotechnology 25:269-275. 48. Seth-Smith HMB, Rosser SJ, Basran A, Travis ER, Dabbs ER, Nicklin S, Bruce NC. 2002. Cloning, sequencing, and characterization of the hexahydro-1,3,5-trinitro-1,3,5triazine degradation gene cluster from Rhodococcus rhodochrous. Applied and Environmental Microbiology 68:4764-4771. 49. Bhushan B, Trott S, Spain JC, Halasz A, Paquet L, Hawari M. 2003. Biotransformation of hexahydro-1,3,5-trinitro-1,3,5-triazine (RDX) by a rabbit liver cytochrome p450: insight into the mechanism of RDX biodegradation by Rhodococcus sp strain DN22. Applied and Environmental Microbiology 69:1347-1351. 50. Coleman NV, Nelson DR, Duxbury T. 1998. Aerobic biodegradation of hexahydro1,3,5-trinitro-1,3,5-triazine (RDX) as a nitrogen source by a Rhodococcus sp., strain DN22. Soil Biology & Biochemistry 30:1159-1167. 51. Coleman NV, Spain JC, Duxbury T. 2002. Evidence that RDX biodegradation by Rhodococcus strain DN22 is plasmid-borne and involves a cytochrome p-450. Journal of Applied Microbiology 93:463-472. 52. Fournier D, Halasz A, Spain J, Fiurasek P, Hawari J. 2002. Determination of key metabolites during biodegradation of hexahydro-1,3,5-trinitro-1,3,5-triazine with Rhodococcus sp strain DN22. Applied and Environmental Microbiology 68:166-172. 53. Nejidat A, Kafka L, Tekoah Y, Ronen Z. 2008. Effect of organic and inorganic nitrogenous compounds on RDX degradation and cytochrome P-450 expression in Rhodococcus strain YH1. Biodegradation 19:313-320. 54. Bernstein A, Adar E, Nejidat A, Ronen Z. 2011. Isolation and characterization of RDX-degrading Rhodococcus species from a contaminated aquifer. Biodegradation 22:997-1005. 55. Seth-Smith HMB, Edwards J, Rosser SJ, Rathbone DA, Bruce NC. 2008. The explosive-degrading cytochrome P450 system is highly conserved among strains of Rhodococcus spp. Applied and Environmental Microbiology 74:4550-4552. 56. Thompson KT, Crocker FH, Fredrickson HL. 2005. Mineralization of the cyclic nitramine explosive hexahydro-1,3,5-trinitro-1,3,5-triazine by Gordonia and Williamsia spp. Applied and Environmental Microbiology 71:8265-8272. 57. Indest KJ, Jung CM, Chen HP, Hancock D, Florizone C, Eltis LD, Crocker FH. 2010. Functional characterization of pGKT2, a 182-kilobase plasmid containing the xplAB genes, which are involved in the degradation of hexahydro-1,3,5-trinitro-1,3,5triazine by Gordonia sp. strain KTR9. Applied and Environmental Microbiology 76:6329-6337. 134 58. Kitts CL, Cunningham DP, Unkefer PJ. 1994. Isolation of three hexahydro-1,3,5trinitro-1,3,5-triazine-degrading species of the family Enterobacteriaceae from nitramine explosive-contaminated soil. Applied and Environmental Microbiology 60:4608-4611. 59. Pudge IB, Daugulis AJ, Dubois C. 2003. The use of Enterobacter cloacae ATCC 43560 in the development of a two-phase partitioning bioreactor for the destruction of hexahydro-1,3,5-trinitro-1,3,5-s-triazine (RDX). Journal of Biotechnology 100:65-75. 60. Cho YS, Lee BU, Oh KH. 2008. Simultaneous degradation of nitroaromatic compounds TNT, RDX, atrazine, and simazine by Pseudomonas putida HK-6 in bench-scale bioreactors. Journal of Chemical Technology and Biotechnology 83:1211-1217. 135