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 OCH-"0 9 “ta-7 " V ‘1'“ xi Qua- SEP 2 2 20.7 EBUIZUU 5 no '17 A I‘D A r 7 T Qp‘fQ 3 '0 7 6’01 cJCIRC/DateDuepBS—p. 15 LINKING PLANT COMMUNITIES TO SOIL MICROBIAL COMMUNITIES AND PROCESSES IN OLD-FIELDS By Laura C. Broughton A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Zoology Ecology, Evolutionary Biology, and Behavior Program 2001 ABSTRACT LINKING PLANT COMMUNITIES TO SOIL MICROBIAL COMMUNITIES AND PROCESSES IN OLD-FIELDS By Laura C. Broughton The resources that support soil microbial communities are primarily derived from plants, so the soil microbial community should respond to changes in plant diversity or productivity, particularly if changes in the plant community affects the quality or quantity of available resources. I investigated the role of soil and plants on the structure and function of the soil microbial community by conducting observational and experimental studies and two manipulative greenhouse experiments. I examined the relationship between plant diversity and productivity and soil microbial community structure and function along a topographic gradient in a successional old—field in Michigan. Variation in plant productivity was confounded by changes in plant community diversity and edaphic characteristics, so I could not determine which of these variables caused the observed changes in the soil microbial community. To further investigate the relationship between the soil microbial community and plant species diversity, I sampled soils from a set of experimental grassland plant communities established as part of the BIODEPTH experiment at Silwood Park, England. Plant species diversity, functional group diversity, and species composition varied across treatments. I found that plant diversity significantly affected soil microbial community structure. However, N-mineralization rates and microbial respiration responded to variation in plant community composition, but not diversity. In a greenhouse experiment I examined how variation in soil fertility influenced the soil microbial community. I found that soil origin had strong effects on the structure and function of the soil microbial community. Higher fertility soils had higher organic nitrogen pools and microbial activities and more eukaryotes in the microbial community. In addition, the presence of Andropogon gerardi also affected the structure and function of the soil microbial community. However, the magnitude of the plant effect on soil microbial respiration was inversely related to soil fertility. In a second greenhouse experiment I further explored the plant species effect on the soil microbial community. I found strong effects of both plant species identity and soil origin on the structure and function of the soil microbial community. In particular, the presence of a legume (T rifolium pratense) increased soil nitrogen cycling processes. Plant species identity had a small effect on soil microbial community structure, but it was dwarfed by the soil origin effect. Results from these studies indicate that several aspects of the plant community, including diversity, composition and individual plant species identities, can strongly influence the structure and function of the soil microbial community. However, other environmental factors that affect soil quality can have strong and persistant effects on the soil microbial community. ACKNOWLEDGMENTS I would like to thank my advisor, Kay Gross, for her unlimited support, patience, and guidance over the last six years. She has been particularly understanding this past year as I have faced health problems that interfered with my work. I will be forever grateful to Kay. I would also like to thank the other members of my committee, Steve Hamilton, Mike Klug, and Gary Mittelbach for their guidance and help in shaping the path of my graduate career. This work could not have been completed without the help of the members of Kay Gross’ lab. I would particularly like to thank Carol Baker, Karen Tindall, and Jaime Fiordalis for their technical assistance and help with field work; and Bryan Foster, Andrea Corbett, Melissa McCormack, Heather Reynolds, Wendy Goodfiiend, Greg Houseman, Rich Smith, and Sarah Emery for their intellectual feedback and companionship. I am grateful to Michel Cavigelli, Mike Kaufman, Christine Easley, and Terry Marsh for help with laboratory techniques. The statistical work could not have been done without timely help from Carl Ramm, Chris Blackwood, and Hal Collins. Additionally, the BIODEPTH portion of my work could not have been completed without the help of my collaborator, Andy Hector, and John Lawton, Asher Minns, and Ellen Bazeley- White. Special thanks to participants in the summer 1996 Advanced Field Ecology and iv Evolution course at the W. K. Kellogg Biological Station who helped with the initial soil and plant sampling for the field survey at Lux Arbor. Living and working at the Kellogg Biological Station can be a lonely endeavor. For their fellowship, dinner groups, game nights, and great parties, I am grateful to my fellow graduates students, particularly Eric Thobaben, Tara Darcy, and Natalie Dubois. I enjoyed singing with Eric in the Bach Festival Chorus and talking about the trivial but fun aspects of our lives. Stefanie Whitmire has been my constant companion for the past four years and I could not have made it through without her. I am eternally grateful to Stefanie. For their encouragement, love, and support throughout my life, I would like to thank my parents, Sharon and Gary Broughton. I could never have made it this far without them. My brother, Keith Broughton, has also shown constant support for my chosen path. The research was funded by NSF grants to the Center for Microbial Ecology at Michigan State University, a Research Training Grant at the W. K. Kellogg Biological Station, and a Dissertation Completion Grant from the College of Natural Science at Michigan State University. I obtained additional monies from the Michigan State University Department of Zoology, the Michigan State University Ecology, Evolutionary Biology, and Behavior Program, and Sigma Xi. TABLE OF CONTENTS LIST OF TABLES .............................................................................. ix LIST OF FIGURES ............................................................................. xii INTRODUCTION .............................................................................. 1 THESIS OVERVIEW ................................................................. 2 FIELD SITES ........................................................................... 6 REFERENCES ......................................................................... 9 CHAPTER 2: PATTERNS OF DIVERSITY IN PLANT AND SOIL MICROBIAL COMMUNITIES ALONG A PRODUCTIVITY GRADIENT IN A MICHIGAN OLD FIELD ............................................................... 10 INTRODUCTION ..................................................................... 10 METHODS .............................................................................. 12 Site Description ................................................................ 12 Sampling Design and Characterization of the Gradient .................. 12 Characterization of the Soil Microbial Community ....................... 14 Statistical Analyses ............................................................ 16 RESULTS ............................................................................... l7 Characterization of the Gradient ............................................. 17 Plant and Soil Microbial Community Relationships with Productivity ............................................................ 1 9 Compositional Shifts in the Plant and Soil Microbial Communities... 21 DISCUSSION .......................................................................... 26 REFERENCES ......................................................................... 34 CHAPTER 3: LINKING PLANT COMMUNITY DIVERSITY TO SOIL MICROBIAL COMMUNITIES: AN EXPERIMENTAL EVALUATION FROM THE BIODEPTH EXPERIMENT ............................................................ 37 INTRODUCTION ..................................................................... 37 MATERIALS AND METHODS .................................................... 39 Site .............................................................................. 39 Soil Sampling and Analysis .................................................. 40 Statistical Analysis ............................................................ 46 RESULTS ............................................................................... 48 Legacies of Site Preparation ................................................. 48 Effects of Species Richness and Functional Group Richness Treatments on the Soil Microbial Community ................... 48 Effects of Functional Groups or Individual Species on the Soil Microbial Community ............................................... 53 vi Effects of Plant Community Composition (Mixture) on the Soil Microbial Community ............................................... 54 Changes in Soil Microbial Community Composition .................... 54 DISCUSSION .......................................................................... 58 Plant Diversity Effects on the Soil Microbial Community ............. 59 Plant Community Composition Effects on the Soil Microbial Community ......................................................... 61 Legacy Effects on the Soil Microbial Community ....................... 63 REFERENCES ......................................................................... 66 CHAPTER 4: PLANT-MEDIATED EFFECTS OF SOIL ORIGIN ON THE COMPOSITION AND FUNCTION OF SOIL MICROBIAL COMMUNITIES ...... 69 INTRODUCTION ..................................................................... 69 METHODS .............................................................................. 71 Site Descriptions ............................................................... 71 Experimental Design .......................................................... 72 Data Collection ................................................................ 74 Statistical Analyses ............................................................ 77 RESULTS ............................................................................... 79 Plant Communities ............................................................ 79 Effects of Soil Origin on Soil and Microbial Processes .................. 80 Effects of Andropogon gerardi on Soil and Microbial Processes ...... 83 Plant and Soil Effects on Soil and Microbial Processes .................. 85 Plant and Soil Effects on Soil Microbial Community Structure. . . . . 87 DISCUSSION ........................................................................... 95 Site Fertility Effects on Soil Microbial Processes ........................ 96 Historical Plant Diversity Effects on Soil Microbial Processes ......... 97 The “Plant” Effect on Soil Microbial Processes .......................... 100 Plant and Soil Effects on Soil Microbial Community Structure ........ 102 REFERENCES ......................................................................... 104 CHAPTER 5: AN EXPERIMENTAL EVALUATION OF THE EFFECTS OF DIFFERENT PLANT SPECIES ON THE STRUCTURE AND FUNCTION OF SOIL MICROBIAL COMMUNITIES ....................................................... 108 INTRODUCTION ..................................................................... 108 METHODS .............................................................................. 110 Site and Species Descriptions ............................................... 110 Experimental Design .......................................................... 1 12 Data Collection ................................................................ 1 15 Statistical Analyses ............................................................ 1 16 RESULTS ............................................................................... 118 Plant Communities ............................................................ l 18 Time Effects on Soil and Microbial Processes ............................ 119 Plant Species Responses to Soil Types ..................................... 119 vii Effects of Soil Origin on Soil and Microbial Processes .................. 125 Plant Effects on Soil and Microbial Processes ............................ 128 Interactive Effects of Plant Species and Soil Origin on Soil and Microbial Processes .................................................. 129 Plant and Soil Effects on Soil Microbial Community Structure. . . 130 A Comparison of the Microbial Communities within and outside the Root Exclosures ..................................................... 135 DISCUSSION ........................................................................... 138 The Effects of Soil Origin on Soil and Microbial Processes and Soil Microbial Community Structure .................................... 139 Plant Effects on Soil and Microbial Processes and Soil Microbial Community Structure ................................................ 140 Implications of the Similarity of Soil Microbial Communities in the Root Exclosures and the Rhizosphere Soil ........................ 145 REFERENCES ......................................................................... 147 CHAPTER 6: SUMMARY .................................................................... 151 viii LIST OF TABLES Table 1.1. Plant productivity and diversity of abandoned old fields from which soil was collected for the greenhouse experiments. Values for peak aboveground plant biomass (an estimate of primary productivity), species richness, and mean percent organic matter are expressed as mean 1" standard deviations (n = 6). Values that are significantly different for a given variable based on Fisher’s LSD test are indicated by different letters .............................................. 