,5... ‘ ‘ éfi . .331? a.» s a. «war, i; z! a f . g! '1 ‘12... .5335 if xiii: x.“ , x5. Vlalel‘nt 33.3??? 5-: . liar... .1 3.. .223... 1...: i... .1 i I??? . , . 2 ... flu (v.4. .111: . , , .2 a LA ‘ am... aw... v 1,.\ l...Y ,. . . 1 . UBRARY F “ L C+A+A IVIIUI “UCIII Ulalc ___9_“5V9r§ilyl -_ — .- ‘oofi-vlv- - u-uv- ”w—‘ This is to certify that the dissertation entitled COMPARISONS OF METHANOTROPH COMMUNITIES IN SOILS THAT CONSUME ATMOSPHERIC METHANE presented by Uri Yitzhak Levine has been accepted towards fulfillment of the requirements for the Doctoral degree in Microbiology and Molecular Genetics \ /" / ‘ V w ' L%}Maj3’r Professor’s Signature 7 — 36 - 09 Date MSU is an Alfinnative AcfiorVEquaI Opportunity Employer PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5/08 KLIProj/AccaPreleIRC/DaleDue.indd COMPARISONS OF METHANOTROPH COMMUNITIES IN SOILS THAT CONSUME ATMOSPHERIC METHANE By Uri Yitzhak Levine A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Microbiology and Molecular Genetics 2009 ABSTRACT COMPARISONS OF METHANOTROPH COMMUNITIES IN SOILS THAT CONSUME ATMOSPHERIC METHANE By Uri Yitzhak Levine Methane is a potent greenhouse gas that is 21-25 times more efficient at trapping heat (infrared radiation) than carbon dioxide. Methane oxidation is mediated by methane- consuming microbes (methanotrophs), but only in upland soils does the activity of aerobic methanotrophs account for a net uptake of atmospheric methane. The conversion of native lands to row-crop agriculture diminishes the strength of the soil methane sink, typically dropping the rate of methane consumption by 70%. To determine the relationship between the rates of methane consumption in soils and the diversity of microbes that catalyze them, we conducted molecular surveys of methanotroph communities across a range of land uses at The Kellogg Biological Station Long Term Ecological Research Site (KBS LTER) and correlated our findings to measurements of the in situ fluxes of methane. Rates of methane consumption and methanotroph diversity were positively correlated, as conversion of lands to row-crop agriculture led to a 7-fold reduction fiom maximal rates of consumption in the native deciduous forests. In fields abandoned from agriculture both methanotroph richness and the consumption of methane were estimated to require approximately 75 years to return to the present diversity and consumption rate of the native deciduous forests. The linear trajectory for recovery of both measures suggested that managing lands to conserve or restore methanotroph diversity would yield increases in the rate of consumption of atmospheric methane in KBS LTER soils. Long-term fertilization is one aspect of row-crop agriculture that is likely to be a significant disturbance to the methanotroph community. We hypothesized that a consequence of this disturbance would be its association with decreases in methane consumption and methanotroph richness in fertilized forest sub-plots at KBS LTER, but neither rate nor methanotroph richness declined due to long-term fertilization alone at KBS LTER. A meta-analysis examined the effect of long-term fertilization in other sites, and revealed no consistent decline in methanotroph richness in fertilized soils. The methanotroph communities did display a distinct biogeography with communities clustering together based on geographic location. As a consequence, the composition of the unique soil methanotroph community probably plays a role in dictating the response of the methanotroph community to changing land use and its disturbances. The causes behind the change in methanotroph richness and the correlated decrease in methane consumption associated with row-crop agriculture at KBS LTER remains unclear as it is not caused by fertilization alone. The quantification of the effect of other variables associated with agricultural management is necessary to determine which management strategies at KBS LTER could be utilized to enhance methanotroph richness. However, a management strategy determined at KBS LTER to be beneficial to the methanotroph community may not be applicable to other paired sites, as the unique methanotroph community and environmental characteristics from each geographic location will probably yield a dissimilar response to the management practice. ACKNOWLEDGEMENTS There are many people who have helped me on my journey through graduate school, and I am grateful to them all. My colleagues in the Schmidt Lab were enormously generous with their time, advice and help. Without their lessons, advice and training I would not be where I am today. Stephanie Eichorst took me under her wing as a rotation student and helped teach me many of the bench skills I would repeatedly utilize during my research. Zarraz Lee was the longest fellow traveler and her help, ideas and encouragement were invaluable. In my first few years of graduate school, John Breznak, John Wertz, Kristin Huizinga, Brad Stevenson, Jorge Rodrigues, Dion Antonoplulous and Kwi Kim were all tremendous sources of advice and help, and later on so too were Clegg Waldron, Tracy Teal, Vicente Gomez, Kevin Theis and Russell Grant. Just as important, all members of the Schmidt lab established a comfortable and collegial working environment that fostered my development, was focused on helping and supporting each other, and is one that I hope to help recreate wherever I go so that I and others can continue to reap its benefits. That atmosphere was a product of my advisor, Tom Schmidt, who set the tone through his own unfailing willingness to help. In addition to setting an excellent example for me to follow, through his tutelage I have become a much better scientist who is more meticulous, thoughtful, purposeful, and independent. The lessons and thinking that he has instilled in me I will take with me throughout my career, and I am extremely grateful. In addition, graduate school has been a period of tremendous growth for me in my personal life, and I am thankful to have had an advisor who was understanding and patient as I adjusted to becoming a husband and father. iv My committee, Phil Robertson, Rob Britton, Terry Marsh, and Ned Walker were invaluable resources, each contributing expertise to different areas of my research, expertly guiding my research, and they never failed to help or to share their own lab’s resources in helping my research. The Britton lab graciously shared much equipment, and without the help of Andrew Corbin, Stacey VanderWulp, Neville Millar and the rest of the Robertson lab, my sampling and experiments at the Kellogg Biological Station Long Term Ecological Research Site (KBS LTER) would have been impossible. The Robertson lab was also responsible for the taking of all of trace gas fluxes used in this study, and without their collaboration my data could not have been placed in its proper context, and its potential significance would remain unclear. I am grateful for their hard work, willingness to help, and the ease of our collaboration. When I traveled to Rothamsted Research I was graciome hosted by the Hirsch lab, and Penny Hirsch, Ian Clark, Keith Goulding and Paul Poulton all devoted considerable time and effort to ensuring that my visit and research was productive. All of this research would not have been possible without funding from an Environmental Protection Agency Star Fellowship, College of Natural Science Recruiting Fellowship and Dissertation Completion Fellowship, a Kellogg Biological Station Long Term Ecological Research small graduate student grant, international travel support from the National Science Foundation’s Long Term Ecological Research program, and support from the Michigan Agricultural Experimental Station. Last, but certainly not least, I would not be here without the support of my friends and especially my family. My wife Rachel and daughter Channa have been my enthusiastic cheerleaders, enduringly patient, and willing to support me in all of my endeavors. I am and will always be, eternally gratefirl for them and their support. vi TABLE OF CONTENTS LIST OF TABLES ................................................................................... ix LIST OF FIGURES ............................................................................... CHAPTER 1. Uncertainties in the Global Methane Budget and Biological Influences on the Concentration of Atmospheric Methane .............................. ....X ...1 Introduction .................................................................................... 1 Summary ...................................................................................... l4 Thesis Overview ................................................................................ 14 References .................................................................................. CHAPTER 2. Agriculture’s Impact on Microbial Diversity and the Flux of ..20 Greenhouse Gases ................................................................................... 26 Abstract ...................................................................................... Introduction ................................................................................. Materials and Methods ................................................................... Results Drscussmn Conclusion............ References .................................................................................. CHAPTER 3. The Impact of Long-term Fertilization to the Methanotroph ..26 ..27 ..30 .37 .42 Acknowledgments.................................. .48 ..63 Communities in Soils ................................................................................ 68 Abstract ...................................................................................... Introduction ................................................................................. Materials and Methods ................................................................... Results Discussion............. Conclusron Acknowledgments............................. References .................................................................................. CHAPTER 4. Conclusions and Future Directions..................... ..68 ..69 ..72 .77 .79 .85 .86 ..97 .100 References ................................................................................... 105 APPENDIX A. Assessment of the PCR bias .................................................. 107 Introduction ................................................................................. 107 Methods ..................................................................................... 108 Results Discussronn .110 111 References ................................................................................... 1 l6 vii APPENDIX B. Ammonia and Nitrate Before and After Fertilization in a KBS LTER Fertilized Sub-plot ......................................................................... 117 Introduction ................................................................................. 1 17 Methods ..................................................................................... 117 Results and D1scussron118 Acknowledgmentsll9 References ................................................................................... 121 APPENDIX C. Biogeography of Methanotrophs is Well Drained Soils... .. ............122 Introduction ................................................................................. 122 Methods ..................................................................................... 122 Results and Discussron123 References ................................................................................... 127 viii LIST OF TABLES Table 2.1. KBS-LTER Sites Investigated in this study ......................................... 49 Table 2.2. Summary of pmoA clone libraries ................................................... 50 Table 2.3. Correlations between the flux of greenhouse gases, species richness and environmental conditions ........................................................................... 51 Table 3.1. Summary of the Sites and pmoA libraries used in this study. .................... 87 Table A. 1. Composition, expected output and actual output of defined artifical communites of pmoA and amoA species used to assess the PCR bias of the primer pair A189-A682. ........................................................................................ 114 Table C.l. Summary of the Sites and pmoA libraries used in this study. ................. 125 ix LIST OF FIGURES Figure 1.1. A simplified schematic of the metabolic pathway of methane oxidation. The key enzyme is the methane monooxygenase, which facilitates the oxidation of methane into methanol .......................................................................................... 18 Figure 1.2. Phylogenetic tree of selected partial pmoA and amoA protein sequences from public databases. The tree is based on 164 amino acid positions using Phylip Protein Maximum Likelihood as implemented in ARB (Ludwig et a1. 2004). Boxed and shaded labels are indicative of pmoA clades that have been found in soils, have not been cultured, and are thought to likely play a substantial role in the oxidation of atmospheric methane. The scale bar represents 10 PAM units ............................................... 19 Figure 2.1. Average monthly carbon dioxide production and methane consumption based on current land management and historical land use at the KBS-LTER: Agricultural management of historically tilled land (Ag HT; V ), early successional plant communities on fields that had been abandoned from agriculture in 1989 (Early HT; I ), mid- successional plant communities on either historically tilled land (Mid HT; J- ) or never tilled land (Mid NT; 0 ), or a late successional deciduous forest (Late DF; 0 ) ............ 52 Figure 2.2. The effect of different land uses on average carbon dioxide emission (a) and net methane consumption (b) at KBS LTER. Different letters represent significant differences (p<0.05) between treatments. Rate measures are the same as those in Fig. 2.1. Error bars represent standard errors. Land use treatments are: agricultural management of historically tilled land (Ag HT), early successional plant communities on fields that had been abandoned from agriculture in 1989 (Early HT), mid-successional plant communities on either historically tilled land (Mid HT) or never tilled land (Mid NT), or a late successional deciduous forest (Late DF) ................................................. 53 Figure 2.3. pmoA rarefaction curves from all of the KBS-LTER pmoA clone libraries constructed from the Late DF and Ag HT treatments. Libraries were constructed from DNA extracted from soil sampled in December 2004, June 2005, and June 2006 with various annealing temperatures (For additional details see Table 2.2, and Methods). All curves were constructed using data from neighbor joining matrixes from Arb (Ludwig et a1. 2004), with curves calculated by DOTUR (3 8). Methanotroph species are defined as pmoA sequences having 94% average nucleotide sequence similarity. Error bars representing 95% confidence intervals were omitted for the sake of c1arity...... . . . . . . . . . . .54 Figure 2.4. Phylogenetic tree of selected partial pmoA and amoA protein sequences from public databases and translated from PCR-based clone libraries fiom KBS-LTER soils. The tree is based on 164 amino acid positions using Phylip Protein Maximum Likelihood as implemented in ARB (Ludwig et al. 2004). Boxed or circled labels are indicative of pmoA clades recovered fi'om KBS-LTER soils. Cluster I and Cluster II clades (boxed and starred labels) were recovered from Ag HT and Late DF, while clades KBSl, JR1 , Upland Soil Cluster or, and MRI (boxed labels) were recovered from Late DF. Clade RA21 (oval) was only recovered from Mid NT soil. The scale bar represents 10 PAM unit ..................................................................................................... 55 Figure 2.5. Neighbor joining phylogenetic tree of the partial nucleic acid sequences of pmoA and amoA fi'om reference sequences, and KBS-LTER clone libraries from soil sampled in December 2004, June 2005, June 2006, and June 2007. The tree is rooted with the branching from the amoA sequences from Nitrosomonas eutropha and Nitrosococcus mobilis. Nodes representing the 27 different methanotroph species (defined as pmoA sequences having 94% average nucleotide sequence similarity) identified at KBS-LTER are highlighted. The colors of the highlights reflect species’ membership in the seven pmoA clades depicted in Figure 5. The pmoA sequences included in this tree are deposited in GenBank under accession numbers F] 529724 - FJ529808, and GQ219582-GQ219583 ............................................................ 56 Figure 2.6. The correlation between methanotrOph diversity and methane consumption at KBS LTER. Relationship between summer methane consumption (June-August) and methanotroph richness (both represent averages of 3 replicate plots) across landscapes at the KBS LTER. A simple linear regression is presented (r2 =0.62, p <0.001) with operational taxonomic units (OTUs) defined as peaks in the tRF LP analysis that have been identified as a pmoA gene. Symbols are as follows: Agricultural management of historically tilled land (Ag HT; ‘7 ), early successional plant communities on fields that had been abandoned from agriculture in 1989 (Early HT; I ), mid-successional plant communities on either historically tilled land (Mid HT; 2 ) or never tilled land (Mid NT; 0 ), or a late successional deciduous forest (Late DF; 0 ) ..................................... 59 Figure 2.7. The recovery of methanotroph diversity and methane consumption at KBS LTER following row-crop agriculture. Increase in methanotmph diversity (open symbols) and methane consumption (closed symbols) as a fimction of time since cesssation of agriculture. Measurements of the deciduous forest (Late DF) are positioned based on projections fiom linear regression used to fit methanotroph diversity (y =0.07x +2.05; r2=0.99, p = 0.020) or methane consumption (y=0.13x + 0.80; 1’2 =0.69, p < 0.001). Error bars represent standard errors; symbols are as follows: Agricultural management of historically tilled land (Ag HT; V ), early successional plant communities on fields that had been abandoned from agriculture in 1989 (Early HT; I ), mid-successional plant communities on either historically tilled land (Mid HT; :3. ) or never tilled land (Mid NT; 0 ), or a late successional deciduous forest (Late DF; 0 ) ..................................... 60 Figure 2.8. Similarity of methanotroph communities among KBS-LTER treatments. The Serenson index was calculated for each pairwise comparison of methanotrophs using OTUs for all confirmed pmoAs, and then plotted using two-dimensional non-metric multidimensional scaling (Hammer et al. 2001). Symbols are as follows: Agricultural management of historically tilled land (Ag HT; V ), early successional plant communities on fields that had been abandoned from agriculture in 1989 (Early HT; I ), mid- successional plant communities on either historically tilled land (Mid HT; .3. ) or never tilled land (Mid NT; 0 ), or a late successional deciduous forest (Late DF; 0 ). Figure 2.9 contains the same data represented as a dendogram) ............................................. 61 xi Figure 2.9. Similarity of methanotroph communities among KBS-LTER treatments. The Serenson index was calculated for each pairwise comparison of methanotrophs using OTUs for all confirmed pmoAs, and then clustered using neighbor-j oining with MEGA (54). Symbols are as follows: Agricultural management of historically tilled land (Ag HT; V ), early successional plant communities on fields that had been abandoned fiom agriculture in 1989 (Early HT; I ), mid-successional plant communities on either historically tilled land (Mid HT; .5». ) or never tilled land (Mid NT; 0 ), or a late successional deciduous forest (Late DF; 0 ) ....................................................... 61 Figure 3.1. Differences between Late DF fertilized and control treatments at KBS LTER in average net methane consumption and the average number of methanotroph species (defined as pmoA sequences having 94% average nucleotide sequence similarity). Fertilization has no effect on either measure (t=0.58, p=0.57 for consumption; and t=1.18, p=0.36 for richness). Measures for net methane consumption are averages from 3 replicates of each treatment, while the methanotroph species are averages fi'om 2 replicates of each treatment. Net methane consumption from the same 2 replicates as those with methanotroph species data yields the same result (20.9:l:4.3 and 19.2:l:3.3 average net methane consumption of the control and fertilized sub-plots, respectively, and t=0.28 p=0.79). Error bars represent standard error ........................................ 88 Figure 3.2. (A) Rarefaction curves from pmoA clone libraries constructed from the control and fertilized sub-plots of the late successional forest at KBS LTER. Libraries were constructed from DNA extracted from soil sampled in June 2007 . (B) Rarefaction curves fi'om pmoA clone libraries constructed fi'om Rothamsted Reasearch soils sampled in Feburary 2008. All curves were constructed using data from neighbor joining matrixes from Arb (Ludwig et al. 2004), with curves calculated by DOTUR (Schloss and Handelsman 2005). Methanotroph species are defined as pmoA sequences having 94% average nucleotide sequence similarity. Error bars representing 95% confidence intervals were omitted for the sake of clarity ...................................... ' .......................... 89 Figure 3.3. Phylogenetic tree of selected partial pmoA and amoA protein sequences fi'om public databases and translated from PCR-based clone libraries from late successional deciduous forest sub-plots. The tree is based on 164 amino acid positions using Phylip Protein Maximum Likelihood as implemented in ARB (26). Bolded clades were recovered from the late successional forest sub-plots. The symbols adjacent to the clade reflect the clade’s recovery from either the control treatment (0) or the fertilized treatment (I). The scale bar represents 10 PAM units. ...................................... 90 Figure 3.4. (A) Similarity of methanotroph communities among fertilized and control sub-plot replicates. (B) Similarity of methanotroph communities between fertilized and control sub-plot replicates with all other pmoA libraries from KBS LTER. Expect for the fertilized and control sub-plots, the pmoA clone libraries were first reported in Chapter 2, and details of their construction can be found there. Both dendograms are based on Serenson index calculations for each pairwise comparison of the methanotroph communities using pmoA species (defined as pmoA sequences having 94% average nucleotide sequence similarity), and then clustered using neighbor-joining with MEGA (35). The scale bars represent at 0.05 change in the Sarenson index. ........................ 91 xii Figure 3.5. Neighbor joining phylogenetic tree of the partial nucleic acid sequences of pmoA and amoA fi'om reference sequences, and KBS-LTER clone libraries from the control and fertilized sub—plots soil sampled in June 2007. The tree is rooted with the branching from the amoA sequences from Nitrosomonas eutropha, Nitrosococcus mobilis, and Nitrosovibrio tenuis. Nodes representing the 17 different methanotroph species (defined as pmoA sequences having 94% average nucleotide sequence similarity) found in late successional deciduous forest sub-plots are circled. The symbols in the circles reflect species’ recovery from either the control treatment (O ) or the fertilized treatment (I). ........................................................................................ 92 Figure 3.6. Methanotroph richness in paired sites featuring long-term fertilized (black shading) and non-fertilized soils (grey shading). The three sites to the left compare fertilized sites in the context of agricultural management, while the two sites on the right are fertilized forests whose only land management is fertilization. Error bars represent standard errors as the averages are reported for sites that featured replicated sites. Additional site details are provided in Table 3.1 .................................................. 94 Figure 3.7. Phylogenetic tree of selected partial pmoA and amoA protein sequences from public databases and translated from PCR-based clone libraries from late successional deciduous forest sub-plots. The tree is based on 164 amino acid positions using Phylip Protein Maximum Likelihood as implemented in ARB (26). Bolded clades were recovered from Rothamsted Research soils. Beneath each bolded clade is a listing of the number of species within that clade found in which Rothamsted Research treatment: Knott Wood (Wood), Broadbalk Wilderness (Wilderness) and Broadbalk Wheat (Wheat). The scale bar represents 10 PAM units. ............................................................ 95 Figure 3.8. Similarity of methanotroph communities in paired sites featuring long-term fertilized and non-fertilized soils. (A) The Sarenson index was calculated for each pairwise comparison of methanotroph species using two-dimensional non-metric multidimensional scaling (21). (B) The same data as in (A), but clustered using neighbor- joining with MEGA (35) and displayed as a dendogram. Labels with a 7:? are sites that have been fertilized long-term. Additional site information is provided in Table 3. 1 . .96 Figure A.1. Expected and actual relative abundance of the artificial pmoA and amoA community Mix 2. ................................................................................. 115 Figure B]. Ammonia and nitrate concentrations in fertilized and control Late DF sub- plots at KBS LTER. Error bars represent the standard error. ............................... 120 Figure C.1. Similarity of methanotroph communities in soils from around the globe. The dendogram is based on Serenson index calculations for each pairwise comparison of the methanotroph communities using pmoA species (defined as pmoA sequences having 94% average nucleotide sequence similarity), and then clustered using neighbor-j oining with MEGA (48). The scale bars represent at 0.1 change in the Serenson index. Additional site information is provided in Table C.l ............................................................ 126 xiii Chapter 1 Uncertainties in the Global Methane Budget and Biological Influences on the Concentration of Atmospheric Methane Introduction To better understand the soil methane sink and its capacity for atmospheric methane consumption, the studies in this thesis attempt to link variation in methane flux associated with different land uses to changes in the richness and composition of communities of methane consuming bacteria (methanotrophs). We then investigate the effect of long-term fertilization, one of the factors associated with the land use changes, to determine if it is associated with methanotroph community changes. As an introduction to these studies, this chapter reviews the role of the soil methane sink in the global methane budget, the methane consuming bacteria that are responsible for the methane sink, the microbial pathway and enzyme that facilitates methane oxidation, the effects of changing land use on the methane sink, and an overview of the following chapters. The Concentration of Atmospheric Methane Methane (CH4) is a potent greenhouse gas whose atmospheric concentration, as of 2005, is 1774 ppb (1). Methane contributes approximately 15% to the atrnosphere’s total radiative forcing (1), and is 21-25 times more efficient at trapping heat (infrared radiation) than carbon dioxide due to a longer lifetime in the atmosphere and greater radiative efficiency (2). The present concentration of atmospheric methane represents its highest concentration in at least the past 650,000 years (1). Since pre-industrial times (ca. 1900), the concentration has increased 250% largely due to human agricultural practices and fossil fuel use (1). Following almost a decade with little change in atmospheric methane concentrations, as of 2007, there are renewed increases measured at all worldwide monitoring stations (3). The balance between methane emissions and the strength of the largest methane sink, photochemical oxidation by hydroxyl radicals in the troposphere, primarily determines the concentration of atmospheric methane (4). Therefore, it is thought that the current increase in methane is either due to increases in wetland emissions in Siberia, or a weakening of the hydroxyl radical sink (3). However, ambiguity in our knowledge of multiple facets of the global methane budget has resulted in our inability to definitively identify the cause of either the current increase or the decade of little change in the concentration of atmospheric methane (3). Many other methane sources apart from wetlands, the largest source of emissions, could be changing the magnitude of their flux, and the strength of other methane sinks could also be changing. Sources of Atmospheric Methane There are numerous environments that are methane sources (yield a net production of methane to the atmosphere). Many anaerobic environments produce methane due to the activity of methanogenic archaea, and their activity accounts for more than 70% of methane emissions. These environments include wetlands, landfills, oceans, domestic and wild ruminant animals, poorly drained soil, and termites (5). Other sources of methane include methane hydrates, wildfires, fossil fuel mining and use, and biomass burning (5). Despite the identification of so many methane sources, the existence and magnitude of emissions from additional methane sources is currently debated. For example, plants producing methane under aerobic conditions was reported by Keppler et al. (6), and originally estimated to be contributing between ~10-40% of annual methane emissions. Critiques of their methods and data fi'om ice cores revised their estimates downwards to 0-10% of total annual emissions (7, 8). Attempts to replicate the initial observation have been inconsistent, with some studies confirming the initial observation of methane emission from plants (9-12), while others have been unable to corroborate those findings (13-15). Some studies have reported that the observed methane is created through UV degradation of plant pectin (9, 11, 12), and if shown to be the causal mechanism it would also explain the inability of some studies to replicate the initial findings as UV light was not included in experimental conditions (13, 14). However, the level of UV light exposure that causes methane emissions due to pectin degradation far exceeded the amount of natural UV light exposure (11), and normal UV light conditions would not yield large methane emissions from plants (13). Therefore, the contribution of plants to methane emissions remains controversial. Another possible contributor to methane emissions that is not typically accounted for in current methane budgets is methane emissions originating from geological sources. The methane is produced from microbial and therrnogenic processes in Earth’s crust and subsequently released into the atmosphere through faults, seepage, and other mechanisms. Current measurements have considerable error associated with them, but they are estimated to possibly contribute 6% of total annual emissions (5). Thus, two potentially significant sources of methane emissions remain controversial, and the magnitude of their flux and their contributions to levels of atmospheric methane remain unconstrained. Methane Sinks The number of methane sinks (environments that yield a net uptake of methane from the atmosphere) is far fewer than the number of methane sources, but there is similar uncertainty in the quantification of their impact on the concentration of atmospheric methane. Approximately 90% of the methane consumed by methane sinks is due to the photochemical oxidation by hydroxyl radicals in the troposphere. The other methane sinks, stratospheric loss and uptake by well-drained soils, split the remaining 10% of methane sink consumption (5). The hydroxyl radicals in the troposphere are very short lived, and producing accurate measurements of the strength of the sink and its variability remains challenging despite recent methodological improvements (4). Consequently, accurate modeling of the troposphere sink and the ability to attribute changes in the concentration of atmospheric methane as an effect of changes to the troposphere sink remains difficult (3). In addition, factors causing variations to the hydroxyl radical sink are unknown. Due to the strength of the troposphere sink, modifications to it will have large implications for the concentration of atmospheric methane. Like other methane sources and sinks, future investigations will have to clarify the role of the troposphere sink in detemrining the concentration of atmospheric methane. Compared to the troposphere sink, the atmospheric methane consumed by well- drained soils is small: 30 z 15 Tg of methane, roughly 3-9% of the total amount of methane removed from the atmosphere (16). However, well-drained soils are the only biological sink of methane (5). The uptake of methane in well-drained soils is due to the direct consumption by aerobic methane-oxidizing bacteria (methanotrophs) in soils (Reviewed by (17)). Methanotrophs are ubiquitous, and will consume a significant portion of the methane produced in most of the environments that are methane sources (i.e. (18)), but only in well-drained soils does their activity yield an environment with a net uptake of methane (see comment above). In a reaction mediated by their methane monooxygenase enzyme (MMO), methanotrophs consume methane via an initial oxidation with oxygen to produce methanol with the following stoichiometry: CH4 + 02 + NADH + H+ —> CH3-OH + H20 + NAD+ After several chemical conversions (Figure 1.1), the carbon from methane is either incorporated into biomass or respired as C02. Aerobic Methanotrophs Aerobic methanotrophs are phylogenetically distinct bacteria, and are found within the Proteobacteria and Verrucomicrobia phyla. Verrucomicrobia methanotrophs have only recently been described, having been isolated from extremely acidic aquatic environments (19-21). Their optimum methanotrophic activities are typically in the range of pH 2.0-2.5 (21), are capable of oxidation down to pH 0.8-1 (20), and do not contain the intracellular membranes typical of Proteobacteria methanotrophs (19). ' Although one strain has had its genome sequenced (22), our knowledge of these methanotrophs remains limited. Thus far, Verrucomicrobia methanotrophs have only been discovered from aquatic environments, and Verrucomicrobia methanotrophs have not been found in soils. Isolates of Proteobacteria methanotrophs have been well studied in culture, but the importance of cultured isolates to the consumption of atmospheric methane and the soil methane sink is limited (see discussion below). Proteobacteria methanotrophs are found in the gamma-Proteobacteria and alpha-Proteobacteria subdivisions. Type I methanotr0phs are gamma-Proteobacteria, and include the genera Methylomonas, Methylobacter, Methylococcus, and Methylomicrobium. Type II methanotrophs are alpha-Proteobacteria, and include the genera Methylosinus and Methylocystis. In addition to their subgroup classifications, Type I and Type II Proteobacterz'a methanotrophs can be distinguished on the basis of the following traits: dominant phospholipid fatty acids, serine verses RuMP carbon assimilation pathways, cellular morphology, ability to fix nitrogen, G+C content of DNA, and the ability to form resting stages (reviewed in (23)). Notably, some authors define Methylococcus capsulatus (Bath) and similar strains as Type X methanotrophs due to the presence of both serine and RuMP carbon assimilation pathways, and possession of traits similar to Type II methanotrophs. Other Methane Oxidizing Microbial Communities Anaerobic methane oxidation has also been observed, and like aerobic methane- oxidation it can attenuate a significant portion of the methane emitted from methane sources (24). Details of metabolism are still being unraveled, but the process is attributed to methanotrophic archaea (ANME) who are most likely performing reverse methanogenesis within a microbial consortia (reviewed by (24)). ANME have typically been found in close association with sulfate-reducing bacteria (25), but other bacteria also appear to be involved in the microbial consortia (26). Denitrifying bacteria anaerobically oxidizing methane without ANME has also been reported (27), further highlighting the possibility that other electron acceptors could be involved (24). ANME have thus far only been found in environments that are net methane sources and due to their requirement for anoxia, it is unlikely that they contribute to the consumption atmospheric methane in well-drained soils. Ammmonia-oxidizing bacteria have also been thought to potentially contribute to the strength of the soil methane sink as MMO is evolutionarily related to ammonium monooxygenase (AMO) (28), and both enzymes are capable of oxidizing a wide range of substrates that includes the other enzyme’s substrate (reviewed in (23)). However, while AMO is capable of oxidizing methane, its Km for methane is much higher than MMO’s Km, such that about a thousand ammonia-oxidizers are required to achieve the same rate of CH4 oxidation as a single methanotroph (29). No evidence has been found for ammonia oxidizers playing a significant role in methane oxidation under low or high meflrane levels in either microcosms or the environment (29-31). Therefore, while it is possible for ammonia-oxidizers oxidize to methane, their in situ contribution to atmospheric methane consumption is, at best, minimal. As a result, we can assume that the strength of the soil methane sink is mostly, if not entirely, due to the activity of aerobic methanotrophs. Methane Monooxygenase Methane monooxygenase (MMO) is the key enzyme in the oxidation of methane, and is found in both soluble and particulate forms. In the methanotrophs that produce both forms, soluble MMO (sMMO) is expressed under low copper conditions, but there is sufficient copper available in nearly all environments that it is rarely expressed in the environment. Conversely, the particulate MMO (pMMO) is found in all known methanotrophs (32) except two Methylocella strains (33). Due to the near ubiquity of the particulate MMO in methanotrophs, the gene encoding the A subunit of pMMO, pmoA, is typically used for non-culture based assessments of the methanotmph community in the environment (17). Methanotrophs can be also be identified through their pmoA gene due to its phylogeny being congruous with their 16S ribosomal genes (17). 168 ribosomal genes can also be used for the identification of environmental methanotrophs, but many of the primer pairs will capture methylotrophs in addition to methanotrophs (reviewed by (17)). In addition, use of pmoA is advantageous for methanotroph detection instead of ribosomal genes because unlike 16S sequences, the inference of methanotrophy is not constrained by our phylogenetic knowledge of cultured methanotrophs. For instance, if a novel pmoA is found it can be assumed to be a novel methanotroph, but if a novel 16S gene is discovered, the assumption of methanotrophy is uncertain. pMMO is made up of three subunits encoded by the genes, pmoC, pmoA and . . . 