7 Table 2.1. Plant species composition in the clusters from the K—means cluster analysis shown in Figure 2.3B. Plant species are listed from most common (by biomass) to least common. Only species that have a total biomass > 1 g for the cluster are listed. Nomenclature follows Gleason and Cronquist (1991) ..................... 25 Table 2.2. Fatty acids used in K—means cluster analysis of FAME profiles .............. 27 Table 3.1. (A) Plant species pool used in establishing the experimental communities and (B) the plant communities sampled for the soil microbial community. Plots sampled only in year 3 (1998) of the experiment are in italics. The control and Rumex plots were not maintained through year 4 and could not be re-sampled. Plots sampled in both years are indicated in bold. Other plots were sampled only in year 4 (1999) of the experiment. Communities are grouped by number of plant functional groups (FG) present ............................................... 41 Table 3.2. Analysis of variance model with sequential sums of squares used to evaluate the effects of the plant community manipulations on soil and microbial community parameters .................................................................. 47 Table 3.3. Significance and direction (for diversity) of treatment effects on soil and microbial parameters for both years as detected by Analysis of Variance using Type I sums of squares (model in Table 3.2) for (a) species diversity, functional group diversity, and composition and (b) presence/absence of legumes or forbs. ANCOVA results using above-ground plant biomass as the covariate are shown in parentheses if the effect changed in significance. NS=not significant. . . . . ...50 Table 3.4. Fatty acids used in principal components analysis of PLF A profiles ........ 51 Table 4.1. Plant productivity and diversity of abandoned fields from which soil was collected for the greenhouse experiment. Values for Annual Net Primary Productivity, species richness, and mean percent organic matter are expressed as mean i standard deviations. Values that are not significantly different for a given variable based on Fisher’s LSD test have the same letter ................... 80 Table 4.2. Effect of soil origin and presence of A. gerardi on (A) plant and (B) soil and microbial variables as detected by Analysis of Variance. NS = not significant, p ix > 0.05. The time factor investigates the results of harvesting half the experiment at 12 weeks, the other half at 16 weeks ................................. 81 Table 4.3. Effect of soil origin and presence of A. gerardi on the soil microbial community as detected by Distance-Based Redundancy Analysis (db-RDA) of CLPP profiles. (A) Significance values for the permutation tests on the environmental factors of the RDA. (B) Variance explained by species data and species-environment correlations for the RDA ....................................... 88 Table 4.4. Fatty acids used in principal components analysis of PLFA profiles. I describe fatty acids using standard nomenclature where the total number of carbon atoms appears before the colon and the total number of C-C bonds appears after it. Cyclo-propane analogs are indicated by "cyclo," and the location of the epoxy bond is indicated by a "c" followed by two numbers. If the cis or trans configuration is unknown, the word "at" is used. The number following "cis," "trans," or "at" indicates the location of the double bond in relation to the carboxyl end of the molecule. Fatty acids with the same retention time are grouped as "sum in feature" and given a unique number designation..91 Table 4.5. Effect of soil origin, time of harvest, and presence of A. gerardi on the soil microbial community as detected by Distance-Based Redundancy Analysis of PLF A profiles. (A) Significance values for the permutation tests on the environmental factors of the RDA. (B) Variance explained by species data and species-environment correlations for the RDA ....................................... 92 Table 5.1. Plant species by functional group and aboveground biomass at the two field sites from which soil was collected. Functional groups are coded by Grass (G), Forb (F), Legume (L), and Woody (W). Plant species used in this greenhouse experiment are in bold ................................................................. 111 Table 5.2. Treatments used in the Randomized Complete Block Design to test for the effects of plant species on the soil microbial community. Each treatment was replicated three times within three Time blocks on two different soils. Species planted were G = Bromus inermis, F = Solidago canadensis, L = T rifolium pratense. 0 = absent, + = present, C = control (no plants) ........................ 114 Table 5.3. Summary of ANOVA results analyzing effect of soil origin and (A) individual plant treatments, (B) individual plant treatments without no-plant controls, or (C) factorial plant treatments on plant and soil and microbial variables. Species codes in 3C are: BROIN, Bromus inermis; SOOCA, Solidago canadensis; TRFPR, T rifolium pratense. NS = not significant, p > 0.05. ANCOVA results, using total plant biomass as the covariate, are shown in parentheses if the effect changed in significance ................................... 122 Table 5.4. Root to shoot ratios for each plant species in each soil and treatment. Values are mean 3: standard error. Significance values from the ANOVA are listed in Table 5.3C ............................................................................... 126 Table 5.5. Fatty acids used in principal components analysis of PLFA profiles. I describe fatty acids using standard nomenclature where the total number of carbon atoms appears before the colon and the total number of C-C bonds appears after it. Cyclo-propane analogs are indicated by "cyclo," and the location of the epoxy bond is indicated by a "c" followed by two numbers. If the cis or trans configuration is unknown, the word "at" is used. The number following "cis," "trans," or "at" indicates the location of the double bond in relation to the carboxyl end of the molecule. Fatty acids with the same retention time are grouped as "sum in feature" and given a unique number designation ............................................................................... 13 1 Table 5.6. Effect of soil origin, time, replicate, and presence of Bromus inermz's (BROIN), Solidago canadensz's (SOOCA), and T rifolium pratense (TRFPR) on the soil microbial community as detected by Distance-Based Redundancy Analysis (db-RDA) of PLFA profiles. (A) Significance values for the permutation tests on the environmental factors of the RDA. (B) Variance explained by species data and species-environment correlations for the RDA ...................................................................................... 134 Table 5.7. Effect of exclosure, soil origin, time, and presence of Bromus inermis (BROIN), Solidago canadensis (SOOCA), and T rifolium pratense (TRFPR) on the soil microbial community as detected by Distance-Based Redundancy Analysis (db-RDA) of PLFA profiles. (A) Significance values for the permutation tests on the environmental factors of the RDA. (B) Variance explained by species data and species-environment correlations for the RDA ...................................................................................... 137 xi Figure LIST OF FIGURES 1.1. Map of the W. K. Kellogg Biological Station with codes illustrating the location of the six field sites (after Foster 1996). Table 1.1 describes these sites .......................................................................................... 8 Figure 2.1. (A) Factor loadings plot for PC 1 and 2 from the PCA of the June 1996 sampling for light at ground level, soil moisture, soil inorganic N, and peak plant biomass; PC 1 and 2 account for 64.8% and 17.1% of the variation, respectively. Correlations between factor 1 and light (r = -0.77), moisture (r = 0.73), N (r = 0.80), and biomass (r = 0.91) were significant at p < 0.001; correlations between factor 2 and light (r = 0.56) and moisture (r = 0.56) were significant at p < 0.005. (B) The relationship between PC 1 and distance down slope for 34 of the 35 sampling points, y = -1.69 + 0.058x; Lines depict regression and 95% confidence intervals .............................................. 18 Figure 2.2. The relationship between productivity index and several variates describing Figure the plant and microbial communities at this site. (A) Plant species richness, y = 10.18 - 0.71x; (B) The number of carbon sources metabolized by the soil microbes after 48 hours; (C) Biolog average well color development after 48 hours, y = 0.89 + 0.14x; (D)ug COz-C/ g dry soil/ hour evolved from SIR control after 40 hours, y = 0.54x — 1.55; (E) ug COZ-C/ g dry soil/ hour evolved from the SIR glucose addition treatment after 40 hours, y = 0.33x — 2.80. Lines depict regression and 95% confidence intervals, 11 = 34 ............................ 20 2.3. Results from K-means cluster analysis evaluating the changes in production and composition of the plant and soil communities in relation to transect position at this site. (A) Contour plot of above-ground plant biomass, values are g/ m2; (B) Plant species composition: cluster 1 has no dominant species, cluster 2 is dominated by Rubus sp., cluster 3 by Solidago canadensis, cluster 4 by S. canadensis and Poa pratense, cluster 5 by Agropyron repens, cluster 6 by Poa pratense, and cluster 7 by Poklfvgonum amphibium var. emersum (see Table 1 for species lists). (C) BiologT carbon source utilization profiles separated into two clusters based on AWCD; and (D) FAME profiles: cluster 2 had smaller proportions of 18:1 cis 9, 16:0, and summed in feature 9 (18:2 cis 9, 12 and 18:0 anteiso, Table 2) than cluster 1 ........................... 23 Figure 3.1. Plant community composition effects on soil microbial biomass in (A) year 3 and (B) year 4 of the Silwood Park BIODEPTH study. Values are Mean :1: 1 SE. Plant community identification codes are from Table 3.1B. No-plant control plots are represented by a blank circle, while non-manipulated reference communities are represented by stars. Monocultures are indicated by circles and mixtures by triangles. Plots that have Hypochaeris radicata have hatch- marks ...................................................................................... 49 xii Figure Figure 3.2. Relationship between number of culturable soil bacteria and (A) plant species richness (R2=0.294, p<0.01), (B) number of plant functional groups (R2=0.288, p<0.01), and (C) aboveground plant biomass in year 3 (R2=0.573, p<0.01). No-plant control plots are represented by blank circles, while non- manipulated reference communities are represented by stars. Monocultures are indicated by circles and mixtures by triangles. Plots that have Lotus corniculatus have hatch-marks slanting up to the right; plots that have Hypochaeris radicata have hatch-marks slanting down to the right. The two plots with both species are mottled .................................................... 52 3.3. Plant diversity effects on the metabolic activity of the soil microbial community as measured by Community Level Physiological Profiles (CLPP) in year 3 (n = 28). Principal component axis 1 accounted for 46.1% of the total variation. Principal component axis 3 accounted for 7.6% of the overall variation and was driven by the ability to metabolize alpha-D-lactose. Significance values for diversity effects are listed in Table 3.3A. Number of functional groups is represented by symbol shape. No-plant control plots are represented by circles, and non-manipulated control communities labeled with “R”. Number of plant species (0, l, 4, 8, or 11) are labeled ........................ 55 Figure 3.4. Plant community composition effects on the structure of the soil microbial community, measured with Phospholipid Fatty Acid profiles of 1998 soil samples. Principal component axis 4 accounted for 4.5% of the overall variation and reflected the amount of 15:0 anteiso in the PLFA profiles. Plant community identification codes are from Table 3.1B. No-plant control plots are represented by a blank circle. The reference plots were not included in this analysis. Monocultures are indicated by circles and mixtures by triangles .................. 57 Figure 4.1. Total plant biomass of A. gerardi produced at 12 and 16 weeks in relation to variation among sites in 1998 field above-ground plant biomass. Soils are coded as in Table 4.1. Values are mean :1: standard error, n = 8. Significance values from the ANOVA are listed in Table 4.2A ................................... 82 Figure 4.2. Variation among sites in (A) total inorganic nitrogen, (B) N-mineralization Figure rate, (C) nitrification rate, (D) microbial respiration, (E) number of culturable bacteria, and (F) microbial biomass. Sites are arranged in order of increasing productivity, as listed in Table 4.1. For each variable, samples are distinguished between treatments with Andropogon gerardi (solid) and no-plant controls (open symbol). Values are mean i standard error, n = 8 for all except number of culturable bacteria and microbial respiration which have n = 4. Significance values from the ANOVA are listed in Table 4.28 ................................... 84 4.3. The magnitude of the plant effect on soil and microbial processes across sites in relation to the plant biomass gradient for (A) total inorganic nitrogen, (B) N-mineralization rate, (C) Nitrification rate, (D) microbial respiration, (E) number of culturable bacteria, and (F) microbial biomass. Values are mean i- xiii Figure Figure Figure Figure Figure Figure standard error, n = 8 except number of culturable bacteria and microbial respiration where have n = 4 ............................................................ 86 4.4. Soil origin and plant effects on the structure of the soil microbial community as measured by CLPP. CLPP patterns are distinguished between soil microbial communities from the A. gerardi (solid symbols) and no plant treatments (open symbols). Soil microbial communities in soils from different sites are indicated by symbols: circles, BA; triangles, FK; upside-down triangles, LL; diamonds, MK; stars, PL; pentagons, UL. Significance values from the RDA are listed in Table 4.4 ...................... ' ............................ 