70 pmoB, Wthh are found consecutively m an operon under the control of a a promoter (34). There are typically two nearly identical copies of the pmoCAB operon in the genome (34, 35). pMMO is an integral membrane protein, and as a result it has proven difficult to study as purified protein preparations are unstable with low specific activities (32, 36). Therefore, our knowledge of pMMO’s biochemistry remains limited and controversial. The pMMO active site remains unknown despite solved protein structures (37, 38), and there are numerous hypotheses as per the location and nature of the active site. The active site could potentially lie in any of the three subunits that make up pMMO (subunits A, B and C), and data supports the active site metal center as either, or some combination of: diiron (36), mononuclear (3 7), dinuclear (3 7, 38) or trinuclear copper metal centers (39, 40). Further hampering our understanding of pMMO is the continued inability to correlate the reaction of substrates or products with any of the hypothesized active sites (reviewed by (41)). Methanotrophs Responsible for Consumption of Atmospheric Methane Regardless of its biochemistry or the role it plays in catalyzing the oxidation of methane, the pmoA gene remains an effective method of characterizing and identifying the methanotroph community. Surveys of pmoA in well-drained soils have rarely found the well-studied cultivable Proteobacteria methanotrophs detailed above (reviewed in (17)). These observations were not completely unexpected, as the inability of cultured methanotrophs to grow on atmospheric methane had led to the hypothesis that uncultured “high affinity” methanotrophs are responsible for the strength of the soil methane sink (42). Most of the cultured methanotrophs have a half-saturation constant (Km) that is too high to allow grth on atmospheric methane. Only three Methylocystis strains (sp. DWT, sp. LR] and sp. SC2) have been reported to approach the required atmospheric methane oxidation kinetics, and to be capable of prolonged survival at atmospheric methane concentrations (43-45). For at least Methylocystis sp. SC2, and in all probability the other strains, their ability to use low methane concentrations is due to a pMMO isozyme (43). Methylocystis sp. SC2 has two pmoCAB operons, each encoding a different pMMO, and Baani and Liesack (43) demonstrate that survival at low methane concentrations (10-100 ppm) is due to one isozyme, while growth at high methane concentrations (>600 ppm) is due to the other isozyme. However, while the strains have been reported to survive for up to three months under low methane concentrations, no growth has been observed when they have been incubated under atmospheric methane concentrations (1774 ppb; 1.74 ppm) (29, 43 -46). It is possible that Methylocystis strains are contributing to the consumption of atmospheric methane in well-drained soils by surviving via atmospheric methane oxidation for maintenance energy in between exposures to higher concentrations of methane that allow for its growth. West and Schmidt (47) found support for such a possibility, as they were able to stimulate methanogenesis after exposing a well-drained arctic tundra soil to anaerobic conditions. Upon the soil’s subsequent return to aerobic conditions atmospheric methane was consumed at a faster rate; indicating that the exposure of methanotrophs to greater methane concentrations is advantageous to the methanotrophic community and the soil methane sink. Thus, so long as the Methylocystis strains are occasionally exposed to greater levels of methane produced by methanogens stimulated via anaerobic conditions from anoxic soil microsites or occasional soil flooding, they will be able to persist for a prolonged time on atmospheric methane and contribute to the strength of the soil methane sink. While these frndings indicate that at least some cultured methanotrophs are contributing to the oxidation of atmospheric methane they are not likely to be the methanotrophs responsible for the majority of atmospheric methane oxidation. Instead, methanotrophs that have not yet been cultured are hypothesized to likely be responsible. Multiple studies looking at soils from throughout the world using culture independent molecular surveys have found as yet uncultured novel numerically dominant and phylogenetically distinct pmoA in well-drained soils (29, 46, 48-61). These pmoA sequences form distinct phylogenetic clusters apart from the cultured gamma- and alpha- Proteobacteria and Verrucomicrobia methanotrophs (Figure 1.2), and a selection of their names are: JRl, JR2, JR3, MR1, Cluster I, Cluster II, Cluster III, Cluster V, Upland Soil 10 Cluster-y (USC-y or WBSF-H), Upland Soil Cluster-or (U SC-or or RA14), and RA21 (44, 46, 48-50, 55, 57). As the sequences from these clusters are numerically dominant and commonly found in well-drained soils throughout the world, it is plausible to conclude that they are the methanotrophs responsible for the majority of biological atmospheric methane oxidation. However, it has not yet been demonstrated that any of these clusters can grow on atmospheric methane. Kolb et al. (29) found in situ pmoA expression for USC-or in a German forest soils, but could not detect Cluster I in situ pmoA expression. Cluster I is the only one of the clusters that has cultured representatives, but the strains have only been minimally characterized (55). None of the exact methane oxidizing capabilities (Km, Vmax, etc.) of any of the above clusters is known, and therefore a determination of the magnitude of their contribution to atmospheric methane oxidation awaits further supporting data. Linking Methane Consumption to the Methanotroph Community Although a full understanding of atmospheric methane oxidation will require the culturing of many uncultured methanotrophs, our ability to measure methanotroph diversity through the retrieval of the pmoA gene allows for the linking of methanotroph diversity and community structure to rates of methane consumption. An example of a model system where the link between the methanotroph community and methane consumption can be tested is in well-drained soils of the same soil type that differ in their land use. The soil methane sink is significantly affected by land use with conversion of soils to agricultural use, based on worldwide measurements, leading to an approximately 11 70% reduction in net methane consumption, and an approximately 100 year recovery to the methane consumption rates of the native land use (reviewed by (62)). Land use conversion to agricultural land-use introduces many environmental changes, and studies examining total methane flux have found that fertilization, tillage, pesticide and herbicide application, changes in water filled pores space, dry bulk density, and pH can attenuate methane consumption in agricultural soils (reviewed in (63, 64)). In particular, fertilization is an acute disturbance to methanotrophs. Ammonia is a known inhibitor of MMO (63, 65, 66), and long-term application of fertilizer negatively impacts the methanotroph community (30, 54). However, paired sites that differ in land use share many environmental characteristics like climate and soil type that better facilitates our ability to identify factors that are impacting the methanotroph community and rates of methane consumption due to conversion to agricultural soils. An example of one such paired site, and the one used to investigate the methanotroph community in this thesis is the Kellogg Biological Station Long Term Ecological Research Site (KBS LTER; http://lter.kbs.msu.edu). KBS LTER features a range of different agricultural treatments as well as an entire successional gradient (early, mid and late successional soils). There are at least three replicates of each land use, and for many of the treatments in situ measures of methane consumption have been regularly made between March-December since 1992. Like other paired sites, the magnitude of the soil methane sink at KBS LTER is affected by land use. Rates of methane oxidation increase along a successional gradient: conventional row-crop agriculture soils consume the least amount of methane, mid- 12 successional fields have intermediate rates of methane consumption, and late successional forests have the highest rates. On a yearly average, the late successional forest soils consume 9.17 g CH4-C ha'1 day-1, roughly 6 times more methane than the conventional row-crop agricultural soils (1.62 g CI-I4-C ha'1 day-1) (67). The robust gas measures, the large rate differences between the soils of the successional gradient, and replicated plots make KBS LTER an ideal choice for exploring the link between methanotroph diversity and community structure to rates of methane consumption. In addition, KBS LTER features sub-plots in the late successional forest that allow for the determination of how one factor associated with row-crop agricultural management, long-term fertilization, affects methane consumption and the methanotroph community. Previous studies have not tried to correlate methanotroph richness to rates of methane consumption as methane consumption rates have been well examined in soils of varying land use (reviewed by (62)), but studies of the methanotroph community in comparable sites have typically only taken a limited number of rates measurements (46, 53, 56, 59, 60). The results from these studies have been inconclusive with no clear relationship observed between methanotroph richness and methane consumption (46, 53, 68). Methanotroph community changes have been observed across other successional gradients, with patterns of methanotroph diversity changing along the successional gradients of reforested comfields (46) and reclaimed pasture lands (68). KBS LTER features more land use types in its successional gradient then either of those study sites and coupled with robust gas measurements can better resolve patterns between the 13 methanotroph community, rates of methane consumption, and land use then any previous study. Summary In order to fully understand the causes of the increase in atmospheric methane, there are many areas of the global methane budget that require greater understanding. One such area is the soil methane sink, the only biological methane sink, where the activity of aerobic methanotrophs results in the net consumption of atmospheric methane. The methanotrophs that are probably responsible for the consumption of atmospheric methane have not yet been cultured, and a range of non-cultured methanotrophs have been revealed in culture-independent molecular surveys. Land use changes, particularly the conversion of native lands to row-crop agriculture, result in a dramatic drop in the rate of methane consumption, but changes in land use have not been linked to changes in methanotroph community diversity and structure. Paired sites of differing land use and changed rates of methane consumption, like those featured at KBS LTER, represent ideal model systems for linking methanotroph diversity to rates of methane consumption. Thesis Overview The main goals of this thesis is to (1) determine how the methanotroph community changes along with rates of methane consumption, (2) to begin to investigate factors that may be causing the changes, and (3) to determine if our findings at KBS LTER are applicable to other sites. By doing so, insights can be gained into how lands might be managed to enhance the methanotroph community, and in turn, the capacity of soil methane sink. Therefore, the main overall questions of this thesis were: 14 (a) Do soils of various land uses at KBS LTER, with different rates of methane consumption, harbor different methanotroph communities? (b) Can the differences in the methanotroph community be directly related to changes in the rate of methane consumption? (c) At KBS LTER is a decline in methanotroph richness associated with long-term fertilization alone? (d) Are there typical changes to methanotroph communities in response to long- term fertilization? To answer questions (a) and (b), chapter 2 presents the findings of assessments of the methanotroph community from the successional gradient at KBS LTER We find that the conversion of native lands to row-crop agriculture causes the loss of methanotroph richness, and a correlated decrease in the rate of methane consumption. We also find that land uses harbor different microbial communities that appear to change in parallel to differences in the plant community. To determine the importance of methanotroph diversity to rates we find that the recovery of both methanotroph richness and methane consumption is linear, concurrent and will take approximately 75 years following abandonment fi'om agriculture. The linear trajectory and lack of a step-wise increase in methane consumption rates following any increase in methanotroph diversity suggests that every methanotroph taxon, and not just a few taxons, are contributing to the rate of methane consumption at the KBS LTER Further supporting the finding of complementarity within the methanotroph community is the comparison of correlations of methane consumption with methanotroph richness, carbon dioxide efflux with total bacterial richness, and of both gas fluxes with moisture 15 and temperature. We find that carbon dioxide efflux, a process that is the result of a highly redundant microbial community does not correlate with total bacterial richness, and that more of the carbon dioxide flux can be explained by moisture and temperature. To answer questions (c) and (d), in chapter 3 I present the findings from assessments of the methanotroph community from KBS LTER late successional forest fertilized sub-plots, from agricultural and forest soils from the Rothamsted Research site, and from previously published studies of methanotroph communities. We find that long term fertilization alone is not causing a decrease in methanotroph richness nor in rates of methane consumption at KBS LTER This result is consistent with our findings in Chapter 2 that linked methanotroph richness and rates of methane consumption, and led to the expectation that the response of richness and rates of methane consumption to long-term fertilization would be the same. No consistent decline in methanotroph richness is observed in other long term fertilized soils, and methanotroph communities display a distinct biogeography with communities clustering together based on geographic location. There is no pattern of typical methanotroph community changes in accordance with long-term fertilization. Based on these findings we conclude that the each methanotroph taxon at KBS LTER provides a similar contribution to rate of methane consumption, and with every additional methanotroph in the soil there is an increase in the rates of consumption. The causes behind the change in richness remain unclear, as they are not caused by long-term fertilization alone, and other agricultural land-use associated changes are implicated as helping to cause the decrease in methanotroph richness in the Ag HT treatment at KBS LTER. The results also indicate that the findings at KBS LTER may not be directly 16 applicable to other soils possibly due to each site’s unique methanotroph community and soil properties. Further experiments that would clarify the effects of land use on the methanotroph community are discussed in Chapter 4. Three appendices are also provided. One details the PCR bias associated with the reaction conditions used in this study, and how the finding led to exclusion of abundance measures in the methanotroph community analyses. The second appendix presents NH4+ and N03' nutrient data fi'om a fertilized late successional forest sub-plot at KBS LTER before and after fertilization. The sub-plot’s methanotroph community was not assessed, but the findings reinforce the expectation that the methanotrophs in the fertilized sub- plots are exposed to ammonia following fertilization, and highlight the active nitrification in KBS LTER late successional forest soils. The third appendix confirms the finding of methanotroph biogeography in long-term fertilized soils by assessing the clustering of methanotroph community composition in a variety of well-drained soils. 17 .6552: 95 0:28.: .6 :osmExo 05 $558.: 533 .ommgwxxoocofi 052:2: 2: m_ 0::an m8— of. detects 2856:. ..o .3358 0:832: 05 .8 ocmfiozom “BE—:86 < .2 8:9“. mmanE 33:8 3.: wouaeomoofi ON: NC w 0% + m TEIIV O ommgwobigew EOOOEA ommaewouwznow OEUEA omwswegsow :0 mo emacowbnoo:oz EU Bane—om efiEeEagom 35522 89502 18 a—proteobacteria Cluster V [Upland Soil Cluster 0493‘] y-proteobacteria amoA y—proteobacteria Type I methanotrophs Cluster Ill Upland Soil Cluster 7 V =Venuoomicrobia pmoA cggnnmr n «cum ll ' mm Unknown Classification / fi-proteobacten’a amoA Crenothrix polyspora) y—proteobacteria \ Cluster ll Presumably a—proteobaclen'a Cmnarachaeota amoA Figure 1.2. Phylogenetic tree of selected partial pmoA and amoA protein sequences from public databases. The tree is based on 164 amino acid positions using Phylip Protein Maximum Likelihood as implemented in ARB (Ludwig et al. 2004). Boxed labels are indicative of pmoA clades that have been found in soils, have not been cultured, and are thought to likely play a substantial role in the oxidation of atmospheric methane. The scale bar represents 10 PAM units. 19 References 1. 10. IPCC (2007) Climate Change 2007: Synthesis Report. Contribution of Working Groups I, II and III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. eds Team CW, Pachauri RK, & Reisinger AGeneva, Switzerland). Forster P, et al. (2007) Changes in Atmospheric Constituents and in Radiative Forcing. Climate Change 200 7: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, eds Solomon S, Qin D, Manning M, Chen Z, Marquis M, Averyt K, Tignor M, & Miller H (Cambridge University Press, Cambridge), pp 129-234. Rigby M, et al. (2008) Renewed Growth of Atmospheric Methane. Geophysical Research Letters 35:doi:10.1029/2008GL036037. Bousquet P, et al. (2006) Contribution of anthropogenic and natural sources to atmospheric methane variability. Nature 443(7110):439-443. Denman KL, et al. (2007) Couplings Between Changes in the Climate System and Biogeochemistry. Climate Change 200 7: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, eds Solomon S, Qin D, Manning M, Chen Z, Marquis M, Averyt K, Tignor M, & Miller H (Cambridge University Press, Cambridge), pp 499-587. Keppler F, Hamilton If G, Brass M, & Rockmann T (2006) Methane emissions fiom terrestrial plants under aerobic conditions. Nature 439(7073): 1 87-191. Kirschbaum MUF, et al. (2006) A comment on the quantitative significance of aerobic methane release by plants. Functional Plant Biology 33(6):521-530. Ferretti DF, et al. (2007) Stable isotopes provide revised global limits of aerobic methane emissions from plants. Atmospheric Chemistry and Physics 7:237-241. Keppler F, et al. (2008) Methoxyl groups of plant pectin as a precursor of atmospheric methane: evidence fi'om deuterium labelling studies. New Phytologist 178(4):808-814. Wang Z-P, Han X-G, Wang GG, Song Y, & Gulledge J (2008) Aerobic Methane Emission from Plants in the Inner Mongolia Steppe. Environmental Science & Technology 42(1):62-68. 20 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. Vigano I, et al. (2008) Effect of UV radiation and temperature on the emission of methane from plant biomass and structural components. Biogeosciences 5(3):937- 947. McLeod AR, et al. (2008) Ultraviolet radiation drives methane emissions fiom terrestrial plant pectins. New Phytologist 180(1):124-132. Kirschbaum MUF & Walcroft A (2008) No detectable aerobic methane efflux from plant material, nor from adsorption/desorption processes. Biogeosciences 5(6):1551-1558. Beerling DJ, Gardiner T, Leggett G, McLeod A, & Quick WP (2008) Missing methane emissions from leaves of terrestrial plants. Global Change Biology 14(8):1821-1826. Dueck TA, et al. (2007) No evidence for substantial aerobic methane emission by terrestrial plants: a 13C-labelling approach. New Phytologist l75(1):29-35. Wuebbles DJ & Hayhoe K (2002) Atmospheric methane and global change. Earth-Science Reviews 57(3 -4):l77-210. McDonald IR, Bodrossy L, Chen Y, & Murrell JC (2008) Molecular Ecology Techniques for the Study of Aerobic Methanotrophs. Appl. Environ. Microbiol. 74(5):1305-1315. Chanton JP, Powelson DK, & Green RB (2009) Methane oxidation in landfill cover soils, is a 10% default value reasonable? Journal of Environmental Quality 38(2):654-663. Islam T, Jensen S, Reigstad LJ, Larsen 0, & Birkeland NK (2008) Methane oxidation at 55 degrees C and pH 2 by a thermoacidophilic bacterium belonging to the Verrucomicrobia phylum. Proceedings of the National Academy of Sciences of the United States of America 105(1):300-304. Pol A, et al. (2007) Methanotrophy below pH 1 by a new Verrucomicrobia species. Nature 450(7171):874-878. Dunfield PF, et al. (2007) Methane oxidation by an extremely acidophilic bacterium of the phylum Verrucomicrobia. Nature 450(7171):879-882. Hou SB, et al. (2008) Complete genome sequence of the extremely acidophilic methanotroph isolate V4, Methylacidiphilum infemorum, a representative of the bacterial phylum Verrucomicrobia. Biology Direct 3. Hanson RS & Hanson TE (1996) Methanotrophic bacteria. Microbiological Reviews 60(2):439-+. 21 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. Caldwell SL, et al. (2008) Anaerobic oxidation of methane: mechanisms, bioenergetics, and the ecology of associated microorganisms. Environmental Science & Technology 42(18):6791-6799. Moran JJ, et al. (2008) Methyl sulfides as intermediates in the anaerobic oxidation of methane. Environmental Microbiology 10(1): 162-173. Pemthaler A, et al. (2008) Diverse syntrophic partnerships from-deep-sea methane vents revealed by direct cell capture and metagenorrrics. Proceedings of the National Academy of Sciences of the United States of America 105(19):7052- 7057. Ettwig KF, et al. (2008) Denitrifying bacteria anaerobically oxidize methane in the absence of Archaea. Environmental Microbiology 10(1 1):3164-3173. Holmes AJ, Costello A, Lidstrom ME, & Murrell JC (1995) Evidence that participate methane monooxygenase and ammonia monooxygenase may be evolutionarily related. FEMS Microbiology Letters l32(3):203-208. Kolb S, Knief C, Dunfield PF, & Conrad R (2005) Abundance and activity of uncultured methanotrophic bacteria involved in the consumption of atmospheric methane in two forest soils. Environmental Microbiology 7(8):1150-1161. Gulledge J, Hrywna Y, Cavanaugh C, & Steudler PA (2004) Effects of long-term nitrogen fertilization on the uptake kinetics of atmospheric methane in temperate forest soils. FEMS Microbiology Ecology 49(3):389-400. Carini SA, Orcutt BN, & Joye SB (2003) Interactions between methane oxidation and nitrification in coastal sediments. Geomicrobiology Journal 20(4):355-3 74. Kitrnitto A, Myronova N, Basu P, & Dalton H (2005) Characterization and structural analysis of an active particulate methane monooxygenase trimer from Methylococcus capsulatus (Bath). Biochemistry 44(33):]0954-10965. Dunfield PF, Khmelenina VN, Suzina NE, Trotsenko YA, & Dedysh SN (2003) Methylocella silvestris sp nov., a novel methanotroph isolated from an acidic forest carnbisol. International Journal of Systematic and Evolutionary Microbiology 53: 123 1-1239. Gilbert B, et al. (2000) Molecular analysis of the pmo (particulate methane monooxygenase) operons from two type II methanotrophs. Applied and Environmental Microbiology 66(3):966-975. Ward N, et al. (2004) Genomic insights into methanotrophy: The complete genome sequence of Methylococcus capsulatus (Bath). Plos Biology 2(10):1616- 1 628. 22 36. 37. 38. 39. 40. 41. 42. 43. 45. 46. 47. Martinho M, et al. (2007) Mosssbauer studies of the membrane-associated methane monooxygenase from Methylococcus capsulatus bath: Evidence for a diiron center. Journal of the American Chemical Society 129(5 1): 15783-+. Lieberman RL & Rosenzweig AC (2005) Crystal structure of a membrane-bound metalloenzyme that catalyses the biological oxidation of methane. Nature 434(7030):177-182. Hakemian AS, et al. (2008) The metal centers of particulate methane monooxygenase from Methylosinus trichosporium OB3b. Biochemistry 47(26):6793-6801 . Chan SI, et al. (2007) Redox potentiometry studies of particulate methane monooxygenase: Support for a trinuclear c0pper cluster active site. Angewandte Chemie-International Edition 46(12): 1 992-1 994. Chan SI & Yu SSF (2008) Controlled oxidation of hydrocarbons by the membrane-bound methane monooxygenase: The case for a tricopper cluster. Accounts of Chemical Research 41(8):969-979. Rosenzweig AC (2008) The metal centres of particulate methane mono- oxygenase. Biochemical Society Transactions 36: 1 134-1 137. Bender M & Conrad R (1992) Kinetics of CH4 Oxidation in Oxic Soils Exposed to Ambient Air or High CH4 Mixing Ratios. FEMS Microbiology Ecology 101(4):261-270. Baani M & Liesack W (2008) Two isozymes of particulate methane monooxygenase with different methane oxidation kinetics are found in Methylocystis sp strain SCZ. Proceedings of the National Academy of Sciences of the United States of America 105(29): 10203-10208. Knief C & Dunfield PF (2005) Response and adaptation of different methanotrophic bacteria to low methane mixing ratios. Environmental Microbiology 7(9): 1307-13 17. Dunfield PF & Conrad R (2000) Starvation alters the apparent half-saturation constant for methane in the type H methanotroph Methylcystis strain LRl. Applied and Environmental Microbiology 66(9):4136-4138. Knief C, et al. (2005) Diversity of methanotrophic bacteria in tropical upland soils under different land uses. Applied and Environmental Microbiology 71 (7):3826- 3 83 1 . West AE & Schmidt SK (2002) Endogenous methanogenesis stimulates oxidation of atmospheric CH4 in alpine tundra soil. Microbial Ecology 43(4):408-415. 23 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. Holmes AJ, et al. (1999) Characterization of methanotrophic bacterial populations in soils showing atmospheric methane uptake. Applied and Environmental Microbiology 65(8):3312-33 18. Henckel T, Jackel U, Schnell S, & Conrad R (2000) Molecular analyses of novel methanotrophic communities in forest soil that oxidize atmospheric methane. Applied and Environmental Microbiology 66(5): 1 801-1808. Horz H-P, Rich V, Avrahami S, & Bohannan BJM (2005) Methane-Oxidizing Bacteria in a California Upland Grassland Soil: Diversity and Response to Simulated Global Change. Appl. Environ. Microbiol. 71(5):2642-2652. Jaatinen K, Knief C, Dunfield PF, Yrjala K, & Fritze H (2004) Methanotrophic bacteria in boreal forest soil after fire. FEMS Microbiology Ecology 50(3): 195- 202. King GM & Nanba K (2008) Distribution of Atmospheric Methane Oxidation and Methanotrophic Communities on Hawaiian Volcanic Deposits and Soils. Microbes and Environments 23(4):326-330. Knief C, Lipski A, & Dunfield PF (2003) Diversity and activity of methanotrophic bacteria in different upland soils. Applied and Environmental Microbiology 69(1 l):6703-6714. Lau E, Ahmad A, Steudler PA, & Cavanaugh CM (2007) Molecular characterization of methanotrophic communities in forest soils. FEMS Microbiol Ecol 60(3):490-500. Pacheco-Oliver M, McDonald IR, Groleau D, Murrell JC, & Miguez CB (2002) Detection of methanotrophs with highly divergent pmoA genes from Arctic soils. FEMS Microbiology Letters 209(2):3 13-3 19. Reay DS, Radajewski S, Murrell JC, McNamara N, & Nedwell DB (2001) Effects of land-use on the activity and diversity of methane oxidizing bacteria in forest soils. Soil Biology & Biochemistry 33(12-13): 1613-1623. Ricke P, Kolb S, & Braker G (2005) Application of a newly developed ARB software-integrated tool for in silico terminal restriction fragment length polymorphism analysis reveals the dominance of a novel pmoA cluster in a forest soil. Applied and Environmental Microbiology 71(3): 1671-1673. Singh BK & Tate K (2007) Biochemical and molecular characterization of methanotrophs in soil from a pristine New Zealand beech forest. FEMS Microbiology Letters 275(1):89-97. 24 59. 60. 61. 62. 63. 65. 66. 67. 68. Singh BK, et al. (2007) Effect of afforestation and reforestation of pastures on the activity and population dynamics of methanotrophic bacteria. Applied and Environmental Microbiology 73(16):5 153-5 161 . Zhou XQ, et al. (2008) Effects of grazing by sheep on the structure of methane- oxidizing bacterial community of steppe soil. Soil Biology & Biochemistry 40(1):258-261. Knief C, Kolb S, Bodelier PLE, Lipski A, & Dunfield PF (2006) The active methanotrophic community in hydromorphic soils changes in response to changing methane concentration. Environmental Microbiology 8(2):321-333. Smith KA, et al. (2000) Oxidation of atmospheric methane in Northern European soils, comparison with other ecosystems, and uncertainties in the global terrestrial sink. Global Change Biology 6(7):791-803. Hutsch BW (2001) Methane oxidation in non-flooded soils as affected by crop production - invited paper. European Journal of Agronomy 14( 4):23 7- 260. King GM (1997) Responses of atmospheric methane consumption by soils to global climate change. Global Change Biology 3(4):351-362. Gulledge J & Schimel JP (1998) Low-Concentration Kinetics of Atmospheric CH4 Oxidation in Soil and Mechanism of NH4+ Inhibition. Appl. Environ. Microbiol. 64(1 1):4291-4298. Suwanwaree P & Robertson GP (2005) Methane Oxidation in Forest, Successional, and No-till Agricultural Ecosystems: Effects of Nitrogen and Soil Disturbance. Soil Science of America Journal 69: 1722-1 779. Robertson GP, Paul EA, & Harwood RR (2000) Greenhouse Gases in Intensive Agriculture: Contributions of Individual Gases to the Radiative Forcing of the Atmosphere. Science 289: 1922-1925. Zhou XQ, Wang YF, Huang XZ, Tian JQ, & Hao YB (2008) Effect of grazing intensities on the activity and community structure of methane-oxidizing bacteria of grassland soil in Inner Mongolia. Nutrient Cycling in Agroecosystems 80(2):l45-152. 25 Chapter 2 Agriculture’s Impact on Microbial Diversity and the Flux of Greenhouse Gases Abstract Row-crop agriculture impacts both the production of carbon dioxide and the consumption of methane by microbial communities in upland soils - Earth’s largest biological sink for atmospheric methane. To determine if there are relationships between the rates of these ecosystem level processes and the diversity of microbes that catalyze them, we measured the in situ fluxes of methane and carbon dioxide, and conducted molecular surveys of methane—oxidizing bacteria (methanotrophs) and total bacterial diversity across a range of land uses. Rates of methane consumption and methanotroph diversity were positively correlated, as conversion of lands to row-crop agriculture led to a 7-fold reduction from maximal values found in native deciduous forests. In fields abandoned from agriculture the diversity of methanotrophs and the consumption of methane increased monotonically, suggesting that managing lands to conserve or restore methanotroph diversity could help mitigate increasing atmospheric concentrations of this potent greenhouse gas. In addition, methanotroph diversity was more capable at explaining methane consumption than was moisture and temperature. Conversely, total bacterial diversity did not correlate with changes in carbon dioxide emission, but did correlate to moisture and temperature. Taken together, these results are consistent with the prediction that ecosystem processes are more likely to be influenced by microbial diversity when microbial communities have limited species richness. 26 Introduction Carbon dioxide (C02) and methane (CH4) are responsible for approximately 80% of the positive radiative forcing of the atmosphere from long-lived greenhouse gases (approximately 62% and 18%, respectively) (1). The atmospheric concentrations of both gases now exceed any of their respective levels over the past 65 0,000 years, and continue to rise (2, 3). Fluxes of both gases are affected by land-use changes — especially deforestation and row-crop agriculture (4, 5). Conversion of native upland soils (e.g. forest, grassland) to agricultural management reduces their capacity for methane consumption by an average of 71%, and recovery of methane consumption is estimated to be ca. 100 years following the cessation of agriculture (reviewed in (5)). Deforestation and agricultural development increases the efflux of carbon dioxide fiom soil (4, 6), and is estimated to have contributed approximately 25% of the increase in radiative forcing from carbon dioxide since 1850 (1). Microbes directly control methane consumption, and portions of soil carbon dioxide efflux (7, 8). Aerobic methane-oxidizing bacteria (methanotrophs) consume methane in well-drained soils, and the consequential net uptake of methane constitutes Earth’s largest known biological sink for atmospheric methane (9, 10). Soil respiration is a large biological source of atmospheric carbon dioxide, and is mainly a reflection of microbial respiration (reviewed in (11)) plus respiration from roots. The relative contribution of root and heterotrophic microbial respiration has proven difficult to disentangle (12). Nevertheless, changes in soil respiration have been associated with variations in the microbial community (8, 13). 27 How row-crop agriculture changes the methanotroph and the heterotrophic microbial community and their associated rates of methane consumption and carbon dioxide efflux remains unclear. The Kellogg Biological Station Long Term Ecological Research Site (KBS LTER), with replicated plots of the same soil series (14) that just differ in land use (Table 2.1), represents an ideal system for determining how row-crop agriculture impacts microbial communities and the greenhouse gas fluxes they control. Past studies have typically investigated either gas flux or the microbial community at analogous sites, but rarely has the function mediated by the microbes and the microbial community been investigated simultaneously in agricultural soils or across a land-use gradient. We correlated in situ fluxes of carbon dioxide and methane to assessments of microbial diversity across a successional gradient featuring 5 different land uses at KBS LTER (Table 2.1). Methane consumption rates have been well examined in soils of varying land use (reviewed by (5)), but studies of the methanotroph community in comparable sites have either not measured rates of methane consumption (15), taken a maximum of two rate measurements (16-19), not featured replicated sites (16, 18, 19), and have only examined a limited number of successional stages (15-19). No clear relationship has been observed between methanotroph richness and methane consumption in any study (18-20), and, to our knowledge, ours is the first study that has attempted to correlate methanotroph richness to rates of methane consumption. Methanotroph community changes have been observed across other land-use gradients, with patterns of methanotroph diversity changing along a gradient of grazing intensities (20), and along a successional gradient of reforested comfields (18). KBS LTER features more land use types in its successional 28 gradient then any previous study, and coupled with robust gas measurements can better resolve patterns between the methanotroph community, rates of methane consumption, ' and land use. Similarly, few studies have attempted to determine the effects of row-crop agriculture on both soil respiration and the heterotrophic microbial community. Studies have documented land-use associated changes to the microbial community due to the conversion of forest to pasture (20, 21), but rates of soil respiration were not measured. Others have linked changes in soil respiration to changes in the microbial community (8, 13, 22), but agricultural soils were not included in any of those studies. We hypothesize that microbial diversity is more likely to be important to a specialized metabolic process like methane oxidation than it is to be a factor in a metabolically redundant process like carbon dioxide production (23-25). Specialized metabolic processes like methane consumption are typically the result of relatively taxonomically narrow microbes, and a limited number of species contribute to the ecosystem process, resulting in a lack of fimctional redundancy. On the other hand, soil respiration is a process that is the result of a multitude of taxonomically diverse microbes and enzymatic processes, and the number of microbes found to be contributing to carbon dioxide production is likely to be high. Therefore, we expect that methanotroph richness will correlate to rates of methane consumption, but total bacterial richness will not correlate to carbon dioxide efflux. In addition, by measuring methanotroph diversity and methane consumption in fields that had been abandoned from agriculture for various lengths of time, we should to be able to infer if the postulated positive relationship between net methane consumption 29 and methanotroph diversity is due complementarity or selection. The complementarity and selection hypotheses are actively debated to explain positive relationships between the magnitude of an ecosystem process and species richness (26, 27). The complementarity hypothesis gives importance to every species in determining the magnitude of the ecosystem process, and postulates that increased ecosystem function results from the combined activity of species in complementary niches. The selection or sampling effect hypothesis offers an alternative explanation: the positive relationship between diversity and function is driven by one or a few dominant species particularly proficient at performing the process under study. Discerning between the two hypothesis is important because if the postulated positive relationship between net methane consumption and methanotroph diversity is observed, it will be useful to identifying whether all methanotroph diversity, or only select species need to be conserved or restored in order to possibly enhance the methanotroph community, and in turn, the capacity of the soil methane sink. Materials and Methods Site Description This study was conducted at the Kellogg Biological Station Long Term Ecological Research Site (KBS LTER; http://lter.kbs.msu.edu) located in Hickory Comers, Michigan. Soils of five treatments were examined in this study: conventional agricultural management of historically tilled land (Ag HT), early successional plant communities on fields that had been abandoned from agriculture in 1989 (Early HT), mid-successional plant communities on either historically tilled land (Mid HT) or never 30 tilled grassland (Mid NT), and a never tilled late successional deciduous forest (Late DF). Additional site descriptions can be found in Table 2.1. All of the KBS LTER soils are located within 3 km of the main experimental site, are all of the Kalamazoo/Oshtemo soil series, and are well-drained Typic Hapludalfs (fine or course loamy, mixed). Rate Measurements In situ rates of methane consumption were measured using closed-cover flux chambers (28). Rate measurements were typically made twice monthly between March and December at 3 or 4 replicate plots between 1992 and 2007 for Ag HT, Early HT and Late DF, and between 1992 and 1997 for Mid NT and Mid HT. Mid HT rate measurements were also made in 2002. Individual rate measurements and detailed methods are available at http://lter.kbs.msu.edu. To determine rate differences based on treatments, analysis of variance (AN OVA) was performed using PROC MIXED (SAS Inc, 2002). Soil Sampling Methanotroph diversity was assessed in 5 soil cores (2.5 x 10 cm) collected fi'om 3 replicates along a gradient of 5 land uses (Table 2.1). These 75 samples, collected from a total of 15 experimental fields on 13 June 2006, as well as samples used for additional molecular surveys collected on 8 December 2004, 6 June 2005, and 13 June 2007 were transported to the laboratory on ice where they were mixed thoroughly and flash frozen in a dry ice/ethanol bath, then stored at -80°C until processing. Total bacterial diversity was assessed in 5 soil cores (5 x 10 cm) from two Ag HT and two Late DF replicates in December 2006. Samples were pooled, transported to the laboratory on ice, sieved and stored at -80°C. 31 DNA Extraction and pmoA PCR Reactions For the assessment of total bacterial diversity, DNA was extracted according to Zhou et al. (29), followed by a cesium-chloride gradient purification (30). For the assessment of methanotroph diversity, DNA was extracted from soil samples with the Mo Bio PowerSoilTM DNA Isolation Kit, following the manufacturer’s protocol, except that mechanical cell lysis was performed by bead beating for 45 seconds. All soil samples were screened for genes coding for the A subunit of particulate methane monooxygenase (pmoA) via PCR amplification with the primer sets A189 (5’- GGNGACTGGGACTTCTGG-B’) -A682 (5’-GAASGCNGAGAAGAASGC-3’) (31), and A189-mb66l (5’-CCGGMGCAACGTCYTTACC-3’) (32) in order to encompass all known pmoA genes (3 3). No amplification was observed from A189-mb661. Amplification reactions contained either 25, 45 or 90 ng of undiluted DNA, 1.25 ul 1% BSA, 10 mM dNTPs, 0.2 uM of each primer, 2.5 ul 10x PCR buffer (200mM Tris-HCl (pH 8.4) and 500 mM KCl), 0.5 pl 50 mM MgC12 and 1.5 U of T aq DNA polymerase (Invitrogen, Carlsbad, CA) in a total volume of 25 ul. Reaction conditions for the 62°C-56°C clone libraries and for tRF LP (see below) were 95°C for 5 minutes, 15 cycles of a ‘touchdown PCR’ of 95°C for 1 minute, 62°C for 1 minute (-O.4°C each cycle to 56°C), and 72°C for 1 minute, 15 cycles using 56°C as the annealing temperature, and a final 10 minute extension at 7 2°C. To ensure that no methanotroph diversity was missed, additional libraries were constructed under the same conditions, but with varying annealing temperatures: 60°C to 51°C (-0.6°C each cycle to 51°C), 48°C, and 51°C (Table 2.2). Clone Libraries 32 Cloning was performed with the TOPO-TA cloning kit (Invitrogen, Carlsbad, CA) using either vector pCR 4 or pCR 2.1 as per the manufacturer’s protocol. Transformants were screened via PCR reactions with the primers F2 (5 ’- CAGTCACGACGTTGTAAAACGACGGC-3’) and R4 (5’- CAGGAAACAGCTATGACCATG-3’) (34). One-1.25 ul of the PCR products were purified via incubation with 0.25 ul of ExoSAP-IT (U sb, Cleveland, 0H) for 30 minutes at 37°C. Sequencing was completed at the Research Technology Support Facility at Michigan State University (RTSF). Sequences identified by BLASTX (3 5) as pmoA or the A subunit of ammonia monoxygenase (amoA), which A189-A682 also amplify, were imported into Arb (36). In Arb, sequences were translated and aligned using Clustal W. Nucleic acid sequences were aligned according to the protein sequence. Sequences from clone libraries were determined to be the same species if they were 2 94% identical (3 7) as determined by DOTUR (average neighbor grouping) (3 8). Soil cores collected in December 2004 and June 2005 were used to construct 62°C-56°C clone libraries from 5 cores from the l”t replicates of Ag HT and Late DF. For June 2005 samples, libraries represent sequences pooled from separate DNA extractions, PCR reactions and cloning reactions. These libraries also include clones whose identities were inferred from identical banding patterns in restriction fragment length polymorphism analyses using the Alwl enzyme (N eb, Ipswich, MA). The same June 