89 4.5. Distance-based Redundancy Analysis of PLFA profiles, investigating the effects of soil origin, plant, and time effects on the structure of the soil microbial community. The PLFA profiles are distinguished between soil microbial communities from the A. gerardi (solid symbols) and no plant treatments (open symbols). Soil microbial communities in soils from different sites are indicated by symbols: circles, BA; triangles, FK; upside-down triangles, LL; diamonds, MK; stars, PL; pentagons, UL. Labels for phospholipid fatty acids are listed in Table 4.3. Significance values from the RDA are listed in Table 4.5 ............................................................................... 93 5.1. Total plant biomass of Bromus inermis, Solidago canadensis, and T rifolium pratense, coded by A) soil origin and plant treatment, treatment codes are in Table 5.2; and plant species by plant treatment for soil from B) Field K (PK) and C) Upper Louden (UL). For (A), values are mean :1: standard error; for (B) and (C), values are means. Significance values from the ANOVA are listed in Table 5.3A ............................................................................... 120 5.2. Root biomass of Bromus inermis, Solidago canadensis, and T rifolium pratense, coded by (A) soil origin and plant treatment, treatment codes are in Table 5.2; and plant species by plant treatment for soil from (B) Field K (PK) and (C) Upper Louden (UL). For (A), values are mean :1: standard error; for (B) and (C), values are means. Significance values from the ANOVA are listed in Table 5.3A ............................................................................... 121 5.3.Variation among plant treatments in (A) total inorganic nitrogen, (B) N- mineralization rates, (C) nitrification rate, (D) microbial respiration, and (E) microbial biomass. For each variable, samples are distinguished between the two sites: FK (filled circles) and UL (hollow circles). Values are mean :1: standard error. Significance values from the ANOVA are listed in Table 5.3A ....................................................................................... 127 5.4. Distance-based Redundancy Analysis of Phospholipid Fatty Acid profiles, investigating the effects of soil origin, plant, and time effects on the structure of the soil microbial community. (A) Canonical principal component plot for the full RDA. The PLFA profiles are distinguished between soil microbial xiv Figure Figure communities from FK soil (solid symbols) and UL soil (open symbols). Soil microbial communities in soils with different plant communities are indicated by symbols: circles. controls; triangles, monocultures; four or five-sided polygons, mixtures. Labels for phospholipids are listed in Table 5.4. Significance values from the RDA are listed in Table 5.5A ...................... 132 5.5. (A-C) Partial canonical principal component plots derived fiom the redundancy analysis of PLFA profiles. (A) Bromus inermis axis (1 of 1) with soil origin, replicate, Solidago and T rifolium treatments, and time partialled out. (B) Solidago canadensis axis (1 of 1) with soil origin, replicate, Bromus and T rifolium treatments, and time partialled out. (C) T rifolium pratense axis (1 of 1) with soil origin, replicate, Bromus and Solidago treatments, and time partialled out. Labels for phospholipids are listed in Table 5.4. Significance values from the RDA are listed in Table 5.5A ...................................... 133 5.6. Distance—based Redundancy Analysis of Phospholipid Fatty Acid profiles, investigating the effects of exclosure, soil origin, plant, and time effects on the structure of the soil microbial community. Canonical principal component plot for the full RDA. The PLFA profiles are distinguished between soil microbial communities from FK soil (solid symbols) and UL soil (open symbols) ....... 135 XV CD IOC Pia CHAPTER 1 INTRODUCTION The role of species diversity and composition in ecosystems is increasingly under scrutiny due to concerns about the potential impacts of the current rapid decline in the Earth’s biodiversity. Species diversity may have important effects on key ecosystem functions like nutrient cycling, water quality, and productivity Rosenzweig 1995, Tihnan 1996). Recent investigations into the relationship between diversity and function have focused mainly on how changes in primary producers and consumers affect ecosystem processes (Schlapfer and Schmid 1999, Rosenzweig 1995). The interaction between aboveground and belowground (soil) communities in mediating these processes has been less studied (Schlapfer and Schmid 1999, Ohtonen et al. 1997). To understand controls on diversity and the role of diversity in ecosystem function it is important to understand the relationships among organisms in the ecosystem. While considerable attention has been paid to factors that affect the composition and function of communities of macroorganisms, very little is known about the factors that affect the structure of soil microbial communities (Ohtonen et al. 1997, Tiedje 1995). Resources available to soil microorganisms are primarily derived from plants. Most of the carbon and nitrogen entering the soil matrix results from litterfall, root exudates, or root death (Paul and Clark 1996). As a result, the composition and productivity of the plant community influences the soil microbial community. Similarly, the productivity OI mic cor C la by Sci relt‘ qua con 9C0. The or diversity of the plant community may be affected by processes mediated by soil microorganisms (e.g. N-mineralization rates). By performing key steps in the cycling of nutrients (carbon, nitrogen, and phosphorus, among others), the soil microbial community plays an essential role in the functioning of terrestrial ecosystems (Paul and Clark 1996). In most temperate grassland and forest systems, plant growth is limited by nitrogen (or a combination of nitrogen and phosphorus, Shaver and Chapin 1980, Schmidt et al. 1997, Jonasson et al. 1999). The soil microbial community controls the release of inorganic nitrogen to plants; however, the soil microbial community is most often limited by carbon (Zak et al. 1994). Therefore, the rate at which limiting nutrients are made available to plants is likely to be influenced by the amount and quality of carbon available to soil microorganisms. Consequently, changes in the plant community likely will change the soil microbial community and potentially affect ecosystem function. Thesis Overview I am interested in the influence of the plant community on the structure of soil microbial communities and the processes they mediate. The challenge is to distinguish between the direct effects of plants (through changes in soil carbon inputs) and indirect effects (due to soil characteristics) on the structure of the soil microbial community. In this dissertation, I explore the relationship between the plant and soil microbial communities through a combination of observational studies and manipulative experiments at two different scales: the plant community scale and the individual plant mi ant C1 C01 pro pro bep H01 COm 1mg species scale. I addressed the following questions through my research: (1) How does plant productivity affect the activity and structure of the soil microbial community? (2) How does plant community diversity affect the structure and function of the soil microbial community? (3) How does plant community composition affect the structure and function of the soil microbial community? (4) How do soil factors influence the structure and function of the soil microbial community and can plants mediate soil effects? Chapter 2 examines the first two questions on the relationships between plant community productivity and diversity and the soil microbial community. I compared patterns of diversity in the plant and soil microbial communities along a productivity gradient in an old field at the Lux Arbor Reserve at the W. K. Kellogg Biological Station in southwestern Michigan. The sampled gradient had a high diversity — low productivity plant community on the ridge top that graded into a low diversity —high productivity plant community down-slope. There was a strong positive relationship between above ground plant biomass and soil microbial respiration at the site. However, this association was confounded by changes in edaphic characteristics (moisture and nitrogen) and with the composition of the plant community that also varied along the gradient. Distinguishing plant from soil effects on the soil microbial community is a necessary first step in determining factors that structure the composition and affect the function of soil microbial communities. In the following chapters I describe the results of field and greenhouse experiments designed to investigate the independent effects of plants and soils on the soil microbial community. Ioz strut fimi Silu exp: were sml conu with . pknn anesi 4i. 1 ( Plant. pmdm dfil‘cm To address questions of how plant community diversity and composition influence the structure and function of the soil microbial community, I sampled soil communities from a series of experimental grassland plots from the BIODEPTH experiment at Silwood Park, England. This research is described in Chapter 3. In the BIODEPTH experiment, the number of plant species and the number of plant functional groups were varied to create plant communities with different plant diversities on the same soil. I found that plant community biomass, composition, and diversity all affected the composition and several functional traits of the soil microbial community. Many field studies have detected differences among soil microbial communities sampled from sites with contrasting plant communities (Zak et al. 1994, Grayston and Campbell 1996, Chapter 2), but in these studies plant and soil effects are confounded. Soil in different sites has been shaped by a variety of factors besides differences in plant community composition, such as parent material, disturbance and management regimes. Therefore, differences among soil microbial communities sampled from sites with different plant communities cannot be attributed solely to the differences in the plant communities because the soil characteristics and histories also differ. To investigate the direct effects of soil origin on the soil microbial community (question 4), I conducted a greenhouse experiment that compared soils from six different local plant communities. The six sites differed in fertility, soil organic matter, and plant productivity, and these factors had detectable, correlated effects on soil processes. To determine if plants could mediate these differences, a common plant species, .‘I’lti COII som 860 soil com com and and thrc dclc {Um soil This 11111: disc Andropogon gerardi, was grown in each soil. Interestingly, when soil microbial communities were grown in the presence of this species, the effect of soil origin on some functions (6. g. microbial respiration and N-mineralization rates) was diminished. Because soil microorganisms are dependent on carbon, and most available carbon in soil comes from plants, the identity of the plant species supplying carbon to the microorganisms may influence the structure and function of the soil microbial community. In Chapter 5, I describe the results of a greenhouse experiment in which I compared the effects of three plant species, grown in two distinct soils, on the structure and function of the soil microbial community. Soils from two of the old fields (high and low fertility) used in the previous experiment were planted with all combinations of three plant species common to local old fields. The experiment allowed me to determine that (1) different plant species can have unique effects on the structure and function of soil microbial communities and (2) the effects of different plant species on soil microbial community structure and function are non-additive. This dissertation suggests that both the origin of the soil and the presence of a plant influence the structure and functioning of the soil microbial community. In Chapter 6 I discuss the overall conclusions from this collection of field and greenhouse studies. Fie abai “as con- Elba Field Sites The greenhouse experiments presented in Chapters 4 and 5 of this dissertation were conducted with soils from six successional old—fields at the W. K. Kellogg Biological Station of Michigan State University in southwestern Michigan (Kalamazoo County; 42° 24’ N, 85° 24’ W). The sites varied in fertility, species richness, and dominant plant type, but all were located on Kalamazoo sandy loam soil (Table 1.1, Figure 1.1, Burbank et al. 1992). The six sites also differed in past land use and time since abandonment ranging in age from 20 to 50 years. McKay (MK) field was abandoned from agriculture in 1973; a section was plowed once in 1981 and then re-abandoned (Burbank et al. 1992). Both the Upper (UL) and Lower (LL) Louden fields were abandoned from agriculture in 1951 (Burbank et al. 1992). The Bailey (Ba) field site was farmed until ten years prior to this sampling (K.L. Gross, personal communication). The Pond Lab Orchard (PL) and Field K (FK) sites had been abandoned for at least twenty years (Foster 1996). Soils from all six sites were used in the first greenhouse experiment (Chapter 4; Table 1.1, Figure 1.1). Soils from two of the six sites (a high and a low fertility site) were used for the second greenhouse experiment (Chapter 5; sites FK & UL; Table 1.1, Figure 1.1). Ll PL FK Table 1.1. Plant productivity and diversity of abandoned old fields from which soil was collected for the greenhouse experiments. Values for peak aboveground plant biomass (an estimate of primary productivity), species richness, and mean percent organic matter are expressed as mean i standard deviations (n = 6). Values that are significantly different for a given variable based on Fisher’s LSD test are indicated by different letters. Site Dominant Peak Plant Biomass Species diversity Soil Organic Plant Form (standing + litter, g/mz) (#/m2) Matter (%) MK Grass 188 i 16 a 2.2 i- 0.4 a 2.40 i 0.43 a UL Forb 320i23b 15.8:1.0e 3.17:0.23b Ba Forb 424i52c ll.3il.0d 3.03:0.17b LL Grass 432 i 27 c 8.5 i 0.4 c 3.84 i: 0.32 0 PL Grass 480 i 48 c 5.7 i 0.8 b 3.63 i 0.28 c FK Grass 592 i 22 d 1.3 i 0.2 a 3.84 i 0.20 c 7 .mozm omofi montage 2 033. A33 Sumo.“ 8me 35 Eva xv. 2: mo c2802 05 wcumbmszm mecca HEB cosfim Hammo—omm wwozox .M .3 05 «o 932 ._._ onE ‘ 25-532 MP...— .._.._J Fc Cu lor 0hr Pau Ros. Sch] Schn Shari Iicdh TUHHM Zak. J. REFERENCES Burbank, D. H., K. S. Pregitzer, and K. L. Gross. 1992. Vegetation of the W. K. Kellogg Biological Station. Michigan State University Agricultural Experiment Station Research Report Number 510. Foster, B. L. 1996. Plant competition and diversity in relation to productivity in old- field plant communities. Ph.D. Dissertation. Michigan State University, East Lansing, MI, USA. Grayston, S. J. and C. D. Campbell. 1996. Functional biodiversity of microbial communities in the rhizospheres of hybrid larch (LarL'x eurolepis) and Sitka spruce (Picea sitchensis). Tree Physiology 16: 1031-8. Jonasson, S., A. Michelsen, I.K. Schmidt, and E.V. Nielsen. 1999. Responses in microbes and plants to changed temperature, nutrient, and light regimes in the arctic. Ecology 80: 1828-1843. Ohtonen, R., S. Aikio, H. Wire. 1997. Ecological theories in soil biology. Soil Biology and Biochemistry 29: 1613-1619 Paul, EA. and FE. Clark. 1996. Soil Microbiology and Biochemistry. 2nd ed. San Diego, CA: Academic Press, Inc. Rosenzweig, ML. 1995. Species Diversity in Space and Time. Cambridge University Press, Cambridge. Schlapfer, F. and B. Schmid. 1999. Ecosystem effects of biodiversity: a classification of hypotheses and exploration of empirical results. Ecological Applications 9: 893-912. Schmidt, I.K., A. Michelsen, and S. Jonasson. 1997. Effects of labile soil carbon on nutrient partitioning between an arctic graminoid and microbes. Oecologia 112: 557-565. Shaver, GR. and F .S. Chapin, III. 1980. Response to fertilization by various plant growth forms in an Alaskan tundra: nutrient accumulation and growth. Ecology 61: 662-675. Tiedje, J. M. 1995. Approaches to the comprehensive evaluation of prokaryote diversity of a habitat. In Microbial Diversity and Ecosystem Function, eds. D. Allisopp, R.R. Colwell, and BL. Hawksworth. Wallingford, UK: CAB International. Tilman, D. 1996. Biodiversity: Population versus ecosystem stability. Ecology 77: 350-363. Zak, J. C., M. R. Willig, D. L. Moorhead, and H. G. Wildman. 1994. Functional diversity of microbial communities: A quantitative approach. Soil Biology and Biochemistry 26: 1101-1108. Th: 20er prm lntr Com 1‘. influ, MlCrr; strum (Obit) Comm CHAPTER 2 PATTERNS OF DIVERSITY 1N PLANT AND SOIL MICROBIAL COMMUNITIES ALONG A PRODUCTIVITY GRADIENT IN A MICHIGAN OLD-FIELD The following chapter was published as the article: Broughton, LC. and KL. Gross. 2000. Patterns of diversity in plant and soil microbial communities along a productivity gradient in a Michigan old-field. Oecologia 125: 420-427. Introduction A central question in ecology is why there are so many different organisms on the earth (Hutchinson 1959). Much of the work focusing on macroorganisms has emphasized the role of factors such as productivity, disturbance, energy, predation, resources, stochasticity, and colonization in determining the diversity of plant and animal communities (Rosenzweig 1995). Considerably less is known about what factors influence the abundance and diversity of microorganisms (Tiedje 1995). Microorganisms have rarely been incorporated into studies of mechanisms that may structure diversity-productivity relationships for plants and other macroorganisms (Ohtonen et al. 1997, Schléipfer and Schmid 1999). Although plant and soil communities are functionally linked, few studies have examined how patterns of diversity in plant and soil microbial communities co-vary. 10 SO; nit: 5011 C017. COI‘.~ inx'e an i diii (6g Olilt Dri' mic 199 CO] Sui Ihi Soil microbial communities are often limited by carbon (D. R. Zak et al. 1994) or nitrogen (Zak et a1. 1990). Because the extant plant community is usually the main source for both of these resources, the composition and diversity of the soil microbial community may be closely associated with the plant community. It is difficult to assess composition and diversity of soil microbial communities. As a result, most investigations use techniques that assay different aspects of a subset of the microbial community. Two tools that are commonly used by ecologists to characterize the soil microbial cormnunity are the Biolog assay and fatty acid methyl ester (FAME) profiles. Biolog assays sole-carbon-source utilization by the microbial community and provides an index of functional diversity. Biolog profiles have been successfully used to differentiate soil microbial communities associated with different plant communities (e.g. J. C. Zak et al. 1994; Goodfriend 1998), especially when used in concert with other techniques, like fatty acid methyl ester (FAME) profiles (e.g. Buyer and Drinkwater 1997). FAME profiles reflect the phenotypic composition of the soil microbial community (Tunlid and White 1992) and can be used to distinguish among microbial communities with different compositions (Haack et al. 1995; Cavigelli et al. 1995). I investigated the relationship between the structure and activity of the soil microbial community and its relationship to the plant community within an ecologically variable site. I hypothesized that the structure of the soil microbial community would vary at this site in relation to: (1) soil characteristics, (2) plant productivity, and (3) plant diversity. I investigated the relationship between the soil microbial community and 11 (A these three variates along a topographic productivity gradient in a mid-successional old- field in southwestern Michigan. Methods Site Description The study site was in a mid-successional abandoned field at the Lux Arbor Reserve of the W. K. Kellogg Biological Station (KBS) in southwestern Michigan. This site had been abandoned approximately 25 years from agricultural production and during that period had not been grazed, burned, or otherwise managed. Successional fields in this area typically attain a stable species composition of herbaceous perennial 5 to 25 years after abandonment (Huberty et al. 1998). There has been no apparent change in the plant community at this site over the past ten years (K.L. Gross, personal communication). The study site was located along a gentle slope, approximately 15° from the top to the bottom of the hill, along which there were apparent changes in plant species composition and productivity. The soil at the site is Kalamazoo sandy loam soil and does not vary across the study area. Sampling Design and Characterization of the Gradient 1 established five parallel transects, 7.5 m apart, perpendicular to the slope of the hill and sampled soil and vegetation in seven 0.25 m2 plots placed at 10 m intervals along 12 .Ab‘ chp dr\ 7‘. pan seas and bior Ur each transect (n = 7 per transect). To characterize the gradient, I measured light at ground level, soil moisture, soil inorganic nitrogen, and aboveground plant biomass. Aboveground plant biomass and species composition were sampled in June 1996 by clipping the plants at ground level (0 cm above the soil surface), sorting by species, drying at 60°C for 48 hours, and weighing. To better estimate peak plant biomass, particularly at the more mesic end of the gradient, which was dominated by wann- season grasses, the same plots were re-clipped in July 1996. Samples were processed and treated as before. Peak plant biomass was calculated as June biomass + July biomass (both living and standing dead). Light availability at ground level was determined prior to clipping in June and July. Measurements were made at midday (1100-1400 hours EDT) using a Sunfleck PAR Ceptometer (Decagon Devices, Inc.). I measured photosynthetically active radiation (PAR) in full sun 1 m above the plots and took four measurements of PAR at ground level within each plot (cardinal directions). I averaged these four data points to obtain an estimate of the percentage of filll sunlight penetrating to ground level. Soils were also sampled in June and July. For the soil analyses, I aggregated five 2.5 cm diameter by 10 cm deep soil cores taken from each 0.25 m2 plot in an X-shaped pattern. Samples were placed in sealed plastic bags and kept on ice for up to 6 hours until they could be returned to the laboratory. There, they were passed through a 2-mm sieve and sub-sampled for gravimetric soil moisture and nitrogen content within 24 13 llOl Silll‘ fill. hours of sampling. Samples were kept at 4°C until processed. Gravimetric soil moisture was determined by weight loss after drying 10 to 15 g soil at 105°C for 24 hours. For the nitrogen assays, I extracted 20 g fresh soil in 100 ml 1 M KCl. These samples were shaken for 1 minute, settled for 24 hours at room temperature, and filtered through a 1 pm Gelman glass filter. The N03' and NH4+ concentrations of the extracts were determined using an Alpkem Auto-Analyzer. The remaining soil was used to characterize the soil microbial community. Soil for FAME analyses was kept at -20°C until the fatty acids were extracted. Characterization of the Soil Microbial Community I characterized the soil microbial community from samples taken in July using a modified substrate-induced respiration (SIR) method, carbon-source utilization (Biolog) and FAME profiles. SIR assesses the microbial biomass of the soil microbial community and is a good indicator of microbial respiration (Hassink 1993). For SIR microbial biomass, soil slurries were shaken with and without glucose in Erlenmeyer flasks sealed with parafilm, and the headspace C02 was measured. For the control, I combined 25 g soil and 25 ml water in 125 ml Erlenmeyer flasks, and for the glucose- addition I substituted 25 ml 30 mg ml'I glucose for the water. Both sets were shaken for 2 hours at 22°C. After 2 hours, I transferred 5 ml of the headspace gas to a serum vial and measured the initial C02 on an ADC series EGA infrared COz gas analyzer l4 (I1 sha For buf Chi Thn in 0 prcc knox take] oficr Slant Smur Hand aVera refidii Slmlh (The Analytical Development Co. Ltd., Hoddesdon, Herts., UK). The flasks were shaken for another 38 hours and headspace C02 again measured. For the Biolog assay, 1 g of fresh, sieved soil was shaken with 99 m1 of 1% phosphate buffer solution for 20 min and 150 pl of the solution was transferred into each well of a GN Biolog microtiter plate (95 Carbon sources; Biolog, Inc., Hayward, CA 94545). Three replicate plates were inoculated for each plot. The plates were incubated at 25°C in the dark and optical densities were measured after 24 and 48 hours using an Emax precision microplate reader (Molecular Devices Corp., Menlo Park, CA). It is well known that inoculation densities from a standard amount of soil can vary for samples taken from different environments (Konopka et al. 1998). Optical density measures are often standardized to account for differences in inoculation densities; however, the standardizations have been criticized for not accurately reflecting growth across samples with different compositions (Konopka et al. 1998). Therefore, instead of standardizing optical densities, I chose to take advantage of differences and used average well color development (AWCD) from Biolog (corrected within plate for water reading) as an index of microbial respiration. Because profiles at 24 h and 48 h were similar only the results from the 48 h time point are presented here. To obtain fatty acids for FAME analysis, I first extracted the lipids from whole soil samples for 2 h using a mixture of dichloromethane (DCM):methanol:phosphate buffer (1:2:0.8 v/v/v), following a modified Bligh-Dyer procedure (Bligh and Dyer 1959). I then saponified the samples using 1 ml NaOl-l (15% w/v) in methanol (50% v/v) at 15 Me 31011 nUml Cafbc $700; Stan’s T0 01 Comp( 100°C for 30 min and methylated the sample with 2 ml 6N HCl in methanol at 80°C for 10 min. I extracted the fatty acid methyl esters into 1.25 ml (1:1 v/v) methyl-tert—butyl ether-hexane for 10 min and washed the extract with 3 ml 1.2% NaOH. FAME analyses were carried out using a HP 5890 series H gas chromatograph (Hewlett Packard Co., Palo Alto, CA) equipped with a 7673 autosampler and flame ionization detector (Microbial ID Inc., Newark, DL). Peaks were identified by comparison with an external standard. I performed all analyses on the fatty acid proportions of the total peak area to correct for differences in overall peak area. I describe fatty acids using standard nomenclature where the total number of carbon atoms appears before the colon and the total number of CC double bonds appears afier it. Cyclo-propane analogs are indicated by "cyclo," and the location of the epoxy bond is indicated by a "c" followed by two numbers. If the cis or trans configuration is unknown, the word "at" is used. The number following "cis", "trans" or "at" indicates the location of the double bond in relation to the carboxyl end of the molecule. A number before "OH" indicates the location of the hydroxyl group in relation to the carboxyl end of the molecule. Those fatty acids with the same retention time are grouped as "sum in feature" and given a unique number designation. Statistical Analyses To obtain an index of productivity along the gradient I performed a principal components analysis (PCA) on those variates expected to be closely related to 16 ‘fi 1110.1 100 1110 Res (ll 4 productivity: light at ground level, gravimetric soil moisture, soil inorganic N, and peak plant biomass. To examine the productivity-diversity patterns, I regressed plant species diversity, number of carbon sources metabolized (Biolog), AWCD (from Biolog), and SIR microbial biomass against this index of productivity. Changes in plant community composition along the gradient were evaluated with indirect gradient correspondence analysis on species-specific aboveground plant biomass. To visually compare the plant and soil community patterns, I performed K-means cluster analysis on plant species- specific biomass data, carbon source utilization profiles (Biolog), and fatty acid methyl ester profiles (FAME). Multidimensional scaling (MDS) was used to predict the number of clusters expected for the Biolog and FAME profile data. As there were more parameters than samples for the Biolog data, I randomly split the parameters into two subgroups that were run through all analyses independently. The results of these two independent analyses were consistent, so the data from only one is presented. Results Characterization of the Gradient Light at ground level, gravimetric soil moisture, soil inorganic N, and peak plant biomass all co-varied along the topographic gradient. Light availability at ground level (%PAR) decreased from 85% at the crest of the hill to 3% at the base of the hill. Gravimetric soil moisture increased from 15% to 36%, soil inorganic N increased from 17 20>..ch 00:20:08 33 30 30230ch 830: 0.33 mxwmod + $2- .I. a .323 3:933. mm 2.: .20 Vm 0.: 32m :30: 032.2: :00 _ Um 300250: 9:30:28 051 Amv .390 v Q :0 300:3me 203 Gmd n c 03208 :5 Smd u 5 2w: 000 N 00:02 300250: 30:20:00 Sood v a :0 300:2:me 0.53 Cod n c 30803 :00 Aowd H 5 Z Ammo n c 23208 find- u b Em: 0:0 _ .582 300250: 30:29:00 0203833: 30:22, 0.: :0 $2 .2 000 {exec 8.: 330000 N 0:0 2 DA $3803 E20 203 :30 .2 030302 20m 6.3208 :8 205. 