2005 soil samples were also used to construct libraries from 48°C, 51 °C, and 60°C- 51°C annealing temperatures. For these annealing temperatures, separate DNA 33 extractions were pooled for 4 replicate PCR amplifications, which were subsequently pooled into 1 cloning reaction. Libraries fiom June 2006 samples were constructed from duplicate 60°C-51°C PCR reactions from each of 5 soil cores from the lSt and 3rd Late DF and 1st and 2lfld Ag HT replicates which were pooled according to replicate. Prior to cloning these PCR products, as well as those from June 2007, were digested with the restriction enzyme PflFI (N eb, Ipswich, MA), and gel extracted with the PrepEase kit (U sb, Cleveland, OH) to reduce the incidence of cloning amoA and non-specific PCR products. Additional 62°C -56°C clone libraries were constructed from a collection of the June 2006 soil cores from various treatments where large unidentified peaks were observed in the tRF LP analysis (see below). The libraries were constructed in order to match more tRF LP peaks with known sequences, and therefore were not sampled until pmoA rarefaction curves were asymptotic. Sequences representing each of the pmoA species found in this study are deposited in GenBank under accession numbers FJ529724 - FJ529808, and GQ219582- GQ2195 83. Terminal Restriction Fragment Length Polymorphism (tRFLP) Analysis For tRF LP (39) each of the 75 soil cores from June 2006 was amplified for pmoA with A682 labeled with fluorophore 6-carboxyfluorescein (6-F am), and template addition was always 45 ng. To obtain enough DNA for tRF LP at least 4 replicate PCR reactions were pooled. PCR products were purified using a MinElute column (Qiagen, Valencia, CA). In a 50 u] reaction, 300ng of purified product was digested with 1.5 U Taul (F ermentas, Glen Burnie, MD) by incubating at 55°C for 1 hr 30 min. To inactivate the 34 enzyme the DNA was precipitated as follows: The sample was diluted to 500 it], followed by the addition of 50 ul 3M sodium acetate, 1 ul or 2.5 pl 10 mg/ml glycogen, and 500 pl isopropanol, and holding on ice for at least 5 minutes. The DNA was then pelleted by centrifugation at 16,000 x g for either 5 (1 ul glycogen) or 10 minutes (2.5 ul glycogen). The supernatant was decantated, and the pellet washed with 500p] of 80% ethanol followed by centrifugation for 2 minutes at 16,000 x g, and removal of the supernatant. After a 30 second centrifugation additional ethanol was removed, and the DNA was air dried for 5-10 minutes before resuspension in 20 ul of water. In an 18-22 ul reaction, either 140 ng (lul glycogen) or 160 ng (2.5ul glycogen) of DNA was digested with 2.5 U SspI (Neb, Ipswich, MA) in a 1hr 30min incubation at 37°C. After heat inactivation at 65°C for 20 minutes 6 U of BstUI (Neb, Ipswich, MA) was added, incubated at 60°C for 1hr 30min, and inactivated by adding 0.8 ul of 0.5 mM EDTA. Capillary electrophoresis of the tRF LP reactions was then performed with a 5 fu cutoff at RTSF. To ensure that the tRFLP profiles could be compared, the distribution of the total fluorescence was compared with PROC UNIVARIATE (SAS Inc, 2002), and any outliers were excluded. Individual peaks were distinguished from the background signal and binned using TRFLP-Stats (40). In TRFLP-Stats default settings were used except for the standard deviation cutoff, which was increased to 4.5. The resulting cutoff of approximately 25 fu excluded any profile whose highest peak area was not greater than 350, and ensured that peak areas below 300 were not identified as true peaks in greater than 90% of the samples. PCR products from clones representing 14pmoA OTUs, 1 amoA, and 21 non-specific (neither pmoA or amoA) were run as tRF LP at least once. 35 These clone controls allowed the identification of specific tRF LP peaks as either pmoA OTUs, amoA, or non-specific PCR products, and allowed for the exclusion of amoA and non-specific PCR product bins. Those bins that were identified as either were excluded from the analysis. To obtain a better representation of an entire replicate, and to control for soil heterogeneity, bins from the same replicate were summed. For every replicate, tRFLPs from at least 3 soil cores were obtained and summed. In total, tRF LP profiles were obtained from 68 of the 75 soil cores. At least two negative controls (no DNA PCR reactions) were included on each plate of tRF LP reactions, and any profile whose highest peak area was not greater than 300, the highest peak area in these controls, was excluded from the analysis. To assess the reproducibility of the method, twenty samples had technical replicates whose results were then pooled after analysis in TRFLP-Stats. Each tRF LP bin then served as an Operational Taxonomic Unit (OTU), and the OTUs were used in a linear regression (PROC REG, SAS Inc, 2002) against the rate of methane consumption at KBS-LTER in summer (J une-August), against the rate of methane consumption between March-August, and to calculate B-diversity with PAST (41) or Estimate S (http://viceroy.eeb.uconn.edu/EstimateS). 168 Tag Sequencing 16S tagged sequencing and analysis was performed on a 454 Life Science’s pyrosequencer as described by Sogin et al. (42). Richness estimates were then regressed against mean carbon dioxide measures taken between 2005-2007 using simple linear regression (PROC REG, SAS Inc, 2002). Multiple regression 36 To determine the ability of soil moisture and maximum air temperature to predict rates of methane consumption and carbon dioxide emission, we used multiple regression (PROC MIXED, SAS Inc, 2002) with data taken between 2005-2007 from Late DF and Ag HT. Moisture and soil temperature measurements were taken on the same day that gas flux was measured. Individual measurements are available at http://lter.kbs.msu.edu/datasets. Results Rate measurements Maximal rates of methane consumption and carbon dioxide emission were observed during the summer months for all treatments (Figure 2.1). Carbon dioxide emissions peaked May through August, and significant treatment effects were found (Figure 2.2a). The conventional row-crop agricultural soil (Ag HT) had the lowest average rate of soil respiration, and the early (Early HT) and mid-successional sites (Mid NT and Mid HT) had maximal rates. On average, the early and mid-successional sites emit approximately 79% more, and Late DF emits 18% more, carbon dioxide than Ag HT per day. Rates of methane consumption also changed according to treatment (Figure 2.2b) with the greatest consumption in Late DF followed by the mid-successional soils, Early HT, and the lowest rates in Ag HT (Figures 2.1 and 2.2b). The overall difference in rates between the highest rate in Late DF and the lowest rate in Ag HT is approximately 7-fold (9.98 g CH4-C ha'1 clay'1 and 1.29 g CH4-C ha'1 day-1, respectively), resulting in 7 times more methane being consumed per day in the late successional deciduous forest as 37 compared to the conventional row-crop agricultural soils. The difference between the Late DF and Ag HT treatments grew 2-fold with the inclusion of rates from 2000-2007 (28). The highest rates of methane consumption were observed June through September (Figure 2.1), the 7-fold difference in rates remained the same during those months, and coincided with when we obtained the majority of the soil samples used to assess the methanotroph community. Methanotroph Diversity and Community Composition pmoA Clone libraries Clone libraries of the pmoA gene, encoding the A-subunit of the particulate methane monooxygenase, the first enzyme in the pathway of methane oxidation and the defining enzyme of aerobic methanotrophs (43), were used to assess methanotroph diversity. Libraries from Late DF and Ag HT were constructed from various annealing temperatures, soil samples (Methods, Table 2.2), and until rarefaction curves were asymptotic (Figure 2.3) to ensure that all methanotroph diversity had been captured. The libraries constructed fi'om the various annealing temperatures revealed that the 60°C- 51°C annealing temperature yielded the most methanotroph species in an individual library (Table 2.2). Due to PCR bias, measures of pmoA abundance were excluded from comparisons of the methanotroph community (Appendix A). Although the A-subunit from some ammonia monooxygenase genes (amoA) amplified with the pmoA primers (31), amoA sequences were distinguished based on diagnostic amino acids (44, 45), and their clustering in phylogenetic trees (Figure 2.4, Figure 2.5). Digestion of the PCR products with the restriction enzyme PflFI (N eb, Ipswich, MA) was found to greatly reduce the incidence of cloning amoA, and to have no 38 change in the overall methanotroph richness recovered from individual libraries (Table 2.2). Phylogenetic analysis of pmoA genes revealed that methanotrophs in KBS LTER soils cluster within seven clades (Figure 2.4). Six pmoA clades - Cluster 1, Cluster II, KBSl , JRl, MR1, and Upland Soil Cluster a - were found in Late DF, compared with just 2 clades, Cluster I and Cluster II, in Ag HT. Cluster RA21 was only recovered from Mid NT (Figures 2.4). Grouping the pmoA sequences from the KBS LTER clone libraries at a species level (94% average nucleotide similarity) revealed an even more dramatic difference in the methanotroph richness between the Ag HT and Late DF land- use treatments: 24 methanotroph species in Late DF, and only 7 species were present in Ag HT (Table 2.2, Figure 2.5). Of the 7 methanotroph species found in Ag HT, 2 were unique to the treatment while the other 5 species were also found in Late DF . In both Late DF and Ag HT, more methanotroph species were found in Cluster I then in any other cluster, with 12 Late DF and 6 Ag HT species. Next was Cluster II with 7 Late DF and 1 Ag HT species. The other clusters had no more than 2 methanotroph species. Cluster KBSl was represented by 1 Late DF sequence fiom the libraries in this study, and additional sequences from the KBSl clade are reported from a clone library constructed from a subplot of the lSt Late-DF replicate that was sampled on 13 June 2007 with 60°C- 51°C as the annealing temperature (Chapter 3) to further confirm the presence of this unique cluster (Figure 2.5). Terminal Restriction Fragment Length Polymorphism (tRFLP) Analysis To survey methanotroph diversity across the entire gradient of KBS LTER land uses (Table 2.1), and in three replicates of each treatment, terminal Restriction Fragment 39 Length Polymorphism (tRF LP) of the pmoA gene was used with soil sampled during the peak period of methane consumption in 2006. The tRF LP assay distinguished 11 Operational Taxonomic Units (OTUs) that were confirmed as pmoA gene fiagments through sequence analysis, and the requirement for comigration with terminal restriction fragments from cloned controls. Similar to the clone libraries, richness differences between treatments were observed. The Late DF sites averaged 7 OTUs compared to 2 OTUs in Ag HT, and intermediate numbers of pmoA OTUs were found in the successional sites (Figure 2.6). In addition, following release from row-crop agriculture (Ag HT), the composition of the methanotroph communities in soils abandoned from agriculture (Early HT and Mid HT) became more like the native sites (Late DF and Mid NT) (Figures 2.8 and 2.9). These methanotroph communities are statistically different from one another according to a one-way analysis of similarities (p < 0.005, ANOSIM), with methanotrophs in the early successional soil beginning to diverge from the row-crop agricultural plots. Divergence continued in the Mid HT soils that have been abandoned from agricultural for ca. 50 yrs., such that these communities began to overlap in composition with communities of methanotrophs in native fields (Figures 2.8 and 2.9). Correlation of methanotroph diversity and methane consumption Simple linear regression revealed a strong positive correlation between summer (June-August) rates of methane consumption and pmoA OTUs (r2=0.62, p<0.001) (Figure 2.6). If pmoA OTUs are regressed against rates of methane consumption from March- December a nearly identical relationship is found (r2=0.64, p<0.001). The same strong positive correlation is also exhibited when, in addition to the pmoA OTUs, tRF LP OTUs 40 that could not be identified (Methods) are included and correlated with summer (r2=0.43, p=0.008), and March-December methane consumption rates (r2=0.48, p=0.004). We tested whether the positive relationship between methane consumption and methanotroph diversity at the KBS LTER was more likely to be explained by either the complementarity or selection hypotheses by determining the trajectory of recovery of richness and consumption in successional soils (Early HT and Mid HT) after intensive row-crop agricultural management (Figure 2.7). There was a linear trajectory for the recovery of both methane consumption (r2=0.69, p < 0.001) and methanotroph richness (r2=0.99, p = 0.020) over time. Extrapolation of the trajectories reveals that if methane consumption and methanotroph diversity continue at the same rates since cessation of agriculture, both would return to the current level of their equilibrium community, the late successional deciduous forest, in approximately 75 years. Total Bacterial Diversity Estimated average total bacterial richness (Chao I) was determined fi'om 454 168 tag sequencing fi'om December 2006 soil samples, and found to be 11,105 in Ag HT, and 7,762 in Late DF . We assumed that every bacterium would contribute to carbon dioxide emissions, and preformed a linear regression with total bacterial richness and average March-December carbon dioxide efflux from 2005-2007. No relationship was found between the two measures (r2=0.22, p=0.522) (Table 2.3). Despite the limited power in the linear regression due to the limited number of samples (n=4), using the same number of samples from comparable methanotroph richness measurements yields a dramatically different result (Table 2.3). Using pmoA clone libraries from Ag HT and Late DF June 41 2006 soil samples (Table 2.2), and average methane consumption measurements from the same 2005-2007 dates, we found a significant positive linear relationship between average March-December methane consumption measures and methanotroph diversity (r2 =0.96, p = 0.013). Multiple Regression Multiple regression with the Ag HT and Late DF gas fluxes from 2005-2007 against soil moisture and maximum air temperature from the same dates and replicates discerned the ability of these general microbial metabolic regulators to explain the observed fluxes. Moisture and temperature were found to be able to explain portions of both fluxes, but were 10 times more effective in explaining the efflux of carbon dioxide (r2 =0.37, p < 0.001) then the rates of methane consumption (r2 =0.03, p = 0.005) (Table 2.3). Discussion Microbial Diversity and the Flux of Greenhouse Gases Correlating ecosystem process rates to their respective microbial diversity measures across the KBS LTER land-use gradient conformed to our hypothesis: methane consumption rose as methanotroph diversity increased (Figure 2.6, Table 2.3), and increases in carbon dioxide emission were not correlated to increases in bacterial richness (Table 2.3). The lowest flux of both greenhouse gases were found in Ag HT as both carbon dioxide and methane consumption fluxes were diminished by row-crop agriculture (Figures 2.1 and 2.2). The difference in the correlations was instead due to the effect of land use on microbial diversity: Row-crop agriculture was associated with a 42 dramatic decrease in methanotroph richness, but total bacterial richness was greater in Ag HT then it was in Late DF (Figure 2.6, Tables 2.2 and 2.3). As postulated by our hypothesis, the ability of microbial diversity to explain the rate of the gas flux is likely the result of the functional redundancy of the microbial community mediating the ecosystem process. Thousands of bacterial species contribute to the efflux of carbon dioxide fi'om soil — reducing the likelihood that diversity influences the process rate. There are comparatively few methanotroph species, which likely results in little or no functional redundancy, as methanotrophs probably occupy non-overlapping niches and every methanotroph species contributes to the rate of methane consumption. Evidence for methanotrophs occupying separate niches at KBS LTER was gained by testing to see if the positive relationship between net methane consumption and methanotroph diversity could best be explained by either the complementarity or selection hypotheses. Ideally this determination would be able to be made with measures of species-specific methane oxidation rates and the reconstruction of defined communities, but the absence of cultured methanotrophs representative of the clades found in KBS LTER soils preclude such measures. As has been found in other soils, the methanotroph community consuming atmospheric methane at KBS-LTER is composed of uncultured methanotrophs. Our findings further indicate that these phylogenetic clusters are those that are largely responsible for the consumption of atmospheric methane (10, 19). No clones were found from the Type I or Type H methanotrophs that are well represented in culture (Figures 2.4 and 2.5). Rather, pmoA clones belonging to 6 clades that have been identified previously in other culture independent investigations of 43 upland soils (10, 18, 19, 45), and 1 new clade, identified as KBS], were found in KBS LTER soils. Of the 7 clades that were recovered, only Cluster I methanotrophs, presumably OC-proteobacteria, have been reported to be cultured, but they remain poorly characterized (46). Despite the absence of cultured methanotrophs, we detemiined whether the complementarity or selection hypotheses could best explain the relationship between methane consumption and methanotroph richness by plotting the recovery of both measures in soils that had been abandoned from agriculture for various lengths of time. The trajectory of the recovery of both methane consumption and methanotroph richness over time is linear, with no indication of abrupt step-wise increases in methane consumption that would be expected to accompany the establishment of a particularly productive methanotroph species in accordance with the selection hypothesis (Figure 2.7). Instead, the concurrent and incremental recovery of methane oxidation and methanotroph diversity following the abandonment fi'om row-crop agriculture is consistent with complementary roles of methanotroph species, and suggests that every methanotroph OTU is important to rates of methane consumption. Additional support for complementarity in the methanotroph community at KBS LTER is found in the ability for general abiotic microbial metabolic regulators, moisture and temperature, to explain 10 times more of the variation in C02 flux then it did for methane consumption (r2=0.37 and 0.03, respectively; Table 2.3). Without an influence of diversity on rates of soil respiration, the ability of moisture and temperature to exert controls on the efflux of carbon dioxide was more apparent. In contrast, with diversity explaining most of the rate of methane consumption, moisture and temperature could 44 only explain a minimal amount of the variation in the rate of methane consumption (Table 2.3). Other studies have similarly failed to find a correlation between total microbial diversity (24, 47, 48) and soil respiration. These studies have also found that changes in the microbial community were correlated to more metabolically specialized processes like N20 production (24) and nitrification potential (24, 48). Studies that have not directly measured total microbial diversity have also found support for the microbial community not influencing carbon dioxide production, but influencing specialized microbially mediated processes ((49), reviewed in (23, 25)). However, not all studies have found that diversity always influences specialized physiological processes. Enwall et al. (50) found rates of denitrification to be independent of microbial diversity, and Wertz et al. (47) found similar results for denitrification and nitrification. There is also evidence of microbial diversity being correlated to soil respiration (8, 13, 22). Additionally, Bell et a1. (51) found bacterial respiration to be correlated with increasing bacterial diversity in laboratory microcosms, but with a maximum of 72 bacterial species, their ecosystem was likely not nearly as functionally redundant as the thousands of bacterial species contributing to carbon dioxide production in soil. Methanotroph richness changing with rates of methane consumption has also been observed in other sites, but an overall pattern is unclear. In Mono Lake, the depth with the highest rate of methane oxidation