030% a £0: 50 05388 82 as; 20.8 <00 a: 800 N 05 _ 00 50 ca .0532 502 2V .2 2&5 05 8% .560 8:320 220.48 2 00 co cm ow on on 3 o o._ m6 o6 We- o. T 0 _ _ T _ 0 P _ _ _ _ _ . 1 v- I r o T .8 o v a one u w. . f m- M m m I. I m.ol d M. D "r. mars—em “cw—mm 7» IA \1/ w - -2 u m % vim Z 2530.: /\ D as U 1 mi \ r We 2332 :8 i < I o._ 18 [a J i) W IP; of r bot: inor‘ has: accu nm’k" grou: addiz 3.29 to 18.58 pg N/ g dry soil, and above-ground plant biomass increased from 89.2 to 309.6 g m'2 along the slope. I performed a PCA to obtain an index of productivity that incorporated these measures of resource availability plus above-ground plant biomass. One sample point at the bottom of the hill was excluded from the analysis because of an abnormally high inorganic soil N value (IO-fold higher than the median). The first principal component based on resource levels measured in June and peak plant biomass (June + July) accounted for 64.8% of the variation in the data set (X = 2.594). Soil moisture, nitrogen, and peak plant biomass were positively correlated with PCI, whereas light at ground level was negatively correlated with PCI (Figure 2.1A). PC2 accounted for an additional 17.1% of the variation, but showed no pattern in relation to the gradient. Therefore, I used PCl as an index of productivity in the remaining analyses (r2: 0.50, Figure 2.1B). A PCA performed on the same variates from the July sampling was indistinguishable from the PCA on the June data, so I will present and use only the June results here. Plant and Soil Microbial Community Relationships with Productivity Plant species richness declined with increasing productivity (PCl) at this site, but productivity accounted for little of the variation in diversity (r2 = 0.17, Figure 2.2A). This relationship was clearly driven by two low diversity points at the high end of the 19 6m u : .2988 0082.800 $3 :8 8888.0: 880: 8:5 .9: 1. fund u a .080: ow 8:: 80880: 8:68 080:8 Em 0:: 80: 8208 80: :80 b: w \U-~OU w: Am: Hm: I 5%: u x .080: ov 8:: .8800 Em 80: 8208 80: E00 b: m CHOU was: as: .o + mad u a .080: we 8:: 8088—08: 880 :03 0m08>0 w0_0_m CV 580: w: 8:: 8:088 :00 0:: 0:: 8288208 80:80 8:80 .8 8:80: 0: H 5: ”025: .. 3.2 u a .0880: 880% 82: 2: 0:0 0:: 8 8:8:8800 3:088 88 82: 0:: 35:08: 82:3 .0880 :8 08:8 53:038.: :0038: 8:30:20: 0: H .N.N 08m; 08:5 £38285 08:5 b38885 vmm_o_-m-m-v. va:o_-m-m-v.. _ . - . r. F . + FL . . . _ . _ . D 1 :5.on .0 D 160v: .0m.. 1 $.oflmm ...N \n} —No"~1~— LN“) «no «a 1v 3 D .0 ..o km :W .33 1» m. hm. 2 m 0m. N.— .... 1 .mA m —W l 0 Q km.- iww m . . _ J _ _ . d . .Ux 08:5 838208: 08::— b_>:0:_005 08:5 538.805 VMN_0_-~-M-Y em~_0_-m-m-v amm_o_-~-m-v Frrb...... ._.._..p. F..:..LL_1N 1 .808 80 1 . 5N N 1 8.0!: ~81”: 1 0m m 1 . ton”: hm h.° 8 .0. o. w.- TC m.- m 4 o o o CVW m I” n 0: m. L .. . . 00m 1. m. v . . .. m. m 2M M coco .0 ”no bwmu 0.. m h. 3 n . .. W m 88 a 4 O O 0 job fiOb EM“ 0 3 00 1 m 00 m Em... q 4 a 4‘4 - - d u u 4 d 1 _4 — d - m 20 produ char (0.25 854 5 hi ghes nucrol Idaho: ”figun nncrol COnSUI Cchn bECaug COIN!) To Vi Sfipar PAM mom (th productivity gradient (Figure 2.2A). However, averaging over transects there was a clear decline in species richness across the gradient: an average of 13 species per plot (0.25 m2) were found at the top of the hill, where productivity was lowest, while as few as 4 species per plot were found at the bottom of the hill, where productivity was highest (Figure 2.2A). Although the number of carbon sources metabolized by the microbial community varied from 23 to 75 across this gradient, there was no relationship between the number of carbon sources metabolized and productivity (Figure 2.2B). However, AWCD at 48 hours, which could be indicative of either microbial respiration or biomass, increased from 0.4 to 1.4 as productivity increased (r2 = 0.52, Figure 2.2C). Similarly, SIR basal activity rate after 40 hours also increased along the gradient (r2 = 0.21, Figure 2.2D), as did the rate at which glucose was consumed between 2 and 40 hours (r2 = 0.44, Figure 2.2B). I used the SIR control and experimental treatments separately in this analysis as measures of microbial respiration because the SIR time course was insufficient to determine microbial biomass. Compositional Shifts in the Plant and Soil Microbial Communities To visually compare patterns in the plant and microbial communities, I performed separate K-means cluster analyses on the plant species biomass data and the Biolog and FAME profiles of the soil microbial communities. A plot of above-ground plant biomass across the study site clearly shows the topographic-productivity gradient (Figure 2.3A) and allows visual comparisons to plant diversity and microbial community measures (Figure 2.3B-D). There were compositional changes in the plant 21 PW“..— Figure 2.3. Results from K-means cluster analysis evaluating the changes in production and composition of the plant and soil communities in relation to transect position at this site. (A) Contour plot of above-ground plant biomass, values are g/ m2; (B) Plant species composition: cluster 1 has no dominant species, cluster 2 is dominated by Rubus sp., cluster 3 by Solidago canadensis, cluster 4 by S. canadensis and Poa pratense, cluster 5 by Agropyron repens, cluster 6 by Poa pratense, and cluster 7 by Polygonum amphibium var. emersum (see Table 2.1 for species lists). (C) BiologTM carbon source utilization profiles separated into two clusters based on AWCD; and (D) FAME profiles: cluster 2 had smaller proportions of 18:1 cis 9, 16:0, and summed in feature 9 (18:2 cis 9, 12 and 18:0 anteiso, Table 2.2) than cluster 1. 22 Distance Down Slnpc (m) Distance Down Slope (m) Distance Down Slope (m) 7.5 Transect (m) 10‘ 20‘ 30-4 0 40 ‘ 60“ " High AWCD :9 s '." '1' I— I l I l 7.5 15.0 22.5 30.0 Transect (m) 15.0 22.5 30.0 {V 1') h— 10- Distance Down Slope (m) 60d 0 I Q r— I I I r I 0.0 7.5 15.0 22.5 30.0 Transect (m) D L l 1 l 4 0— a a Q 9 9 I- 104 E o 20" ’ c. .9 m E 304 o D <1) 8 40-~ £3 2 Q 50— a 1 o .. .— 60— I {32 A -'-'- ‘: I— T I I I I 0.0 7.5 15.0 22.5 30.0 Transect (m) Figure 2.3. 23 comn‘ indlca l comp? analys produt comm n0 €le perenn Agropi Plots ; domin; MDS < relied 1 with n numbC AWCI the [(1 lower There the Sr: community along the productivity gradient. MDS on the species biomass data indicated three to seven valid clusters: I chose to illustrate seven in order to more completely portray the variation in plant composition (Figure 2.3B). The cluster analysis revealed an inverse relationship between plant community diversity and productivity. More specifically, the cluster analysis showed a shift in the plant community from the top of the slope, where there was a mixed community of forbs and no clearly dominant species (Table 2.1, cluster 1), to a mid-slope region dominated by perennial herbs (Table 2.1, clusters 3-4) to a low diversity community dominated by Agropyron repens near the bottom of the slope (Table 2.1, clusters 5-7, Figure 2.33). Plots at the bottom of the hill with highest soil moisture and productivity were dominated by Polygonum amphibium var. emersum (water smartweed) (Figure 2.33). MDS of the Biolog profiles indicated two strong clusters; however, cluster formation relied solely on AWCD and not number or type of carbon sources. This is consistent with the soil microbial community - productivity relationship (Figure 2.2B,C). The number of carbon sources was not related to the productivity index (Figure 2.2B), while AWCD was significantly related to the productivity index (Figure 2.2C). Similarly, in the K—means cluster analysis, sites located at the top of the slope were characterized by . lower AWCD, while the sites at the base of the hill had higher AWCD (Figure 2.3C). There was no difference in the number or types of carbon sources metabolized across the gradient (data not shown). 24 Table shom comir Nome Clustc -1311)” 1(13) "‘1 (6) Table 2.1. Plant species composition in the clusters from the K-means cluster analysis shown in Figure 2.38. Plant species are listed from most common (by biomass) to least common. Only species that have a total biomass > 1 g for the cluster are listed. Nomenclature follows Gleason and Cronquist (1991). Cluster (#plots) 1 (13) 2 (2) 3 (9) 4 (6) 5 (2) 6(1) 7 (2) Plant species Centaurea maculosa, Rubus occidentalis, Hieracium sp., Achillea millifolium, Rubus allegheniensis, Rumex acetosella, Poa compressa, Panicum sp., A gropyron repens, Solidago canadensis, Potentilla recta, Solidago graminifolia, Poa pratense, Aster pilosus, Plantago lanceolata, Phleum pratense. Cerastium vulgatum, Daucus carota, T rifolium pratense. Lespedeza capitata, Dactylis glomerata Rubus occidentalis, Poa pratense. Solidago canadensis, Phleum pratense, Hieracium sp., Rumex acetosella, Agropyron repens, Polygonum amphibium var. emersum, Panicum sp. Solidago canadensis, Rubus occidentalis, Poa pratense, Achillea millifolium, Monardafistulosa, Phleum pratense, Poa compressa, Rumex acetosella, Agropyron repens, Daucus carota, Potentilla recto, Solidago graminifolia, Cornus racemosa, Apocynum cannabinum, Hieracium sp., Lespedeza capitata, Rubus allegheniensis, Centaurea maculosa, T rifolium pratense. Cerastium vulgatum, T araxacum oflicinale, Rumex crispus, Hypericum perforatum Poa pratense, Solidago canadensis, Agropyron repens, Achillea millifolium, Monardafistulosa, Phleum pratense. Aster strigosa, Potentilla recta, T araxacum oflicinale, Galium aparine, Daucus carota, Rumex crispus, Solidago graminifolia, Rumex acetosella Agropyron repens, Solidago canadensis, Monardafistulosa, Polygonum amphibium var. emersum, Rubus occidentalis, Solidago graminofolia, Galium aparine, Poa pratense, Achillea millifolium Poa pratense, Agropyron repens, Solidago canadensis, Achillea millifolium, Polygonum amphibium var. emersum Polygonum amphibium var. emersum, Agropyron repens, Polygonum persicaria, Solidago canadensis, Rumex acetosella, Poa pratense, Poa compressa 25 Altho no dl.‘ used : analys 18:1 C than 3 did n1“ diversi Discus I had h be rel; Becaus index Comnu Comml [he di\ Plant I Although there was variation in the FAME profiles from these samples, MDS showed no distinct clusters in this data set along the productivity-diversity gradient. Fatty acids used in MDS and cluster analyses are listed in Table 2.2. When I forced the cluster analysis to create two clusters, soils from cluster 1 had larger proportions of fatty acids 18:1 cis 9, 16:0, and summed in feature 9 (18:2 cis 9, 12 and 18:0 anteiso, Table 2.2) than soils from the cluster 2 (Figure 2.3D). However, the cluster-based FAME profiles did not show any pattern concordant with peak plant biomass (Figure 2.3A), plant diversity (Figure 2.3B), or Biolog AWCD patterns (Figure 2.3C). Discussion I had hypothesized that the structure of the soil microbial community at this site would be related to plant community diversity, plant productivity, or soil characteristics. Because these three factors covaried at this site (Figure 2.1), I combined them into an index of productivity, but still could not detect any relation to the soil microbial community structure. Neither Biolog nor FAME assays of the soil microbial community were strongly related to variation in productivity. There were changes in the diversity and composition of the plant community associated with soil fertility and plant biomass; however, these differences in plant community composition had no detectable effect on the composition of the soil microbial community. I did find evidence, however, that suggested the respiration (or biomass) of the soil microbial community varied in relation to plant productivity, paralleling the edaphic 26 Tabl F311} T210 11:0 C911 14:0 15:0 15:0 15:1 15:0 16:0 16:1 16:1 16:0 150 1 amei 1?:0 sumr Table 2.2. Fatty acids used in K-means cluster analysis of FAME profiles. Fatty Acid 12:0 17:0 anteiso 11:0 iso 30H 17:1 cis 10 C9 dicarboxylic acid 17:0 cyclo 14:0 18:3 cis 6,12, 14 15:0iso 18:1cis9 15:0 anteiso 18:0 15:1 cis7 19:0 cycloC11-12 15:0 18:0 20H 16:0 iso 20:4 cis 16:1 cis 9 20:0 16:1 cis 11 22:0 16:0 23:0 iso 17:1 G 22:0 20H anteiso 17:1 at 9 24:0 17:0 iso 23:0 20H summed feature 9: 18:2 cis 9, 12; 18:0 anteiso summed feature 10: 18:1 cis 11; 18:1 trans 9; 18:1trans6 27 C0111 1‘ relate can b “V0 differ AWC comn is 815. and p micm resmr Most 3C1'035 PrOdu COmn. Comm Chang along 111% El gradient. I have two lines of evidence that support the idea that soil microbial community respiration (or biomass) increases with plant productivity at this site. First, I detected an increase in AWCD of the Biolog plates in relation to productivity (Figure 2.2C). Although Biolog AWCD is not a direct measure of respiration, it is strongly related to inoculum density (Garland and Mills 1991, Haack et al. 1995) and, as such, can be interpreted as an indicator of total number of bacteria (biomass). Conversely, two wells with the same inoculation density may differ in AWCD because of differences in microbial respiration (Konopka et al. 1998). In either case, the higher AWCD in the sites at the base of the hill indicates a more productive microbial community and this corresponds to areas along the gradient where the plant community is also the most productive. This is consistent with the higher amounts of N, moisture, and plant biomass at the base of the hill, which should make more C available to the microorganisms. Secondly, the modified SIR analysis indicates higher rates of respiration at the base of the hill where productivity was highest (Figure 2.2D,E). Most studies that have reported changes in soil microbial community composition across community types have sampled sites that differed in plant species composition, productivity, and soil type. From these studies, it is unclear whether the plant community or the underlying edaphic factors are influencing the soil microbial community structure. For example, J. C. Zak et al. (1994) used Biolog to investigate changes in functional diversity of the soil microbial community from grasslands located along an elevational and moisture gradient in New Mexico. They found differences in the Biolog profiles of the soil microbial community from six distinct plant communities 28 dong sues. [flant Goodl cmnm LautC< edaph Bkflcu SCVCTL acnls] dlSIlnE 1n the (180551 mICrok along this gradient. However, because soil characteristics also varied among these sites, it is not clear whether the differences in Biolog profiles were due to changes in plant community composition, edaphic factors, or some other variable. Similarly, Goodfn'end (1998) found that Biolog distinguished among the soil microbial communities of eight sites representing a variety of wetlands in the southwestern United States. However, it was not clear whether plant community composition or edaphic characteristics were more important in influencing the grouping of those Biolog profiles into habitat types. Several authors have argued that phospholipid fatty acids (PLFA’s), a subset of fatty acids present in the phospholipid membrane, may provide a more sensitive indicator to distinguish among microbial communities. Phospholipid fatty acids break down easily in the soil and are thus thought to represent the active soil microbial community (Bossio and Scow 1998). Zelles et al. (1992) used PLFA profiles to compare soil microbial community patterns in grassland and agricultural fields under different management regimes and found that profiles differed among the different fields, but they did not distinguish between plant community and edaphic effects. Bossio et al. (1998) concluded that soil type has stronger effects on the soil microbial community structure than plant community type. They found that the addition of a cover crop (an increase in plant community diversity over time) was less influential in changing PLF A profiles than soil type. The differences in edaphic characteristics at this site, although substantial, were not as striking as differences between soil types would be. 29 There are several possible reasons why I did not detect changes in the soil microbial community composition along this gradient: (1) there is no connection between the structure of the soil microbial community and the soil characteristics, plant diversity, or plant productivity; (2) the soil microbial community structure is very stable and affected mainly by factors like long-term plant community composition or historical C inputs to the soil; (3) the spatial scale or time of year I sampled was inappropriate for detecting differences in the soil microbial community; or (4) the techniques I used to assay the soil microbial community were not specific enough to detect what differences were there. The first two reasons seem unlikely because there should be a linkage between the microbial (consumer) community and the resources (plant carbon) that they utilize (Paul and Clark 1996). Much of the carbon available to soil microorganisms is being provided to the soil microorganisms each year by the extant plant community, and although this is a successional community, the plant community composition at this site has remained stable for the past decade (K. L. Gross, personal communication). Even if soil microbial community structure is not affected by plant community composition, increasing plant diversity or productivity should provide additional resources to the extant soil microbial community and thus influence soil microbial community composition. Additionally, past agricultural use at this site likely would have depleted soil C (Drinkwater et al. 1998, Robertson et al. 1993), and therefore made the current community inputs of C important in determining the structure and activity of the soil microbial community. 30 It is NEW: temp chos plant reve; mid-1 samp carbc rhizo It is p “fight Westr T111205 and C 01 rhi Hon-e 5133113 more Produ, d1 1TB 1' c It is difficult to know if sampling at a different time of year or spatial scale would have revealed associations between the plant and microbial communities at this site. Both temporal and spatial scales are important in the observation of ecological phenomena. I chose to sample in mid-summer on the assumption that at this time of year both the plant and soil microbial communities would be most active. Other researchers have revealed associations between plant and microbial communities using soil sampled in mid-summer. Bossio et al. (1998) detected differences in PLFA patterns of soil sampled in July from different agricultural treatments in California. Similarly, using carbon source utilization patterns, Westover et al. (1997) differentiated among rhizosphere soils sampled in August from several grass species in Washington. It is possible that if this sampling had been done at a smaller, more fine-grained scale I might have detected associations between microorganisms and specific plant species. Westover et al. (1997) detected differences among soil microbial communities of rhizosphere soils of several grass species in both the field and greenhouse. Grayston and Campbell (1996) used Biolog to differentiate between the microbial communities of rhizosphere soils from two tree species, Larix eurolepis and Picea sitchensis. However, others have found associations between plant and microbial communities at spatial scales similar to the scale used in this study. Plant community composition is more likely to affect soil microbial community composition than plant diversity or productivity. A recent experimental study by Wardle et al. (1999) did detect differences in PLFA composition of the soil microbial community that were 31 The hnnta there to dist used 1 1995; lechni Closet: significantly related to the plant removal treatments. This suggests that 3-4 years of abandonment is sufficient to detect changes in the soil microbial community. Broughton et al. (2001) saw a similar relationship between plant community composition and PLFA patterns of the soil microbial community at the Silwood Park BIODEPTH site after three years (results presented in Chapter 3). The inadequacy of tools to assess microbial diversity has been a long-standing limitation to this understanding of soil microbial communities (Tiedje 1995). While there are clearly limitations to the ability of fimctional tools such as Biolog and PLF A to distinguish among microbial communities, as noted above, a number of studies have used these tools to successfully differentiate among communities (Zelles et al. 1992, 1995; J. C. Zak et al. 1994; Goodfriend 1998). The development of molecular techniques more sensitive to shifts in composition may reveal natural shifts from one closely related microorganism to another along a gradient, just as there are shifts among closely related plant species along gradients. Additionally, molecular techniques may allow us to better address the roles of dominance and plasticity in structuring soil microbial communities. The correlative nature of this study does not allow us to determine what factors may underlie the observed variation in the soil microbial community at this site. Despite our expectation that there should be a close association between the plant community and the soil microbial community, at this spatial scale (within a site), using these tools, I were not able to detect any association between plant community composition and soil 32 micr and that patte microbial community composition. The similarity between patterns of plant biomass and soil microbial respiration is intriguing, however, and suggests that the resources that limit each of these communities co-vary. In contrast, the differences between patterns of plant diversity and soil microbial community structure suggest that different mechanisms are responsible for structuring diversity in these associated communities. 33 Brou Buye Cavh h \ Dnnk Gafla GOOd] Bra-VS Haack REFERENCES Bligh E. G., W. J. Dyer. 1959. A rapid method of total lipid extraction and purification. Canadian Journal of Biochemistry and Physiology 37: 911-917 Bossio, D. A. and K. M. Scow. 1998. Impacts of carbon and flooding on soil microbial communities: Phospholipid fatty acid profiles and substrate utilization patterns. Microbial Ecology 35: 265-278. Bossio, D. A., K. M. Scow, N. Gunapala, and K. J. Graham. 1998. Determinants of soil microbial communities: effects of agricultural management, season, and soil type on phospholipid fatty acid profiles. Microbial Ecology 36, 1-12. Broughton, L.C., K.L. Gross, and A. Hector. 2001. Linking plant community diversity to soil microbial communities: an experimental evaluation from the BIODEPTH experiment. Journal of Ecology (submitted). Buyer, J. S. and L. E. Drinkwater. 1997. Comparison of substrate utilization assay and fatty acid analysis of soil microbial communities. Journal of Microbiological Methods 30: 3-11. Cavigelli M. A., G. P. Robertson, and M. J. Klug. 1995. Fatty acid methyl ester (FAME) profiles as measures of soil microbial community structure. Plant and Soil 170: 99-113 Drinkwater L. E., P. Wagoner, and M. Sarrantonio. 1998. Legume-based cropping systems have reduced carbon and nitrogen losses. Nature 396: 262-265 Garland J .L. and A. L. Mills. 1991. Classification and characterization of heterotrophic microbial communities on the basis of patterns of community-level sole-carbon- source utilization. Applied and Environmental Microbiology 57: 2351-2359 Gleason H. A. and A. Cronquist. 1991. Manual of Vascular Plants of Northeastern United States and Adjacent Canada. The New York Botanical Garden, New York. Goodfiiend W.L. 1998. Microbial community patterns of potential substrate utilization: a comparison of salt marsh, sand dune, and seawater-irri gated agronomic systems. Soil Biology and Biochemistry 30: 1169-1176 Grayston SJ. and C. D. Campbell. 1996. Functional biodiversity of microbial communities in the rhizospheres of hybrid larch (Larix eurolepis) and Sitka spruce (Picea sitchensis). Tree Physiology 16: 1031-8 Haack, S.K., H. Garchow, M.J. Klug, and L. J. Fomey. 1995. Analysis of factors affecting the accuracy, reproducibility, and interpretation of microbial community carbon source utilization patterns. Applied and Environmental Microbiology 61: 1458-1468 Hassink J. 1993. Relationship between the amount and the activity of the microbial biomass in Dutch grassland soils: Comparison of the fumigation-incubation method and the substrate-induced respiration method. Soil Biology and Biochemistry 25: 533-538 34 Huberty L. E., K. L. Gross, and C. J. Miller. 1998 Effects of nitrogen addition on successional dynamics and species diversity in Michigan old-fields. Journal of Ecology 86: 794-803 Hutchinson GE. 1959. Homage to Santa Rosalia or Why are there so many different kinds of animals? American Naturalist 93: 145-159 Konopka A., L. Oliver, and R. F. Turco, Jr. 1998. The use of carbon substrate utilization patterns in environmental and ecological microbiology. Microbial Ecology 35: 103-115 Ohtonen R., S. Aikio, and H. Wire. 1997. Ecological theories in soil biology. Soil Biology and Biochemistry 29: 1613-1619 Paul, EA. and FE. Clark. 1996. Soil Microbiology and Biochemistry. 2nd ed. San Diego, CA: Academic Press, Inc. Robertson G. P., J. R. Crum, and B. G. Ellis. 1993. The spatial variability of soil resources following long-term disturbance. Oecologia 96: 451-456 Rosenzweig, ML. 1995. Species Diversity in Space and Time. Cambridge University Press, Cambridge. Schléipfer, F. and B. Schmid. 1999. Ecosystem effects of biodiversity: a classification of hypotheses and exploration of empirical results. Ecological Applications 9: 893-912. Tiedje, J. M. 1995. Approaches to the comprehensive evaluation of prokaryote diversity of a habitat, pp. 73-87, In Microbial Diversity and Ecosystem Function (eds D. Allisopp, R.R. Colwell, and D.L. Hawksworth). CAB International, Wallingford, UK. Tunlid, A., and DC. White. 1992. Biochemical analysis of biomass, community structure, nutritional status, and metabolic activity of microbial communities in soil, pp. 229-262, In Soil Biochemistry, vol. 7 (eds G. Stotsky and J .M. Bollag). Wardle, D.A., K.I. Bonner, G.M. Barker, G.W. Yeates, K.S. Nicholson, R.D. Bardgett, R.N. Watson, and A. Ghani. 1999. Plant removals in perennial grassland: vegetation dynamics, decomposers, soil biodiversity, and ecosystem properties. Ecological Monographs 69: 535-568. Westover, K. M., A. C. Kennedy, and S. E. Kelley. 