was found to have the greatest richness of methanotrophs (52), and two pine forests had greater methanotroph richness and methane consumption rates than paired pasture soils in New Zealand (16). However, that same study found that a paired shrubland and pasture soil had the same methanotroph 4s richness despite different rates of methane consumption, and other studies have been inconclusive with no clear relationship observed between methanotroph richness and methane consumption (18-20). Therefore, the relationships we have observed between methane consumption and methanotroph diversity is consistent with some previous findings, and the lack of an overall pattern may at least be partially due to methodological differences. Most assessments of the methanotroph community are performed at a broader phylogenetic level, with fewer replicates, and/or compared to less robust measures of the rate of methane consumption. In addition, this is the first study to correlate methanotroph richness to rates of methane consumption. The Impact of Agriculture on Methanotroph Diversity At KBS LTER, there is unambiguous evidence linking the diversity of aerobic methanotrophs to the observed rates of methane consumption, and for agricultural management to diminish methanotroph species richness. The 24 methanotroph species that we recovered in Late DF (Figure 2.5, Table 2.2) is, to our knowledge, the most methanotroph richness recovered from a single soil; other studies have reported between 1-13 methanotroph species (l6, 18, 19, 45). Conversion of soils to row crop agriculture dramatically reduced that richness to a small subset of the methanotroph richness found in Late DF. In addition, the recovery of methanotroph diversity and methane consumption after cessation of agriculture is projected to take approximately 75 years (Figure 2.7). This slow recovery of methanotroph richness and methane consumption at the KBS LTER is consistent with worldwide observations that suggest a recovery period of ca. 100 46 years for methane consumption following the cessation of agriculture (5). The lack of a quicker recovery following the stopping of agricultural management practices, especially those like fertilization that represent a disturbance to the methanotroph community, indicate that many aspects of methanotroph niches are distressed due to row-crop agriculture. It will be important to identify these pivotal variables and their applicability to other sites if we are to manage lands to conserve or restore methanotroph diversity and enhance the capacity of soil to serve as a sink for this potent greenhouse gas. The methanotroph community changes across the successional gradient at KBS LTER mirrors changes in the plant community, and further underscores the notion that methanotroph niches are slowly re-established and colonized following row-crop agriculture (Figures 2.8 and 2.9). There is a discemable pattern of recovery, and an apparent succession of methanotrophs. Similar methanotroph community changes across other successional gradients have also been observed, with patterns of methanotroph diversity along successional gradients of reforested comfields (18) and reclaimed pasture lands (20). In each of these studies, as well as our study, the methanotrophs changed along with the plant community. King and Nanba (53) also found distinct methanotroph communities in volcanic deposits with different plant communities. However, while these observations suggest that plant diversity and/or community composition influences methanotroph diversity, there is not an obvious causal connection between plant and methanotroph diversity. Conclusion 47 In conclusion, our results suggest that every methanotroph OTU is important to methane consumption rates in KBS LTER soils, and both methane consumption and the diversity of methanotrophs decline in response to row crop agriculture. The decline of both methane consumption and methanotroph diversity in row-crOp agricultural soils, and the long time required for recovery of methanotroph diversity suggests that multiple aspects of the methanotrophs habitat are disrupted. It will be important to identify and quantify the effect of these pivotal variables if we are to manage lands to conserve or restore methanotroph diversity, and enhance the capacity of soil to serve as a sink for this potent greenhouse gas. There is no relationship between soil respiration and bacterial richness, and the contrasting result from methane consumption and methanotroph richness is consistent with the prediction that microbial diversity is more likely to be important to a specialized metabolic process whose a microbial commrmity is likely to be of limited richness and consequently functionally redundant. Acknowledgements All methane rate measurements and were taken by G. Philip Robertson and members of his laboratory. Soil sampling and DNA extraction for the assessment of total bacterial diversity was performed by Dion Anotonopoulos. 48 .3828 SE8. 8.88 8 85 £828: 083. 8284—95 88 28 8:8 88280888823— 05 .8 08 £8 =< 3:82 820388 28 82.38% 288088 .8288 82:0 .8288 :898m-Eoo-~8:3 85:; E8688 2 28 .32 2 232288 89c noc¢u§n< 8288 32:58 2 2.8 28:: :89 .88: .83 .8 8.88: :8: ~88“— ooiéma— .8253 22888 9.322%.“ Eat 8.88822 8:: 88 88: 82...: :8 "own 88» co. :8 88:9: 28 .88 soon 88: 822 88232 8:5 2: .8 83. .888 88:28“. 286888 83 am; :_ 8 m :8.:mE.m€_.8~<\H&E 8 28$ up :8 288—8 82m 28222. H: w< 88::otw< 2:288:00 H: bumm— _m=2m88:m Esm— ._.Z 22 82880 28588522 .5.— 22 .88"— 88888522 "5 33 ~88m 80:298. 88% 88 N. _ 48:8 25 E .8232ng 8:m MmEH—fimx .—.N 03:. 49 22 £238— .n0ZV 0850 862me Sam 05 .23 8888.0 083 0888 :8 085 89¢ 80:88 MD“: 50 8 v: 8:88,..- 8 0.3 33.5 058 5 8m 8838:8053 z «8 .8888 “E 838 v 8 3-8 38.6. 2:: N ,8 8 m 8 8.8 $58 2:: _ ,8 8 m 8 8 8m 2:: _ 8 0< N 8 8-8 88 as: _ ,8 m... _ mm 8-8 38 .2888 _ 8 0< w 8 :8 88 2.3 m 8 33 o. a. 3-8 88 2.3 _ 8 33 E 8 :8 8m 2:; _ 8 93 8 3 8 88 as: _ ,8 23 a o: 8 8N 2:; _ 8 33 : S 8-8 88 2:: _ 8 23 v a 8-8 88 .3888 _ 8 23 858.888 . _ 5 8% all; 482 84:88 a .58. $3888 85138.38 88888.8 88% 82.8.8: 08.0 «BER-8 gm .N.N 03:. Table 2.3. Correlations between the flux of greenhouse gases, species richness and environmental conditions. g P Value C02 Production]: vs. bacterial richnessz 0.22 0.522 vs. temp and moisture4 0.37 <0.001 CH4 Consumption]: vs. methanotroph richness3 0.96 0.018 vs. temp and moisture4 0.03 0.005 1 Rate measurements were typically made bimonthly between March and November of 2005 through 2007. 2 Linear regression (n=4) with total bacterial richness estimated with Chao I and based on 220,996 168 tag sequences detemiined from samples collected in December 2006. Rate measures are arithmetic averages fi'om the same plots fiom which richness was determined: The second and fourth replicates of Ag HT, and the first and third replicates of Late DF. 3 Linear regression (n=4) with methanotroph richness determined at a 94% pmoA sequence identity from clone libraries from samples collected in June 2006. Rate measures are arithmetic averages from the same plots from which richness was determined: The first and second replicates of Ag HT, and the first and third replicates of Late DF. 4 Multiple regression (n=280) with maximum air temperature and soil moisture for the day rates were measured. Rate measures are from four replicates of Ag HT, and from three replicates of Late DF. 51 RR —7 T T i‘i: p -2- i\§ i\ \i7!23 v.2... 2 28323. 3:23 N and .325» £3395 «£8383. 33:22.3— »uBm m2... 9. 80352 @225 _ mom 48:3 xfinvam «29.8.9.3— Baegzom beam 3:. mm Eovmfiv. @833 N mom $02.62.? ..72—Roi 4283QO 338.353— xcam m3... 9 Eocwcg @825 _ mom $2525., 2.332685 45.883. 35:853— xvam m2... .3... Eovwfiv— 28:5 N com .383 :20— ..22383- 38:82qu zuBm m2: mm EonwEv. 32:5 _ and 683 .35— ..28383. 368.353— N .2926 2. 322m .23: N ..3. .35 23 33.5 8.33 32222.. 32.3. N .2936 8.. 8.2m .2225 M. as. .3... 23 3.25 822m 323.2... 32.3. N 323.30 N3 835 32.35 N 3.3. H: w... 3.3.. 822m 322.33. m3.2.3. N .236 wNN 322m .223: N. as. .2. w... 3...: 822m 323.22 mm2.3. 3 a a 3% mzlm 3.403% .333... mi. 5 33: 823.5: WEEK 28 82m 05 mo ban—8.5 ._.m 03$. 87 Net Methane Consumption Methanotroph Species (g CH4-C ha"1 day‘) 28- —14 24— -12 20— I -10 16j - 8 T 12- - 6 8 - - 4 4 - — 2 o _ Control Sub-plot Fertilized Sub-plot Control Sub-plot Fertilized Sub-plot Figure 3.1. Differences between Late DF fertilized and control treatments at KBS LTER in average net methane consumption and the average number of methanotroph species (defined as pmoA sequences having 94% average nucleotide sequence similarity). Fertilization has no effect on either measure (t=0.58, p=0.57 for consumption; and t=1.18, p=0.36 for richness). Measures for net methane consumption are averages from 3 replicates of each treatment, while the methanotroph species are averages from 2 replicates of each treatment. Net methane consumption from the same 2 replicates as those with methanotroph species data yields the same result (20.9243 and 19.2233 average net methane consumption of the control and fertilized sub-plots, respectively, and t=0.28 p=0.79). Error bars represent standard error. 88 P Number of Methanotroph Species Number of Methanotroph Species 4o - __ Fertilized Sub—plot 1 35 4 — — - Fertilized Sub-plot 2 """"" Control Sub-plot 1 30 + ' ' - - Control Sub-plot 2 25 _ 20 _ 15 - 1o - __________________________________ 5 .. ,5 1,5". g ’: 2’ —————————————— /" o I I I I T 1 I Number of Sequences 40 _ Broadbalk Wheat 1 35 d ---------- Broadbalk Wheat 2 --------- Broadbalk VWdemess 1 30 d W Broadbalk VWdemecs 2 “WW Knott Wood 1 25 ‘ ------------ ~— Knott Wood 2 20 - Number of Sequences Figure 3.2. (A) Rarefaction curves from pmoA clone libraries constructed from the control and fertilized sub-plots of the late successional forest at KBS LTER. Libraries were constructed from DNA extracted from soil sampled in June 2007. (B) Rarefaction curves from pmoA clone libraries constructed from Rothamsted Reasearch soils sampled in Feburary 2008. All curves were constructed using data fiom neighbor joining matrixes from Arb (Ludwig et al. 2004), with curves calculated by DOTUR (Schloss and Handelsman 2005). Methanotroph species are defined as pmoA sequences having 94% average nucleotide sequence similarity. Error bars representing 95% confidence intervals were omitted for the sake of clarity. 89 a—proteobacteria Upland Soil . I Clustera ’N l Type II methanotrophs y-proteobacteria { amoA y-proteobacten'a Type I Verrucomicrobia methanotrophs} pmo A K Upland Soi Clustery MR1. I f O I Cluster ll :RA21 Unknown Classification Unknown Classification M90-P96 \ O KBS1 B-proteobacteria 4 GMOA Cmnothn‘x polyspora) y-proteobacteria Presumably 0.10 Crenarachaeota Clusterl . '- a—proteobacteria _ amoA Figure 3.3. Phylogenetic tree of selected partial pmoA and amoA protein sequences fi'om public databases and translated from PCR-based clone libraries fiom late successional deciduous forest sub-plots. The tree is based on 164 amino acid positions using Phylip Protein Maximum Likelihood as implemented in ARB (26). Bolded clades were recovered from the late successional forest sub-plots. The symbols adjacent to the clade reflect the clade’s recovery fiom either the control treatment (0 ) or the fertilized treatment (I). The scale bar represents 10 PAM units. 90 Control Subplot 1 Control Subplot 2 Fertilized Subplot1 Fertilized Subplot 2 0.05 B. __{——Late DF Rep 1, June '05,60°C-51°C Control Subplot 1 Late DF Rep 1, June '05, 62°C-56°C Control Subplot 2 Fertilized Subplot 1 Late DF Rep 1, June '06, 60°051°C Late DF Rep 3, June '06, 60°C-51°C Fertilized Subplot 2 —- l LateDFRep1,June'O5,51°C ' LateDFRep1,June'05,48°C !" Ag HT Rep 1, June '06, 60°C-51°C l——- Ag HT Rep 2, June '06, 60°C-51°C Lae DF Rep 1, Deoerrber'04,62°C-56°C Ag HT Rep 1, June'05,48°C I Ag HT Rep 1, June '05,62°C-56°C 1 AgHTRep1,oewrper'o-1,62°056°c l——l 0.05 Figure 3.4. (A) Similarity of methanotroph communities among fertilized and control sub-plot replicates. (B) Similarity of methanotroph communities between fertilized and control sub-plot replicates with all other pmoA libraries from KBS LTER Expect for the fertilized and control sub-plots, the pmoA clone libraries were first reported in Chapter 2, and details of their construction can be found there. Both dendograms are based on Serenson index calculations for each pairwise comparison of the methanotroph communities using pmoA species (defined as pmoA sequences having 94% average nucleotide sequence similarity), and then clustered using neighbor-joining with MEGA (3 5). The scale bars represent at 0.05 change in the Sorenson index. 91 Figure 3.5. Neighbor joining phylogenetic tree of the partial nucleic acid sequences of pmoA and amoA from reference sequences, and KBS-LTER clone libraries from the control and fertilized sub-plots soil sampled in June 2007. The tree is rooted with the branching from the amoA sequences fiom Nitrosomonas eutropha, Nitrosococcus mobilis, and Nitrosovibrio tenuis. Nodes representing the 17 different methanotroph species (defined as pmoA sequences having 94% average nucleotide sequence similarity) found in late successional deciduous forest sub-plots are circled. The symbols in the circles reflect species’ recovery from either the control treatment (0) or the fertilized treatment (I ). 92 %W——l:l) Crenothrix polyspora —I:7) B-proteobacteria amoA )Type I methanotrophs y— proteobacteria Nitrosococcus oceani strain AFCZ4P amoA (AF509002) Nitrosococcus oceani strain SW amoA (AF509003) ) Y— —pr0teObaCteria amOA )T ell methanotro hs, oc— roteobacteria Uncultured bacterium MR4 (AF200727) Fertilized sub—plot 2, 0708 CO7 Fertilized sub-plot 1, Fert BOS - Control sub— —plot 2, 0708 G04 . I Upland SOII CIUSter 0‘4 Control sub- -plot 2, Control H02 Uncultured bacterium Hold :1 (AF148523) Control sub— —plot 1 1CDF- Uncultured bacterium DGGE band5 E5FB- f (AJ579663) Uncultured bacterium clone CL10 (A J699 O7) Uncultured bacterium DGGE band CF- h (3AJ868282) cControl sub- plot 1,1—CDF 1 CO 02 . gontr ol sub- plot 2, 0708 E05 Control sub- -plot 2, 0708 G06 . Cluster |l Control sub- lot 1, 1CDF B10 0 j Fertilized su plot 2 0708 COB Fertilized su -plot 2, 0808 H02 I] Control sub- -p|ot1, 1CDF- 1 804 Control sub- plot 1, C1DF- 1 805 —E%%t‘i%.it%%& 1’ 18% 19302 0 j)KBs1 M90 P69 { Uncultured bacterium DGGEb and E56F- -a (A J579662) Uncultured bacterium MR1 (AF200729) Control sub- plot 1,1CDF-1 A08 Control sub-plot 2, Control C12 . I MR1 Fertilized sub plot 1,0708 A02 Fertilized sub— plot 1,0708 A01 Uncultured bacterium amo/pmoPN_ O- 5#2 (EF452672) )RA21 )Verrucomicorbia pmoA Uncultured bacterium clone B7233- 79 (AJ()564439)313 v Uncultured bacterium clone pmoA— U2- D0295 Fertilized sub- p—Iot 2, 0708 C03 I Fertilized sub- plot 2 0708 B03 Uncultured bacterium DGGE6 band E33b- -a (AJ579661) Control s-pub ot,1 1CD FD O I Fertilized subl— —ptlot’1, 13x1 E02 Fertilized sub- plot 1 1208 A02 Control sub- plot 1 1CDF— 18 06 Control sub- plot 1,1CDF CO4 L _ Methanotroph K3 16 (AF547178) —- Uncultured bacterium clone pLWPmoA- -24 (DQ067064) Uncultured bacterium clone LOPA12 4 (AF358041) Uncultured bacterium clone CL27 (AJ699310) Fertilized sub-plot 1, Fert 807 Control sub- plot 2, Control E07 . I Fertilized sub- -p|ot 1 13x1 A04 Control sub- —p|ot2, Control 805 Cluster I Uncultured bacterium clone CL49 (AJ699315) Fertilized sub— plot 1 1208 (302 ! Methanotroph K1 -8 (AF547175) Methanotroph K2- 14 (AF54718707) Methanotr orph K3— 21 (AF 547 Uncultured bacterium clone pmoAa4 (D0367742) Uncultured bacterium c|c[>3n19 CL5 (AJ699316) Control sub- -p|ot1, 1CDF . I Fertilized sub- plot 2, Fert E04 Control sub- p—Iot 2, 0708 F02 Uncultured bacterium DGGE band L58-b2 (AJ868244) Fertilized sub- plot 2, 0708 D07 .1 Fertilized sub- plot 2 0708 D06 Control sub- plot 1 tCDF- 1 CO7 Fertilized sub- -plot 2, Fert F05 . I Fertilized sub- plot 2, 0708 COB Control sub- -p|ot 2, Control E01 Control sub- plot 1, iCDF B12 . J Control sub-plot 1, 1CDF Bot Figure 3.5 0.10 93 14 7 .Non-fertilized Soils I Fertilized Soils 12 4 10-i Methanotroph Species Figure 3.6. Methanotroph richness in paired sites featuring long-term fertilized (black shading) and non-fertilized soils (grey shading). The three sites to the left compare fertilized sites in the context of agricultural management, while the two sites on the right are fertilized forests whose only land management is fertilization. Error bars represent standard errors as the averages are reported for sites that featured replicated sites. Additional site details are provided in Table 3.1. 94 cat-proteobacteria Upland Soil JR1 Cluster a Type II / methanotrophs y-proteobaoteria amoA f Type I r- methanotrophs Verrucomicrobia - _ \ pmoA y—proteobactena Upland Soil Cluster y 1 Wood I 1Wildemess .5 r 0 “we“ MR1 Cluster II L, 1 Wood RA21 4 Wheat - - M90-P96 Classification Unknown 0 Wood Classification KBS1 OWilderness 0 Wood 1 Wheat 0.1 0 Wilderness — 1 Wheat p—proteobaoten‘a amoA ‘ Crenothrix polyspora) y-proteobacteria ‘ Presumably Crenaracheeota Cluster I a—proteobacteria amoA 3 Wood 4 Wildemess 7 Wheat Figure 3.7. Phylogenetic tree of selected partial pmoA and amoA protein sequences fi'om public databases and translated from PCR-based clone libraries fi'om late successional deciduous forest sub-plots. The tree is based on 164 amino acid positions using Phylip Protein Maximum Likelihood as implemented in ARB (26). Bolded clades were recovered from Rothamsted Research soils. Beneath each bolded clade is a listing of the number of species within that clade found in which Rothamsted Research treatment: Knott Wood (Wood), Broadbalk Wilderness (Wilderness) and Broadbalk Wheat (Wheat). The scale bar represents 10 PAM units. 95 0.241 0.18 - 0.12- Coord‘nate 2 O ' Sakerat Experimental -0.06 . Station -0. 12 - 127 -0.18 - -0.24 - -0 3 30.48 oi -o.é2 -o.é4 -o.1'6 ode d 0.68 0.16 0.524 Coord'nate1 B. Broadbalk Wildemess 2 HE m w... 2 Knott Wood 1 H— Broadbalk Wildemess 1 Broadbalk Wheat 11:} Broadbalk Wheat 2 o Fertilized Sub-plot 2i} \‘ Rothamsted Research Fertilized Sub-plot 11? Late DF 3 KBS LTER Control Sub-plot 2 Control Sub-plot 1 Late DF 1 Evergreen Forest . ___r—i:—Reforested Plantation ) 322%? ”mm” Comfieldfi Pine Forest, Control —‘i:—r_l=ine Forest, Fertilizedfi ) Harvard Forest l—-t Hardwood Forest, Fertilizedtf} 0.1 Figure 3.8. Similarity of methanotroph communities in paired sites featuring long-term fertilized and non-fertilized soils. (A) The Sarenson index was calculated for each pairwise comparison of methanotroph species using two-dimensional non-metric multidimensional scaling (21). (B) The same data as in (A), but clustered using neighbor- joining with MEGA (35) and displayed as a dendogram. Labels with a 7:? are sites that have been fertilized long-term. Additional site information is provided in Table 3.1. 96 References l. 10. 11. 12. Smith KA, et al. (2000) Oxidation of atmospheric methane in Northern European soils, comparison with other ecosystems, and uncertainties in the global terrestrial sink. Global Change Biology 6(7):791-803. IPCC (2007) Climate Change 2007: Synthesis Report. Contribution of Working Groups I, II and IH to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. eds Team CW, Pachauri RK, & Reisinger AGeneva, Switzerland). Rigby M, et al. (2008) Renewed Growth of Atmospheric Methane. Geophysical Research Letters 35:doi:10.1029/2008GLO36037. Hutsch BW (2001) Methane oxidation in non-flooded soils as affected by crop production - invited paper. European Journal of Agronomy 14( 4):237- 260. King GM (1997) Responses of atmospheric methane consumption by soils to global climate change. Global Change Biology 3(4):351-362. Suwanwaree P & Robertson GP (2005) Methane Oxidation in Forest, Successional, and No-till Agricultural Ecosystems: Effects of Nitrogen and Soil Disturbance. Soil Science of America Journal 69: 1722-1779. Gulledge J, Hrywna Y, Cavanaugh C, & Steudler PA (2004) Effects of long-term nitrogen fertilization on the uptake kinetics of atmospheric methane in temperate forest soils. FEMS Microbiology Ecology 49(3):389-400. Chan ASK, Steudler PA, Bowden RD, Gulledge J, & Cavanaugh CM (2005) Consequences of nitrogen fertilization on soil methane consumption in a productive temperate deciduous forest. Biology and Fertility of Soils 41(3):182- 1 89. Conrad R (1996) Soil microorganisms as controllers of atmospheric trace gases (HZ, CO, CH4, OCS, N20, and NO). Microbial. Rev. 60(4):609-640. Gulledge J & Schimel JP (1998) Low-Concentration Kinetics of Atmospheric CH4 Oxidation in Soil and Mechanism of NH4+ Inhibition. Appl. Environ. Microbial. 64(11):429l-4298. Lau E, Ahmad A, Steudler PA, & Cavanaugh CM (2007) Molecular characterization of methanotrophic communities in forest soils. FEMS Microbial Ecol 60(3):490-500. Maxfield PJ, Hornibrook ERC, & Evershed RP (2008) Acute impact of agriculture on high-affinity methanotrophic bacterial populations. Environmental Microbiology 10(7): 191 7-1924. 