1997. Patterns of rhizosphere microbial community structure associated with co-occurring plant species. Journal of Ecology 85: 863-873. Zak D. R., D. R. Gri gal, S. Gleeson, and D. Tilman. 1990. Carbon and nitrogen cycling during old-field succession: constraints on plant and microbial biomass. Biogeochemistry 11: 111-129 Zak D. R., D. Tilman, R. R. Parmenter, C. W. Rice, F. M. Fisher, J. Vose, D. Milchunas, and C. W. Martin. 1994. Plant production and soil microorganisms in late-successional ecosystems: a continental-scale study. Ecology 75: 2333- 2347 35 Zak, J. C., M. R. Willig, D. L. Moorhead, and H. G. Wildman. 1994. Functional diversity of microbial communities: A quantitative approach. Soil Biology and Biochemistry 26: 1101-1108. Zelles, L., Q.Y. Bai, T. Beck, and F. Beese. 1992. Signature fatty acids in phospholipid fatty acids and lipopolysaccharides as indicators of microbial biomass and community structure in agricultural soils. Soil Biology and Biochemistry 24: 317-323. 36 CHAPTER 3 LINKING PLANT COMMUNITY DIVERSITY TO SOIL MICROBIAL COMMUNITIES: AN EXPERIMENTAL EVALUATION FROM THE BIODEPTH EXPERIMENT These results have been submitted to Journal of Ecology in an article: Broughton, L.C., K.L. Gross, and A. Hector. 2001. Linking plant community diversity to soil microbial communities: an experimental evaluation from the BIODEPTH experiment. Journal of Ecology (submitted). Introduction Most studies to date investigating the relationship between species diversity and ecosystem function have focused on how changes in primary producers and consumers affect ecosystem processes (Schlapfer and Schmid 1999, Rosenzweig 1995). The interaction between above-ground and below-ground (soil) communities in mediating these processes has been less studied (Schlapfer and Schmid 1999, Ohtonen et al. 1997). While considerable attention has been paid to factors that affect the composition and function of communities of macroorganisms, very little is known about the factors that affect the structure of soil microbial communities (Ohtonen et al. 1997, Tiedje 1995). Soil microorganisms play an essential role in the functioning of terrestrial ecosystems because the soil microbial community provides key steps in the cycling of nutrients 37 31; prc pen C011. 0111. may or d llllCr. C0mr €COS} T0 dc C0mm $thm (carbon, nitrogen, and phosphorus, among others) through the ecosystem (Paul and Clark 1996). In most temperate grassland and forest systems, plant growth is limited by nitrogen (or a combination of nitrogen and phosphorus, Shaver and Chapin 1980, Schmidt et al. 1997, Jonasson et al. 1999). The soil microbial community controls the release of inorganic nitrogen to plants; however, the soil microbial community is most often limited by carbon. Therefore, the rate at which limiting nutrients are made available to plants is likely to be influenced by both the amount and quality of carbon provided by plants and available to soil microorganisms. Because plant species differ in carbon content and quality, plant species identity has the potential to affect nutrient process rates through litter quality effects, which consequently affect the soil microbial community (Paul and Clark 1996, Wardle and Giller 1996). As a result, the composition and productivity of the plant community may influence the soil microbial community. Similarly, differences in the productivity or diversity of the plant community may be affected by processes mediated by soil microorganisms (e.g. N-mineralization rates). Consequently, changes in the plant community and the resulting change in the soil microbial community potentially affect ecosystem function. To determine the effect of plant community diversity on soil microbial community diversity and processes, I studied the soil microbial community in experimental plant communities at the BIODEPTH site in Silwood Park, UK, where plant community structure and diversity were experimentally manipulated. I asked the following 38 questions: (1) How does plant community diversity affect the structure of the soil microbial community? (2) How does plant community composition affect the structure of the soil microbial community? (3) How does plant productivity affect the relationship between plant community diversity and the soil microbial community? and (4) Do specific functional groups or plant species have detectable effects on the soil microbial community? Materials and Methods Site The study was conducted at the Imperial College site at Silwood Park, Ascot, UK (National Grid Reference 51°22’N, 00°37’W) and was part of the BIODEPTH experimental network of sites (BIODiversity and Ecosystem Processes in Terrestrial Herbaceous systems, Hector et al. 1999). The site was previously used for horse- grazing and has sandy-loam soil with an average pH of 5.26. In Fall 1995, the field was fenced, herbicided (Round Up, Dow Elanco), and tilled (Hector et al. 2000). The soil was fumigated in April 1996 with methyl bromide (Check Fumigation Ltd., Reading, UK) to remove the soil seed bank. Fumigation should also have killed much of the soil microbial community. In May 1996, two replicate blocks each with 33 plant assemblages, plus no-plant controls, were established in 2 x 2 m plots. The assemblages consisted of different combinations of plant species that varied in species richness and firnctional group richness (Hector et al. 1999). The species sown were all 39 herbaceous perennials representative of grassland species common to this part of England. Maximum species richness was 11 species per 2 m x 2 m quadrat, the average number of species in the area. Plant species were classified into one of three functional groups: legumes, non-leguminous forbs, and grasses. To address concerns about individual species effects on species richness curves (see Huston 1997), species thought to have strong effects on productivity were included in all mixtures. To minimize plant biomass effects, all mixtures included at least one grass species. This limits our ability to detect the effects of individual plant species and to evaluate the effects of grasses. Plant assemblages were maintained by hand-weeding for all 4 years of the experiment. Several undisturbed reference plots were also established adjacent to the manipulated plots. Peak plant biomass was clipped 5 cm above-ground level both years to provide an estimate of annual net primary productivity (Hector et al. 1999). Soil Sampling and Analysis To determine the relationship between plant diversity and the soil microbial community, I selected a subset of plots, encompassing the full range of species and functional group diversity, to sample for soil and microbial characteristics. In October of year 3 of the experiment (1998) I sampled soils from 28 plots (2 replicates of 14 different plant compositions, Table 3.1). The following year (September 1999, year 4), I sampled soils from 36 plots (2 replicates of 18 different plant compositions, Table 3.1). These included 12 of the 14 plots sampled in year 3, plus an additional 6 plots to expand the coverage of the species richness gradient. 1 were unable to re-sample the 40 Tab and $8.111I lo: ih bl the - pres (a) £58" :1 gl‘r’ ..~1 / 0f: prartl Ant/i odor. Arr/r clam C 1 Hr Cl‘lSlr. Dar! 1 F8511! H()/( I. Liam“. P/l/t’li’ Trr’sor flares. Table 3.1. (A) Plant species pool used in establishing the experimental communities and (B) the plant communities sampled for the soil microbial community. Plots sampled only in year 3 (1998) of the experiment are in italics. The control and Rumex plots were not maintained through year 4 and could not be re-sampled. Plots sampled in both years are indicated in bold. Other plots were sampled only in year 4 (1999) of the experiment. Communities are grouped by number of plant functional groups (F G) present. (a) Grasses Abbrev. Legumes Abbrev. Forbs Abbrev. Agrostis capillaris AgC Lotus LC Achillea AM Alopecurus AP corniculatus millefolium pratensis Medicago ML Cerastium CF Anthoxanthum AO lupulina fontanum odoratum T rifoli um TR Hypochaeris HR Arrhenatherum AE repens radicata elatius T rifoli um TP Plantago PL Cynosurus CC pratense lanceolata cristatus Vicia hirsuta VH Potentilla PE Dactylis glomerata DG V icia saliva VS erecta F estuca rubra FR Vicia VT Rumex acetosa RA Holcus lanatus HL tetrasperma Stellaria SG Luzulla campestris LC graminea Phleum pratense PhP T araxacum TO T risetum TF oflicinale flavescens Veronica VC chamaedrys 41 Table 3.1 (cont’d). (b) E E E Trt 0 g Trt 1 FG g Trt 2 FG g Trt 3 FG Code PG 5 Code i Code 3 Code Co co {1 AB 1 15 AE, LC lRer reference n- E 2 FR E 1 plots tro 5 3 HR 3 16 AE, FR, LC, : ls g 4 RA g TR g 11 AE, TR, i 5 LC 1 5 RA, HR 56 TR 3 18 AgC, AE, FR, 5 5 5 HL, AM, HR, 5 12 FR,AE, :13 AgC, FR, ; RA, PL ; LC, TR, g HL, AE g ; HL, RA, 1 a 17 AgC, AE, FR, 5 PL, AM 3 14 AgC,AE, 3 HL, TR, LC, 3 5 FR, HL, 2 TP, vs 5 19 HL, AgC, § AP, A0, 2 3 LC, TR, FR, g CC, TF g 8 AgC, AE, ; PL, RA, HR 5 i FR.HL,RA, i 5 7 AgC,AE, 3 PL, AM, so, 310 AgC,AE, ; FR,HL, ; VC, PE, TO 5 FR,HL, E AP,AO, 3 : LC,TR. g CC, TF, g 9 AgC, AE, 5 AM, CF, 5 DG, PhP, 1 FR, HL, TR, 5 HR, PL, 3 LZ 3 LC, TP, vs, 5 RA ; ; ML, VH, VT 3 42 control (no plant) and Rumex monoculture plots because the plots were not maintained in the fourth year. In year 4, I focused on a subset of the soil and microbial variables. In both years the sampled plots included undisturbed reference plots and combinations of 4, 8, and 11 species varying from 1 to 3 functional groups. I sampled soil to a depth of 10 cm, then sieved the sample through a 3.35 mm sieve, and stored it in sealed plastic bags at 4°C until analyzed. All analyses were done at the WK. Kellogg Biological Station of Michigan State University within 3 days of sampling. I determined gravimetric soil moisture for each sample by drying 10 g soil at 105°C for 48 hours (Nelson and Sommers 1982). A subsample of the dried soil was ashed at 500°C for 4 hours to determine organic matter content (Nelson and Sommers 1982). Soil pH was determined using a Corning pH meter 420 after mixing 5 g of air-dried soil in 50 ml millipure H20. For nitrogen analyses, I extracted 20 g of fresh soil in 100 ml 1M KCl. The samples were shaken for 1 min and allowed to settle for 24 h at room temperature. The supernatant mixture was filtered through a l-um Gelman glass-fiber filter and NO3' and NH4+ concentrations were measured using Alpkem auto-analyzer. To determine potential N-mineralization and nitrification rates, a companion 20 g sample was incubated for 21 days at 25°C and 15% humidity and then extracted using the same methods as above. 43 1 determined microbial biomass using the chloroform fumigation incubation method (Paul et al. 1999b). Two 25 g soil samples were pre-incubated for 5 days then one sample was fumi gated with chloroform for 24 hours to kill the microorganisms. After a vacuum was created and the chloroform evaporated, 0.5 g of original soil was added to both samples. I measured initial headspace C02 and accumulated C02 after 10 days on an ADC series EGA infrared C02 gas analyzer (The Analytical Development Co. Ltd., Hoddesdon, Herts., UK). I calculated microbial biomass as [ 1.73 * (10 day accumulated COz-C — initial COz-C for the fumigated samples) — 0.56 * (10 day accumulated COz-C — initial COz-C for the control samples)] (Paul et al. 1999b). To determine microbial respiration I used a separate set of 10 g soil samples that were pre- incubated 5 days in a 160 ml glass qorpak bottle. I measured initial headspace CO; and accumulated C02 after 1 and 5 days and calculated the rate of COz-C respired per day. Community-level physiological profiles (CLPP) were determined using both Biolog and Ecolog plates (Biolog, Inc., Hayward, Calif, USA). For both assays, 1 g of fresh, sieved soil was shaken with 99 ml 1% phosphate buffer solution for 20 min. 150 1.11 of the mixture was transferred to each well of the microtiter plate (GN Biolog, 95 Carbon sources + 1 non-Carbon control; or Ecolog, 3 replicates of 31 Carbon sources + 1 non- Carbon control). The plates were incubated in the dark at 25 °C and optical densities were measured at 14 h intervals from 0 h to 64 h using an Emax precision microplate reader (Molecular Devices Corp., Menlo Park, Calif, USA). Because the 5 incubation 44 tin: 165'. CW: (VarI polar phos met pl 0\ times gave consistent results and the Biolog and Ecolog plates were similar in their results, I present here data only from the 64 h Ecolog measurements. For the PLFA analysis, I extracted lipids from 6 g whole soil samples for 2 h using a mixture of dichloromethane (DCM): methanol: phosphate buffer (1:2:0.8 v/v/v), following a modified Bligh-Dyer procedure (Bligh and Dyer 1959). Phase separation was achieved by adding DCM and saturated sodium bromide solution (1:4 v/v). I isolated the phospholipid fatty acids from the dried lipid extracts by solid phase extraction. The lipid material was added to a polar column consisting of 100 mg silica (Varian Bond Elut LRC Columns, Product # 1211-3010). Lipids of low or intermediate polarity were eluted with chloroform and acetone and discarded. Subsequently, phospholipid fatty acids were eluted with 1.5 m1 methanol for preparation of fatty acid methyl esters. I saponified the samples using 1 m1 NaOH (15% w/v) in methanol (50% v/v) at 100°C for 30 min and methylated the samples with 2 m1 6M HCl in methanol at 80°C for 10 min. I extracted the fatty acid methyl esters into 1.25 ml (1:1 v/v) methyl- tert-butyl etherhexane for 10 min and washed the extract with 3 ml 1.2% NaOH. Phospholipid amounts were measured using a HP 5890 series 11 gas chromatograph (Hewlett Packard Co., Palo Alto, Calif, USA) equipped with a 7673 autosampler and flame ionization detector (Microbial ID Inc., Newark, Del., USA). Peaks were identified by comparison with an external standard. I performed all analyses on the phospholipid fatty acid proportions of the total peak area to correct for differences in overall peak area. 45 ator C it indi. unlo the 1 acid: numl Star 13;. in I describe fatty acids using standard nomenclature where the total number of carbon atoms appears before the colon and the total