97 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. Kolb S, Knief C, Dunfield PF, & Conrad R (2005) Abundance and activity of uncultured methanotrophic bacteria involved in the consumption of atmospheric methane in two forest soils. Environmental Microbiology 7(8):1150-1161. Carini SA, Orcutt BN, & Joye SB (2003) Interactions between methane oxidation and nitrification in coastal sediments. Geamicrabialagy Journal 20(4):355-3 74. Knief C, et al. (2005) Diversity of methanotrophic bacteria in tropical upland soils under different land uses. Applied and Environmental Microbiology 71(7):3826— 3 83 l . Ambus P & Robertson GP (2006) The effect of increased 11 deposition on nitrous oxide, methane and carbon dioxide fluxes from unmanaged forest and grassland communities in Michigan. Biogeochemistry 79(3):315-337. Willison TW, Oflaherty MS, Tlustos P, Goulding KWT, & Powlson DS (1997) Variations in microbial populations in soils with different methane uptake rates. pp 85-90. Hutsch BW, Webster CP, & Powlson DS (1994) METHANE OXIDATION IN SOIL AS AFFECTED BY LAND-USE, SOIL-PH AND N-FERTILIZATION. Soil Biology & Biochemistry 26(12):]613-1622. Robertson GP, Paul EA, & Harwood RR (2000) Greenhouse Gases in Intensive Agriculture: Contributions of Individual Gases to the Radiative Forcing of the Atmosphere. Science 289:1922—1925. Millar N & Kahmark K (2008) Field Gas Sampling Protocol (Bucket Method). (W. K. Kellogg Biological Station Long-Tenn Ecological Research Project, Michigan State University, Hickory Comers, Nfl). Hammer 0, Harper DAT, & Ryan PD (2001) PAST: Paleontological statistics software package for education and data analysis. Palaeontalagia Electronica 4(1):l-9. - Costello AM & Lidstrom ME (1999) Molecular Characterization of Functional and Phylogenetic Genes from Natural Populations of Methanotrophs in Lake Sediments. Appl. Environ. Microbial. 65(11):5066-5074. Bourne DG, McDonald IR, & Murrell J C (2001) Comparison of pmoA PCR primer sets as tools for investigating methanotroph diversity in three Danish soils. Applied and Environmental Microbiology 67(9):3802-3 809. Stevenson BS, Eichorst SA, Wertz JT, Schmidt TM, & Breznak JA (2004) New Strategies for Cultivation and Detection of Previously Uncultured Microbes. Appl. Environ. Microbiol. 70(8):4748-4755. 98 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. Altschul S, et al. (1997) Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res 25:3389 - 3402. Ludwig W, et al. (2004) ARB: a software environment for sequence data. Nucl. Acids Res. 32(4):]363-1371. Konstantinidis KT & Tiedje JM (2004) Trends between gene content and genome size in prokaryotic species with larger genomes. Proceedings of the National Academy of Sciences of the United States of America 101(9):3160-3165. Schloss PD & Handelsman J (2005) Introducing DOTUR, a Computer Program for Defining Operational Taxonomic Units and Estimating Species Richness. Appl. Environ. Microbiol. 71(3):]501-1506. Castro MS, Steudler PA, Melillo JM, Aber JD, & Bowden RD (1995) Factors Controlling Atmospheric Methane Consumption by Temperate Forest Soils. Global Biageachemical Cycles 9:1-10. Gulledge J, Doyle AP, & Schimel JP (1997) Different NH4+-inhibition patterns of soil CH4 consumption: A result of distinct CH4-oxidizer populations across sites? Soil Biology and Biochemistry 29(1): 1 3-21. Nyerges G & Stein LY (2009) Ammonia cometabolism and product inhibition vary considerably among species of methanotrophic bacteria. FEMS Microbiology Letters 297(1): 1 3 1-136. Aber JD, Alison M, Boone R, Melillo JM, & Steudler P (1993) Plant and Soil Responses to Chronic Nitrogen Additions at the Harvard Forest, Massachusetts. Ecological Applications 3(1):156 - 166. Compton JE, Watrud LS, Porteous LA, & DeGrood S (2004) Response of soil microbial biomass and community composition to chronic nitrogen additions at Harvard forest. F arest Ecology and Management 196(1): 143-158. Aber J, et al. (1998) Nitrogen Saturation in Temperate Forest Ecosystems. BioScience 48(11):921-934. Kumar S, Nei M, Dudley J, & Tamura K (2008) MEGA: A biologist-centric software for evolutionary analysis of DNA and protein sequences. Brief Biainform 9(4):299-306. 99 Chapter 4 Conclusions and Future directions In this thesis there are multiple lines of evidence suggesting that multiple aspects of the methanotrophs’ habitat are disrupted by row-crop agriculture at KBS LTER: (a) The decline of both methane consumption and methanotroph diversity in row-crop agricultural soils; (b) the 75 years required after the cessation of agricultural land management for methane consumption rates and methanotroph diversity at KBS LTER to achieve the current rate of methane consumption and diversity of the native soils (Chapters 2 and 3). An array of row-crop agriculture related factors, in addition to, and perhaps in interaction with long-fertilization that was investigated in Chapter 3, may all be disturbing the methanotroph community and are discussed below. Together, the net effect of row-crop agriculture at KBS LTER is the apparent destruction of methanotroph niches and subsequent loss of methanotroph richness and rates of methane consumption. Each of the row-crop agriculture related factors could be explored in the future, as our ability to potentially manage lands to conserve or restore methanotroph diversity, and enhance the capacity of the methane soil sink, will rely on understanding the effect of these potentially pivotal variables on the methanotroph community. Only 1 of the pmoA phylogenetic clusters found at KBS LTER has not been reported in other upland soils (Chapter 2), and we recovered that cluster in the Broadbalk wheat soil at Rothamsted Research (Chapter 3). Therefore, while each sites’ methanotroph community composition is unique (Chapter 3, Appendix C), at least some of the methanotroph species are going to be closely related to the methanotrophs that we find at KBS LTER. The discovery of the effects of land management practices to specific methanotroph 100 species would likely be applicable to other sites where the same or closely related species are present. Previous studies have found that the common row-crop agricultural practices of tillage, fertilizer, pesticides, and herbicide application can negatively impact methane consumption (reviewed in (1)). These practices are all featured in KBS LTER’s conventional row-crop agricultural treatment, and therefore may at least be partially causing the 7-fold land-use related decrease in methane consumption and methanotroph diversity observed in Ag HT (Chapter 2). In addition, land use related changes to soil properties like pH, water filled pores space, and dry bulk density may be contributing to the decline of methane consumption and the methanotroph community in Ag HT. Each has been shown to directly influence rates of methane consumption (reviewed by (1 -3)), and changes in the methanotroph community have been found to be associated with different pHs (4, 5), changes in the successional stage of the plant community (Chapter 2 and (5, 6)), temperatures, and precipitation levels (7). None of these factors were investigated in the present study as we chose to only investigate long-term fertilization. The short-term effect of tillage on methane consumption at KBS LTER was negligible (8), and led to the assumption that its long-term effect would be minor. However, reduced tillage has been shown increase methane consumption (reviewed by (1), and the disturbance to soil structure as a result of tillage is potentially a contributor to the 75-100 years it takes to for methane consumption rates and methanotroph richness to recover fi'om agricultural land use (Chapter 2 and (3)), and with long-term fertilization’s minor effects, tillage’s long-term effect on methanotroph communities merits further investigation. 101 For instance, tillage may be changing the soil structure such that dry bulk density and water filled pore spaces change and cause less methane to be available to the methanotroph community. Increases in both dry bulk density and water filled pore spaces limit gas diffusion, and have been correlated to decreases in rates of methane consumption (2, 3). Dry bulk density at KBS LTER has only occasionally been measured previously at KBS LTER (http://lter.kbs.msu.edu), and therefore we cannot determine if either dry bulk density or water filled pore spaces (the determination of water filled pore spaces depends on the measure of dry bulk density) are affecting methane consumption. Pesticide and herbicide application, other agricultural practices at KBS-LTER, have also been shown to negatively impact the rate of methane consumption at other sites (1, 9-11), and can alter the methano/methylotroph community (9). The magnitude of the chemical impact on methane consumption and the methanotroph community can greatly vary depending upon chemical and soil type (1, 9, 10, 12). For instance, long term application of atrazine and metolachlor, two herbicides among those used at KBS-LTER, has been found to not cause a difference in the rate of methane consumption, and to only cause minor changes in the composition of the methano/methylotroph community (10). Also, a study contrasting the effect herbicides and fertilization found that the methano/methylotroph community clustered according to the type of fertilization, and that methane consumption rates did not significantly decline due to herbicide treatment (12). The conventional agricultural soil at KBS LTER has been treated with a variety of chemicals (http://lter.kbs.msu.edu), so a definitive determination of herbicide or pesticide effects on the KBS LTER methanotroph community will be difficult. In 102 addition, the organic agricultural soil at KBS LTER despite having no herbicide or pesticide application since 1989 still has a low rates of methane oxidation (13); indicating that the influence of herbicide and pesticide application on rate and the methanotroph community is likely to minimal. However, herbicide or pesticide application may nevertheless be influencing the loss of methanotroph richness at KBS LTER. Decreasing pH has also been correlated to reduced methane consumption (reviewed by (3 )), but the reverse trend has been observed at KBS LTER. Neither rates of methane consumption nor methanotroph richness decline with pH at KBS LTER. The Late DF soils have the lowest pH (approximately 5.3, http://lter.kbs.msu.edu), but have the most methanotroph richness and the greatest rates of methane consumption, while the Ag HT soils have the highest pH (approximately 6.2) with the least methanotroph richness and the lowest rates of methane consumption. Notably, while declining pH is not affecting methanotroph richness at KBS LTER, it may be limiting methanotmph richness in the Rothamsted Research soils. There, Knott Wood has the least methanotroph richness, and the lowest pH (4.0) while Broadbalk wheat has the highest pH (6.7) as well as the most methanotroph richness (14, 15). Future directions into determining the cause behind the decrease in methanotroph diversity and methane consumption associated with row-crop agriculture at KBS LTER might best be focused upon tillage, changes in soil structure, changes associated with the plant community (discussed in Chapter 2), and the possible interaction of these factors with fertilization and pesticide and herbicide application. Each of these directions would shed light on the niche for KBS LTER methanotrophs, and lead to insights that could 103 predict and explain the response of methanotrophs and, most importantly, rate of methane consumption to management changes at KBS LTER. 104 References 1. 10. 11. Hutsch BW (2001) Methane oxidation in non-flooded soils as affected by crop production - invited paper. European Journal of Agronomy 14( 4):237- 260. King GM (1997) Responses of atmospheric methane consumption by soils to global climate change. Global Change Biology 3(4):351-362. Smith KA, et al. (2000) Oxidation of atmospheric methane in Northern European soils, comparison with other ecosystems, and uncertainties in the global terrestrial sink. Global Change Biology 6(7):791-803. Knief C, Lipski A, & Dunfield PF (2003) Diversity and activity of methanotrophic bacteria in different upland soils. Applied and Environmental Microbiology 69(11):6703-6714. Knief C, et al. (2005) Diversity of methanotrophic bacteria in tropical upland soils under different land uses. Applied and Environmental Microbiology 71(7):3826- 383 1 . Zhou XQ, et al. (2008) Effects of grazing by sheep on the structure of methane- oxidizing bacterial community of steppe soil. Soil Biology & Biochemistry 40(1):258-261. Horz H-P, Rich V, Avrahami S, & Bohannan BJM (2005) Methane-Oxidizing Bacteria in a California Upland Grassland Soil: Diversity and Response to Simulated Global Change. Appl. Environ. Microbiol. 71(5):2642-2652. Suwanwaree P & Robertson GP (2005) Methane Oxidation in Forest, Successional, and No-till Agricultural Ecosystems: Effects of Nitrogen and Soil Disturbance. Soil Science of America Journal 69:1722-1779. Mertens B, Boon N, & Verstraete W (2005) Stereospecific effect of hexachlorocyclohexane on activity and structure of soil methanotrophic communities. Environmental Microbiology 7(5):660-669. Seghers D, et al. (2003) Effect of long-term herbicide applications on the bacterial community structure and function in an agricultural soil. FEMS Microbiology Ecology 46(2): 139-146. Seghers D, et al. (2003) Pollution induced community tolerance (PICT) and analysis of 16S rRNA genes to evaluate the long-term effects of herbicides on methanotrophic communities in soil. European Journal of Soil Science 54(4):679- 684. 105 12. 13. 14. 15. Seghers D, Siciliano SD, Top EM, & Verstraete W (2005) Combined effect of fertilizer and herbicide applications on the abundance, community structure and performance of the soil methanotrophic community. Soil Biology & Biochemistry 37(2):l87-193. Robertson GP, Paul EA, & Harwood RR (2000) Greenhouse Gases in Intensive Agriculture: Contributions of Individual Gases to the Radiative Forcing of the Atmosphere. Science 289:1922-1925. Goulding KWT, Willison TW, Webster CP, & Powlson DS (1996) Methane fluxes in aerobic soils. Environmental Monitoring and Assessment 42(1): 175-187. Hutsch BW, Webster CP, & Powlson DS (1994) METHANE OXIDATION IN SOIL AS AFFECTED BY LAND-USE, SOIL-PH AND N-FERTILIZATTON. Soil Biology & Biochemistry 26(12): 1 61 3-1622. 106 Appendix A Assessment of PCR bias Introduction To determine if relative abundance, in addition to richness, could be used in determining and comparing the methanotroph community structure, the bias from the polymerase chain reaction (PCR) that amplified pmoA in the tRF LP analysis and some of the clone libraries was assessed. PCR bias is the over-amplification of specific templates which results in the post-amplification concentrations of those templates being much greater then their pre-amplification concentrations (1, 2). If there is PCR bias then the abundance measures fiom PCR based community analyses are misleading, erroneous conclusions could be made regarding the dominant and rare organisms in a given community, and measures (i.e. diversity indices) that rely on relative abundance will be skewed. The methanotroph community analyses in this thesis (Chapters 2—3) utilize the A189-A682 primer pair (3) to amplify pmoA. Any PCR bias associated with the primer pair has not been quantified, nor has the consistency of the output from the primers and the specific PCR conditions been addressed previously. Thus, PCR bias was assessed with tRF LP profiles whose initial templates were defined artificial communities of pmoA and amoA (The A subunit of the ammonia monooxygenase (amoA) is also amplified by A189-A682) that varied the concentration of initial template. Ideally, if there is PCR bias it will be consistent throughout the mixtures so that despite the bias the relative abundances measures could be included in the community analyses of the pmoA 107 community. While not suitable for quantitative comparisons, the measures would be the same regardless of concentration changes within the mix, justifying their usage. Methods The template for amplification reactions was 4 different mixes of purified plasmids, and each of the purified plasmids. The plasmids contained pmoA PCR products fiom soils at the Kellogg Biological Station Long Term Ecological Research site TOPO TA cloned into vector pCR 4 or pCR 2.1 (Invitrogen, Carlsbad, CA) (Chapter 2). The plasmids represented 6 pmoA species and 1 species of the A subunit of the ammonia monooxygenase (amoA), which A189-A682 also amplifies. The plasmids were mixed such that the concentrations of all the species were held approximately constant except for one pmoA species, Cluster I A (Table 1). Each mix was run as a tRF LP either 2 or 3 times with each tRF LP beginning with new PCR amplifications. The PCR reactions contained a total of 30 pg of template, 1.25 pl 1% BSA, 10 mM dNTPs, 0.2 pM of each primer (A682 was labeled with fluorophore 6- carboxyfluorescein (6-Fam)), 2.5 pl 10x PCR buffer (200mM Tris-HCl (pH 8.4) and 500 mM KCl), 0.5 pl 50 mM MgClz and 1.5 U of Taq DNA polymerase (Invitrogen, Carlsbad, CA) in a total volume of 25 pl. Reaction conditions were 95 °C for 5 minutes, 15 cycles of a ‘touchdown PCR’ of 95°C for 1 minute, 62°C for 1 minute (-O.4°C each cycle to 56°C), and 72°C for 1 minute, 15 cycles using 56°C as the annealing temperature, and a final 10 minute extension at 72°C. For each tRF LP, 4 replicate PCR reactions were pooled. PCR products were purified using a MinElute column (Qiagen, Valencia, CA). In a 50 pl reaction, 300ng of 108 purified product was digested with 1.5 U Tau I (F ermentas, Glen Burnie, MD) by incubating at 55°C for 1 hr 30 min. To inactivate the enzyme the DNA was precipitated as follows: The sample was diluted to 500 pl, followed by the addition of 50 pl 3M sodium acetate, 1 pl or 2.5 pl 10 mg/ml glycogen, and 500 pl isopropanol, and holding on ice for at least 5 minutes. The DNA was then pelleted by centrifugation at 16,000 x g for either 5 (1 pl glycogen) or 10 minutes (2.5 pl glycogen). The supernatant was decantated, and the pellet washed with 500p] of 80% ethanol followed by centrifugation for 2 minutes at 16,000 x g, and removal of the supernatant. After a 30 second centrifugation additional ethanol was removed, and the DNA was air dried for 5-10 minutes before resuspension in 20 pl of water. In an 18-22 pl reaction 140 ng of DNA was digested with 2.5 U SspI (Neb, Ipswich, MA) in a 1hr 30min incubation at 37°C. After heat inactivation at 65°C for 20 minutes 6 U of BstUI (Neb, Ipswich, MA) was added, incubated at 60°C for 1hr 30min, and inactivated by adding 0.8 pl of 0.5 mM EDTA. Capillary electrophoresis of the tRF LP reactions was then performed with a 5 fu cutoff at RTSF. Individual peaks were distinguished fiom the background signal and binned using TRFLP-Stats (4). In TRFLP-Stats default settings were used except for the standard deviation cutoff, which was increased to 4.5. The resulting cutoff was approximately 25 fu. The tRF LP profiles of the individual plasmids ensured that the specific bin belonging to each species was correctly identified. Any other peaks were excluded from the analysis, and the relative abundance of the remaining peaks were normalized such that the absolute abundance of every treatment was 100. The amount of relative abundance 109 excluded from any one tRF LP profile before normalization was between 0-11% with all tRFLPs except one having <6% excluded. Results The pmoA species Cluster I B was consistently over-amplified in every mix, while species Cluster I A was over-amplified except for when its initial abundance was less then 10% of the initial artificial community (Table 1). The over-amplification of Cluster I A and Cluster I B species was dramatic, and in some mixes was approximately double their expected output (Figure 1, Mixes 2-4). In addition, when Cluster I A’s concentration in the artificial community was increased, its relative abundance in the output also increased. However, the increase was not proportional, and increasing initial concentrations of Cluster I A led to increased over-amplification. For instance, between Mix 2 and Mix 4, the concentration of Cluster 1 A in the output was expected to increase 2.3 fold, but its actual output increased 2.9 fold. Cluster II A was the only species whose actual output was consistently close to its expected output, although when Cluster I A was expected to be less then 10% of the output (Mix 1), it too was over-amplified (Table 1). pmoA species Cluster II B was also overamplified in Mix 1 but otherwise, it, along with pmoA speices Cluster I C, Cluster II D, and amoA species A, was under-amplified (Figure 1, Table 1). Cluster I C had the lowest amount of starting template in all the mixes, and its under-amplification resulted in the species not being recovered in two mixes (Mix 1 and 4), and amoA species A was also not recovered in Mix 4. 110 Discussion The cause of the observed PCR bias is unclear. It is possible that it is at least partially caused by the different binding energies from the degenerate positions within the primers (1). The GC content differences from the degenerate positions can result in a 2°C difference in annealing temperature, and subsequently a greater proportion of GC- rich template can bind to the primers and cause over-amplification as compared to primers with A or T at those positions (1). Due to the ambiguous positions there are 16 template sequences that a reverse primer could bind, and 4 sequences that a forward primer could bind. Only 2 of the species in this experiment (species Cluster II B and Cluster II D) actually have identical primer binding sites. As a result, the likelihood of the primer binding sites having different binding energies is high, and is possibly contributing to the PCR bias. However, if the degenerate primers were the primary cause of the bias we would expect to see (1) all GC-rich primer binding sites over-amplified, and (2) consistency in the pattern of PCR bias as the same species should be always be over-or under-amplified. We do not see either. With the exception of the consistent over-amplification of species Cluster I B, the results for all species were inconsistent; with some species being under- amplified in some mixes and over-amplified in others (i.e. Cluster II B). In addition, Cluster I B only has A or Ts in degenerate positions, and not G or Cs. We therefore conclude that the primer pair A189-A682 is not the primary cause of the PCR bias, and its biases are likely no worse than most other primer pairs. Regardless of the underlying cause, the inconsistent pattern of over- and under- amplification argues for the exclusion of the relative abundance data. Supporting that 111 conclusion is that as the concentration of the Cluster I A species increased with each mix, its over-amplification became worse. This is disconcerting because it indicates that communities that are seemingly dominated by an over-amplified species are more likely to have erroneous measures of relative abundance. The over-amplified Cluster I A dominates the tRF LP profiles fiom Ag HT (Chapter 2), and therefore the relative abundance measures from the Ag HT methanotroph communities are those that are most likely to be erroneous due to the PCR bias. For instance, the first replicate of Ag HT had Cluster I A at 98% relative abundance, and Cluster II A and MRI were both at 1% relative abundance. Looking at that result, or if put into a diversity index (i.e Simpson or Shannon), the presence of Cluster II A and MRI would be discounted due to Cluster I A’s dominance of the community. However, the PCR bias indicates that the 98% relative abundance is likely over-amplified at least 2-fold, so the true relative abundance of Cluster 1A is probably closer to 50%. If Cluster II A and MR1 are being under-amplified, as our results indicate for Cluster II A (see Mixes 3 and 4; MRI was not included in this experiment), then their true relative abundances are much higher then 1%. Therefore, the fairest and most representative way to present the data is to just consider whether a species is present or absent. Despite species Cluster I C not being recovered in two mixes, its lack of recovery was not associated with either high or low species Cluster I A input; indicating that the loss of its presence may have been independent of the amount of bias in a reaction. Mix 1, with the least over-amplification, did not recover cluster I C, and neither did Mix 4, which had the greatest over-amplification of species Cluster I A. Communities with 112 highly over-amplified species are expected to have an increased chance of not being able to recover methanotroph OTUs (as shown by the amoA A species being recovered in all mixes but Mix 4), but the inability to recover Cluster I C in the least biased mix suggests that the recovery of methanotroph OTU or species that are in low abundance is variable, and just as possible in communities with limited PCR bias and those with considerable PCR bias. Considering that 5 tRF LP profiles were summed to represent one replicate (Chapter 2), the expectation of recovering a low-abundance species even in treatments with the highest over-amplification of species Cluster I A is reasonable. Therefore, due to the PCR bias and the presence of nearly all methanotrophs in each mix, all methanotroph community analyses are presented only with presence/absence data (richness, Sorenson index), and exclude the relative measures of methanotroph abundance. The over-amplification of pmoA species Cluster I A and Cluster I B, and the considerable deviation of all species from their expected output led to the conclusion that abundance measures should not be included in the tRF LP or clone library community analyses of the pmoA community. If future investigators can identify more reproducible amplification conditions, then the use of the relative abundance measures could be justified and include in the analyses of the methanotroph community. 113 o 0 NH m.m < <23 H H NH mgr. a HH .336 <23 H 4 NH m.m m HH .386 <23 o 0 NH m.m < HH .336 <23 v .5... o o N H.N u H.330 <23 o 2 NH m.m m H.330 <23 H. 3 mm 3: < H.336 <23 m m 4H NH. < <23 4 H. «H N... a HH .336 <23 o 4 4H NH. m HH .336 <23 N 2 4H N... < HH .336 <23 n x... N N m m.N u H.336 <23 H H 4H NH. m H.335 <23 S 3 HN 3 < H .336 <23 N N 2 3. < <23 m m H 0.4 0 HH .386 <23 H N H a... a HH .336 <23 4 NH mH 2v < HH .336 33 N x... N N m N.N u H.325 <23 N NH. 3 3. m H.336 <23 m N H o... < H .336 <23 N H. 2 a... < <23 N S 2 we a HH .335 <23 H NN 3 as. a HH .336 <23 H N 3 a... < HH .325 <23 H x... o o 3 m.N u H.336 <23 H on 3 o... m H.325 <23 H m m m.N < H .325 <23 .95 35 33.6 .253 05 3 15 2.9.6 388.5 2: 33 gonna! 2: 5 (an 3.8% taste-um 00:23:30 250.0: 009.25 he 8:23:50 9583. 38.—Ea... 3 252.3 (OE. .3 (OER 360% «SEN. new «SEQ .8 33:38:26 .8595 cacao“. .8 ”59:0 .253 can 593 uofioaxo .coEmanoU .H.< use. .Nwoarmmgx :3 ..mEtQ 9.: no mm... «UN. 05 “Woman 3 flow: 114 I Expected 50 '1 Observed Relative Abundance in tRFLP Output (% of Total) Cluster IA Cluster I B Cluster I C Cluster H A Cluster H B Cluster 1] D amoAA pmoA or amoA Species Figure A.1. Expected and actual relative abundance of the artificial pmoA and amoA community Mix 2. 115 References 1. P012 MF 8:. Cavanaugh CM (1998) Bias in Template-to-Product Ratios in Multitemplate PCR. Appl. Environ. Microbiol. 64(10):3724-37 30. Suzuki M & Giovannoni S (1996) Bias caused by template annealing in the amplification of mixtures of 16S rRNA genes by PCR. Appl. Environ. Microbiol. 62(2):625-630. Holmes A], Costello A, Lidstrom ME, & Murrell IC (1995) Evidence that participate methane monooxygenase and ammonia monooxygenase may be evolutionarily related. FEMS Microbiology Letters 132(3):203-208. Abdo Z, et al. (2006) Statistical methods for characterizing diversity of microbial communities by analysis of terminal restriction fragment length polymorphisms of 16S rRNA genes. Environmental Microbiology 8(5):929- 938. 116 Appendix B Ammonia and Nitrate Before and After Fertilization in a KBS LTER Fertilized Sub-plot Introduction In addition to sampling one of the KBS LTER Late DF fertilized sub-plots, we also established a 10x Fertilized sub-plot. The 10x sub-plot was designed to allow investigation into the short-term effect of fertilization on the methanotroph community, but its community was not investigated after we found that long-term fertilization had no effect on methanotroph richness (Chapter 3). However, the results from the nutrient analyses of the inorganic N from the 10x sub-plots from before and after fertilization is presented here to illustrate that the nitrogen in the applied urea fertilizer is converted to ammonia that can be recovered from the soil. Even though we have no direct evidence that the ammonia from the fertilization inhibits the activity from the 1 OX sub-plot methanotrophs, we can confirm that the ammonia concentration after urea fertilization in the soil is high enough to be predicted to inhibit the rate of methane consumption. In addition, the data is also provided to illustrate the turnover of ammonia to nitrate that is indicative of considerable nitrification in the 10x sub-plot soils. Methods A 2x2 m sub-plot was established in each Late DF replicate adjacent to the other fertilized sub-plots. The 10x sub-plot received one application of 33 g N m'2 yr'l as urea, which was applied using a backpack sprayer on June 5th 2007. Three soil cores (2.5 x 10 cm) were collected from all 3 replicates on the following dates: June 1St 2007, June 117 13th 2007, June 27th 2007, August 14th 2007, and October 17th 2007, representing 4 days before fertilization, and 8, 22, 47 and 111 days after fertilization. Approximately 10g of each soil core was extracted with 100ml of 1 M KC], and prepared for analysis on a Flow Injector Analyzer as per the protocol available at http://lter.kbs.msu.edu/ (Soil Inorganic N). The three measures fiom each plot were then averaged to be a composite measure for that replicate, and the average of three biological replicates is represented in the Figure B. 1. Results and Discussion In the 10x sub-plot, 8 days afier fertilization, a big spike in ammonia concentrations was observed as the enzyme urease quickly converted urea to ammonia (l , 2). The concentration of ammonia at that date was 8 NH4-N ppm (or 8 pg NH4-N g soil- 1). That concentration, based on modeling by Hiitsch (3) is predicted to inhibit methane consumption. By 22 days after fertilization, nitrate concentrations have begun to rise as the ammonia is niuified, and by the third measurement, 47 days after fertilization, nitrate has peaked, and concentrations of ammonia have fallen. This conversion of most of the ammonia to nitrate in 47 days after fertilization is evidence of robust nitrification, and probably lessens the effect of the fertilization disturbance upon the methanotroph community as exposure to ammonia may be relatively limited. Therefore, we conclude that the urea fertilizer was converted rapidly to ammonia, the concentrations of ammonia were high enough to have inhibited methane consmnption, and ammonia was quickly nitrified to nitrate. 118 Acknowledgments Neville Millar was an equal partner in the soil extractions, and was responsible for running the samples through the Flow Injector Analyzer. 119 I l i 10.0 '1 r 5.0 i E +Control ‘ i l i ‘ 9.0 3‘ NH4-N ‘ 4.5 l t 8.0 ‘ +10x ‘ 4.0 i Fertilized ‘ ‘ 7.0 i NH4'N l? 3.5 i ‘ +Control \ 6.0 4 N03“ r 3.0 E l 3. ‘ a i a. +10x i (Z, 5‘0 \ Fertilized t 2'5 z.“ ‘t :5? 1 NO3-N ; O )2 4.0 ‘ » 2.0 Z . l l 1 . l 3.0 1 - 1.5 ' l l 2.0 t t 1.0 I l l l . 3 1.0 ~ 1 0.5 l - . 0.0 w e . . 3&1 0.0 -5 20 45 70 95 120 Days Since Fertilization Figure B. 1. Ammonia and nitrate concentrations in fertilized and control Late DF sub- plots at KBS LTER. Error bars represent the standard error. 120 References 1. Malhi SS, Grant CA, Johnston AM, & Gill KS (2001) Nitrogen fertilization management for no-till cereal production in the Canadian Great Plains: a review. Soil and Tillage Research 60(3-4):101-122. Rochette P, et al. (2009) Banding of Urea Increased Ammonia Volatilization in a Dry Acidic Soil. J Environ Qual 38(4):]383-1390. Hutsch BW (2001) Methane oxidation in non-flooded soils as affected by crop production - invited paper. European Journal of Agronomy 14( 4):237- 260. 121 Appendix C Biogeography of Methanotrophs in Well Drained Soils Introduction To confirm and extend the biogeography observed in the long-term fertilized sites in Chapter 3 (Figure 3.8), and to determine the biogeography of methanotrophs in well drained soils, a meta-analysis was performed using 27 clone libraries of the A subunit of the particulate methane monooxygenase (pmoA) (Table C.1). pmoA is found in all known methanotrophs (1) except two Methylocella strains (2). Due to the near ubiquity of pmoA, and the inference of function fiom the presence of the gene, it was chosen over comparing libraries that assess the methanotroph community via 16S ribosomal genes. Methods Using only pmoA sequences allowed for all of the sequences to be imported into Arb (3) where they were translated and aligned using Clustal W. Nucleic acid sequences were then aligned according to the protein sequence. Sequences from clone libraries were determined to be the same species if they were _>_ 94% identical (4) as determined by DOTUR (average neighbor grouping) (5). In order to facilitate the comparison between the methanotroph communities of the different soils species were used to calculate fi- diversity with the Sorenson index (PAST, (6) and Estimates (http://viceroy.eeb.uconn.edu/EstimateS)). The clustering between the communities was visualized using a neighbor-joining dendogram. Only pmoA libraries fiom well drained soils that consume atmospheric methane were included in the analysis. Libraries from environments that are net sources of 122 atmospheric methane like landfill cover soils, mine soil, rice paddies, wetlands, etc, were excluded fiom the analysis. If a study fi'om a well drained soil included a library from an experimental treatment of the same land use type (e.g. increased atmospheric C02; (7)) that library was excluded, and only the pmoA libraries from the control treatments were included. Only pmoA libraries that had at least 10 sequences were included in order to ensure that under-sampled methanotroph communities would not confound the analysis. Doing so excluded many pmoA libraries that were based on results from denaturing gradient gel electrophoresis (i.e. (8, 9)), and the Harvard forest and Sakerat Experimental Stations clone libraries included in the biogeography analysis of long-term fertilized sites (Chapter 3). Included in the final analysis were 27 pmoA libraries from well drained soils from Germany, the United Kingdom, the United States, and New Zealand (Table C. 1). The libraries were from 9 studies (Chapter 2, Chapter 3, (7, 10-14)) , and included a total of 1,560 pmoA sequences from a variety of forest, pasture, shrub land and agriculture soils. Results and Discussion The meta-analysis of methanotroph community compositions in well drained soils (Figure C. 1) revealed a distinct biogeography, and confirmed the patterns seen in Chapter 3 with long-term fertilized sites. Methanotroph communities generally clustered according to geographic location. The two exceptions to that pattern was 1 library from a German forest clustering with the Rothamsted Research libraries, and the two Hawaii soils not sharing any species, and therefore clustering separately. The Hawaii forests are 123 on soils that vary in age by approximately 50,000 years so it is not surprising that those libraries did not conform to the pattern of biogeography evident elsewhere (Table C.1). 124 :5 ”SN 352 2... wee. :5 ”SN £52 H...e 35. E «SN .3 a no: 2: HSN .He He Nam a: HSN .He .0 N8... 2: HSN ._e a Ne3H Hm: HSN .He .0 N3... 8: EN 2.; HesW .Hmem : HH NSN .He .0 swam HH HH 88 .He .0 .HweHm HH : 82 ._e .0 swam HH HH 88 .He .0 35m 2 HH EN .3 He .35 2: EN .H.w He :35 HNHH 38 He He 35. H .326 N .236 m .3920 m .0325 m .3925 m .2920 N .225 m 8320 N .2325 N .8325 N .325 N 5525 g .20 38» can boHaEHxER—a m. 3.28 on... N 55: 96. 2o .8» occdnA a co v2.82 2 86.8 2t. e .cowebi v5 62.8.9095 938883 .50 .3 £0.6— EoBE< a 6.23 :8 28:92 2 .52 332.30% H. .88on snoboHHafioE o no maggot 09.26 SH .23 Am bag—UV 355: 3.6365 0 .Ho 9:883 05 85838: .on9on £32859:— wd .Ho maggot omflo>a SH .23 AN BEE—0:01.25: .3239: v .8 9:85;». 05 358505 mNN moocozmom 62 23.3 302 3:3 3.2 22.3 .32 BanN 302 233 302 van—uoN 302 .3250 5835. use: 5839 uses 5830. 3...: .835. 3:5 5835. 3...: Scenes. use: 83% 93:5 8.3 3...: 8.3 Bee: 8.3 Bee: 8.3 Bee: 8.5. 8%: 33d N368.— Eceuv. @822 e698"— unEeguE @222 mace—mafio 33m 05.. 38m 33m 333m .3332 2.3m HBO 25m but? .83“— :ooom 530er20 838.. 22385 3.38; 25H 22mg 2.: .88... N as. $2.3 H323 H 33. .325, vHHaHHceeaH N as. 38:32.3 3333 H as. 30:52.3 3323 N .3 683 .55. H .3. 683 .85. N e3. .235 .380 an 23 H e3. 83.25 .228 ..E 23 N .5. .an 23 MH e3. .3 23 N e... H: w< NH .3 H: m< 3% 323:3. HauHmo—oB a: H 2 Hugo. 328:3.”— H 3c: «>2 wov— 2:82 33H 8220 ..c 525 .2339» nose—U. anti 5%.... 5:35 E320 .5530 .5520 cuss— E33380 8:033 .5533 32333 32383 3:55 3:5.— 33... 59580 4:288”— 38:553— ..noHaomox “c2255.:— vnufimox vonEasHox «£888.: BHmEafiox «£859.3— vonEafio: 4.3.383— 33:853— ...HmHHH 8:3 .8323. 32.3. Rm: 8:3 .3323 32.3. Rm: 8:3 22335 32.3. :35 8:3 .8365 MW3:3. Rm: 8:3 Hansen 33:3. Rm: 8:3 HeeHweHeHm 32.3. am $95 25 5 com: 8:95: V2.3 can 8H5 2: .8 >355 HQ 035... 125 Broadbalk Wilderness 2 Knott Wood 2 Knott Wood 1 Gottingen Forest Rothamsted Research Broadbalk Wilderness 1 Broadbalk Wheat 1 Broadbalk Wheat 2 Late DF Control Sub-plot 1 —|—— Late DF Control Sub-plot 2 Late DF 1 Late DF 3 KBS LTER Oak Stand __: Spruce Stand Acacia Forest, Mauna Kea, Hawaii Alder Stand Glsburn Pine Stand Jasper Ridge Grassland Metrosideros Forest, Kilauea Volcano, Hawaii Westview Pine Old-growth Beech Forest Waiouru Pasture t l——l 0.1 Westview Pasture New Zealand Waiouru Shrubland Puruki Pine Puruki Pasture Figure C.1. Similarity of methanotroph communities in soils from around the globe. The dendogram is based on Sarenson index calculations for each pairwise comparison of the methanotroph communities using pmoA species (defined as pmoA sequences having 94% average nucleotide sequence similarity), and then clustered using neighbor-joining with MEGA (48). The scale bars represent at 0.1 change in the Sorenson index. Additional site information is provided in Table C. 1. 126 References 1. 10. ll. Kitmitto A, Myronova N, Basu P, & Dalton H (2005) Characterization and structural analysis of an active particulate methane monooxygenase trimer from Methylococcus capsulatus (Bath). Biochemistry 44(33):]0954-10965. Dunfield PF, Khmelenina VN, Suzina NE, Trotsenko YA, & Dedysh SN (2003) Methylocella silvestris sp nov., a novel methanotroph isolated from an acidic forest cambisol. International Journal of Systematic and Evolutionary Microbiology 53:1231-1239. Ludwig W, et al. (2004) ARB: a software environment for sequence data. Nucl. Acids Res. 32(4):]363-1371. Konstantinidis KT & Tiedje JM (2004) Trends between gene content and genome size in prokaryotic species with larger genomes. Proceedings of the National Academy of Sciences of the United States of America 101(9):3160-3165. Schloss PD & Handelsman J (2005) Introducing DOTUR, a Computer Program for Defining Operational Taxonomic Units and Estimating Species Richness. Appl. Environ. Microbiol. 71(3):1501-1506. Hammer 0, Harper DAT, & Ryan PD (2001) PAST: Paleontological statistics software package for education and data analysis. Palaeontologia Electronica 4(1): 1-9. Horz H-P, Rich V, Avrahami S, & Bohannan BJM (2005) Methane-Oxidizing Bacteria in a California Upland Grassland Soil: Diversity and Response to Simulated Global Change. Appl. Environ. Microbiol. 71(5):2642-2652. Knief C, Lipski A, & Dunfield PF (2003) Diversity and activity of methanotrophic bacteria in different upland soils. Applied and Environmental Microbiology 69(11):6703-6714. Zhou XQ, et al. (2008) Effects of grazing by sheep on the structure of methane- oxidizing bacterial community of steppe soil. Soil Biology & Biochemistry 40(1):258-26l. Singh BK & Tate K (2007) Biochemical and molecular characterization of methanotrophs in soil from a pristine New Zealand beech forest. FEMS Microbiology Letters 275(1):89-97. Singh BK, et al. (2007) Effect of afforestation and reforestation of pastures on the activity and population dynamics of methanotrophic bacteria. Applied and Environmental Microbiology 73(16):51 53-5 161 . 127 12. 13. 14. Ricke P, Kolb S, & Braker G (2005) Application of a newly developed ARB software-integrated tool for in silico terminal restriction fragment length polymorphism analysis reveals the dominance of a novel pmoA cluster in a forest soil. Applied and Environmental Microbiology 71 (3): 1 671-1673. Reay DS, Radajewski S, Murrell J C, McNamara N, & Nedwell DB (2001) Effects of land-use on the activity and diversity of methane oxidizing bacteria in forest soils. Soil Biology & Biochemistry 33(12—13): 1613-1623. King GM & Nanba K (2008) Distribution of Atmospheric Methane Oxidation and Methanotrophic Communities on Hawaiian Volcanic Deposits and Soils. Microbes and Environments 23(4):326-330. 128