number of C-C bonds appears after it. Cyclo-propane analogs are indicated by "cyclo," and the location of the epoxy bond is indicated by a "c" followed by two numbers. If the cis or trans configuration is unknown, the word "a " is used. The number following "cis, trans," or "a " indicates the location of the double bond in relation to the carboxyl end of the molecule. Fatty acids with the same retention time are grouped as "sum in feature" and given a unique number designation. For analysis, I included only those phospholipid fatty acids that were present in all samples and reported their abundance as the proportion of the total phospholipid fatty acid amount in each sample. Of the 25 lipids detected, 16 phospholipid fatty acids met this criterion in both years. Statistical Analysis Using an analysis of variance (ANOVA) model and sequential (Type 1) sums of squares, I tested the effect of number of species (richness), number of functional groups, block, and plant community composition (MIXTURE) on the following response variables: pH, soil moisture, soil organic matter, total N, N-mineralization rate, nitrification rate, number of culturable bacteria, microbial respiration, microbial biomass, CLPP, and PLFA profiles (Table 2). The effect of number of species was tested separately from the effect of number of functional groups; therefore, both tests were non-conservative. Hector et al. (1999) found that aboveground plant biomass increased with increased numbers of plant species at this site (and most of the other 46 1211. the par. Son 11.: 810 Oh i 1111' Bio [111 lune legit F P111): I Plan‘ 8101‘ Dit't‘ Mitt ago Table 3.2. Analysis of variance model with sequential sums of squares used to evaluate the effects of the plant community manipulations on soil and microbial community parameters. Source of variation Mean square Variance ratio Main GLM model (ANOVA): BIOCK M83 M83/ MSBtM Diversity MSD MSD / MSM Mixture MSM MSM / MSBI-M Block*Mixture MSWM -- [Where Diversity is 1) species diversity, 2) functional group diversity, 3) presence/absence of legumes, or 4) presence/absence of forbs] Plant Biomass as a covariate (ANCOVA): Plant Biomass (covar) MSC MSC / MSM BIOCK M83 M83/ MSBt-M Diversity MSD MSD / MSM Mixture MSM MSM / MSBtM Block*Mixture MSB-M -- BIODEPTH sites, species number, p<0.001; functional group number, p<0.01). Consequently, I used aboveground plant biomass as a covariate in these analyses to investigate the influence of plant diversity independent of plant biomass effects (Table 3.2). I used Principal Components Analysis (PCA) to determine if there was an underlying structure to the soil microbial community as detected by CLPP and PLFA. I used these principal component axes as response variables in the ANOVA and ANCOVA to test for main treatment effects of plant species richness, functional group richness, and community composition on soil microbial community metabolic activity and structure. 47 Results Legacies of Site Preparation The effect on the soil microbial community of soil fumigation with methyl bromide during site preparation is clearly illustrated by the soil microbial biomass data. In both years, the undisturbed (non-fumigated) reference plots had the highest microbial biomass, as much as four times higher than the manipulated plots (Figure 3.1). Even after 4 years (1999 sampling), the biomass of the soil microbial communities had not recovered from the disturbance effect of fumigation (Figure 3.1B). Effects of Species Richness and Functional Group Richness Treatments on the Soil Microbial Community The initial analyses of these data excluded aboveground plant biomass as a covariate and I found little effect of species richness or functional group richness on soil or microbial community parameters (Table 3.3A). There were a few exceptions. In year 3, the numbers of colony forming units (CFU’s) after 48 hours increased at higher species richness (Figure 3.2A, r2=0.29, p< 0.01). Similarly, number of CFU’s after 48 hours increased at higher functional group richness (Figure 3.2B, r2=0.29, p< 0.01). However, neither of these relationships was significant when aboveground plant biomass was used as a covariate in the year 3 analysis (Table 3.3A). The number of culturable bacteria was significantly correlated with aboveground plant biomass in year 48 9.35-53: 3:: 33.35 atmcfiombfl 3:: 35 $07. .8353: 3 3:ng :5 33:6 3 @8365 2: 85:28:02 28m 3 cap—omega: 8: 33:36:50 00:20.3: 333358-:0: 223 .286 x52: : E “68:88:“: 2: 32: 65:8 22902 .m:.m 038. Eat 2: 3:3 5:85:52 €55.88 Ema .mm _ H :32 2: 82:5 .353 Imam—DOE fem vooB=m of no v SPA Am: 2.: m :8» A5 E 3:62: 3595:: :8 :0 float» 52895: $53888 Ema ._.m “:sz arm—.5923? $3.2m m N o m m N 5mm +fi45511fi71___ : : a m A. m 4 .W m - m LF. fiflfififlfil m 1&1 .L I J Zoom- .oow- .o .oow CON oom .oov com com ([Ios Kip 8 /;) 8n) ssequIq [etqorotw NF :owm m N. o m e m Ntomoo fififififir 4- 0 OO O \— OF. .IN 0) no N 8 L0 V ([105 Kip 3 /3 Bit) sseurorq lerqomr 49 Table 3.3. Significance and direction (for diversity) of treatment effects on soil and microbial parameters for both years as detected by Analysis of Variance using Type I sums of squares (model in Table 3.2) for (a) species diversity, functional group diversity, and composition and (b) presence/absence of legumes or forbs. ANCOVA results using above-ground plant biomass as the covariate are shown in parentheses if the effect changed in significance. NS=not significant (a) Plant Species Diversity F G Diversity Composition Variable: Biomass Year 3 pH <0.05 NS NS < 0.05 (< 0.05) Soil moisture <0.10 + < 0.01 (< 0.05) NS NS Soil organic matter NS NS NS NS Total N NS NS NS NS N-mineralization NS NS NS NS rate Nitrification rate NS NS NS < 0.05 (NS) Culturable bacteria + <0.05 + < 0.05 (NS) + < 0.05 (NS) NS Microbial NS NS NS < 0.01 (< 0.01) respiration Microbial biomass NS NS NS NS CLPP PC] + <0.01 + < 0.05 (NS) + < 0.05 (< 0.10) NS CLPP PC3 NS NS + < 0.10 (<0.05) NS PLFA PCl - <0.05 NS NS NS PLFA PC2 NS - <0.05 (<0.10) NS NS PLF A PC4 NS NS NS <0.05 Year 4 Soil moisture NS NS NS NS Total N + <0.10 - <0.10 (NS) NS NS N-mineralization NS NS NS NS rate Nitrification rate NS NS NS NS Microbial NS NS NS NS respiration Microbial biomass NS NS NS NS PLFA NS NS NS NS 50 Table 3.3 (cont’d). (b) Variable: Plant Biomass +/- Legumes +/- Forbs Year 3 pH NS NS NS Soil moisture + <0.05 NS + <0.10 (<0.10) Soil organic matter NS NS NS Total N NS NS NS N-mineralization rate NS NS NS Nitrification rate NS NS NS Culturable bacteria + <0.01 NS + < 0.05 (NS) Microbial respiration NS NS + < 0.05 (<0.05) Microbial biomass NS NS + < 0.05 (<0.05) CLPP PC4 NS + < 0.05 (0.10) + < 0.10 (< 0.10) CLPP PCS NS NS - < 0.01 (<0.01) PLF A NS NS NS Year 4 Soil moisture NS NS NS Total N NS + <0.05 (<0.05) NS N-mineralization rate NS + <0.05 (<0.10) NS Nitrification rate NS NS NS Microbial respiration NS NS + <0.10 (<0.10) Microbial biomass NS NS NS PLFA NS NS NS Table 3.4. Fatty acids used in principal components analysis of PLF A profiles. Phospholipid Fatty Acids C9 Dicarboxylic acid 14:0 15:0 iso 15:0 anteiso 16:0 iso 16:1 cis 9 16:1 cis 11 16:0 17:0 iso 17:0 anteiso 17:0 cyclo 18:2 cis 12 18:1 cis 9 18:0 19:0 cyclo c11-12 summed feature 8: 18:1 trans 9 51 .339: 2: 36on 50: 5:3 32: 03: of. .Ewm: on: 9 :30: wEEEm 3.38.53: 02:: 88:8: atmgoomxt 02:: :9: 32a ”Em: on: o: a: @352: 358.53: 96; 2:339:58 32:: 3:: 8...: 32m .mo_w::_:: B 3:238 28 83:6 :3 38:65 0:: 85:30:32 9:3 3 325850: 0:: mo_:_::EEo: 858.3: “BEE—€278: 223 520:: :23: E 820850: 0:: 82: .0550 233-02 .Codvq .mnmdnmmv m :3» E 3:805 :52: 33833:: ADV cc: .Codv: .wwmdnemv 820% 3:285: ES: :0 :onEz: Am: fodva .vomdummv 33:2: 36on :52: Ao:< cog 8o— oow coo co: com o anew 3:285: mo :38: Z 860% Ema? 82:32 m N _ o N: w e o _ ~ ~ — i1 — Mul O _ q _ _lO _ q _ :10 a o W 4 o N o w 4 00— w .1 4 O m 0 O l e e. e )m m 18:)m m 18:) 4 co 3 4 . J 3 4 o 13 ma 4 ® mm 0 a 4 o 13 q 1:8 n m u n .1... n H m. . 108x w e . roamx 4 0 u 4 I o l. ‘ 9mm. ‘ 0. m. a Or.) 18: (m a... m ( a w room w room mm 4 ova u 1:9. u. m 4:8 < 18:. 1131; re saguoloojo iaqum N 52 3 (F 1; numk With } I numb when plots E ffcct Becau 0f prc pararm moistu year 3. diSIingl 0f ’egUfl and N- j The Sm; Very few 3 (Figure 3.2C, r2: 0.57, p< 0.01). PLF A PC2 was also significantly related to species number (Table 3.3A). Soil moisture and CLPP PCl were both positively correlated with plant species number, while CLPP PCI and PC3 were positively correlated with number of fimctional groups (Table 3.3A). These relationships were still significant when plant biomass was included as a co-variate; however, when the no-plant control plots were excluded from the analyses, the relationships were not significant. Effects of Functional Groups or Individual Species on the Soil Microbial Community Because grass species were present in all mixtures, I were only able to test for the effect of presence/absence of forbs and legumes on the soil and microbial community parameters (Table 3.2). The presence of forbs had a significant positive effect on soil moisture, number of culturable bacteria, microbial respiration, microbial biomass in year 3, and microbial respiration in year 4 (Table 3.38). CLPP PC4 and PCS also distinguished among plots with and without forbs in year 3 (Table 3.3B). The presence of legumes influenced CLPP PC4 in year 3 and corresponded with increased Total N and N-mineralization rates in year 4. The small number of plots sampled only allowed us to evaluate the direct effects of a few species on the soil microbial community; most species were either in all or only a very few mixtures. I were able to detect the effects of two species, Lotus corniculatus (legume) and Hypochaeris radicata (forb), on soil and microbial parameters. Soil from L. corniculatus monocultures had a significantly lower pH than all other plots. The L. 53 corniculatus monocultures also were distinguished from the other plots by having higher values for CLPP on the third PC (Figure 3.3), indicating the soil microorganisms in these plots were better able to metabolize a-D-lactose. Hypochaeris radicata monocultures had the highest soil microbial biomass of manipulated plots in both years (Figure 3.1A,B), suggesting a strong plant species effect. However, I saw no evidence that the presence of this species in mixtures increased the overall soil microbial biomass for the mixture. Effects of Plant Community Composition (Mixture) on the Soil Microbial Community I were able to detect plant community composition effects on the soil microbial community even when aboveground plant biomass was a covariate. Year 3 PLFA PC4 was significantly related to plant species composition. Soil microbial respiration was significantly related to plant community composition in year 3 (Figure 3.1A, Table 3.3A, p< 0.05), but not in year 4 (Figure 3.1B). Soil microbial biomass in these plots was consistent across years (r2=0.39, p <0.01), although not related to diversity or composition of the plant community. Changes in Soil Microbial Community Composition There was considerable variation in CLPP among plots, and much of this could be accounted for by the plant diversity or composition treatments. PCl explained 46.1% of the variation and was correlated with increased species diversity and increased 54 CLPP PC3 (7.6%) — 1 at. 1 11 .1 Oil. R0 211‘} 1. %§;figfil}fi4 l 1 _Number of PG ' 0 . 0 -3: ii 1 . 34 1 1 I -4 ' -4 -3 -2 -1 O l 2 CLPP PCl (46.1%) Figure 3.3. Plant diversity effects on the metabolic activity of the soil microbial community as measured by Community Level Physiological Profiles (CLPP) in year 3 (n = 28). Principal component axis 1 accounted for 46.1% of the total variation. Principal component axis 3 accounted for 7.6% of the overall variation and was driven by the ability to metabolize alpha-D-lactose. Significance values for diversity effects are listed in Table 3.3A. Number of functional groups is represented by symbol shape. No-plant control plots are represented by circles, and non- manipulated control communities labeled with “R”. Number of plant species (0, l, 4, 8, or 11) are labeled. 55 nmnh 150“ \anar 560 w p<0l’ was $1 anabs tflant posuil (Tabh: negau' The V3 m'0 hr indicat Comnn 1411 1 more 1 indicat PC3ac (lite-r31“. Vallalir abere numbers of functional groups (Table 3.3A), while PC2 accounted for an additional 15.0% of the variation, but was not significantly correlated with any explanatory variables. However, PC3 (which accounted for 7.6% of the variation in the CLPP data set) was significantly related to number of functional groups (Figure 3.3, Table 3.3, p<0.05). PC3 was driven by the ability to metabolize a-D-lactose. This relationship was still significant when aboveground plant biomass was included as a covariate in the analysis, indicating that the functional group diversity effect was independent of any plant biomass effect. CLPP PC4 accounted for 6.8% of the variation and was positively correlated with both the presence of legumes and the presence of forbs (Table 3B, p<0.05, p<0.10), while PCS accounted for 5.4% of the variation and was negatively correlated with the presence of forbs (Table 3B, p<0.01). The variation in PLFA profiles of the soil microbial community showed structure at two levels. PC 1 (45.3% of the variation) was significantly related to Block (p<0.01), indicating location in the field was important in structuring the soil microbial community at this site. Soils from the first block contained more C9 dicarboxylic acid, 14:0, 15:0 iso, 16:0 iso, 17:0 iso, and 17:0 anteiso; while the second block contained more 16:1 cis 11, 18:2 cis 12, 18:1 cis 9, and summed in feature 8: 18:1 trans 9, indicating a higher proportion of eukaryotes, most likely fungi (Cavigelli et al. 1995). PC2 accounted for 21.5% of the variation and was significantly correlated with species diversity (Table 3.3A, p<0.05). PC3 accounted for an additional 16.8% of the total variation, but was not significantly correlated with plant diversity, plant productivity, or aboveground biomass. A much smaller amount of the variation in PLFA profiles (PC4, 56 .0 o 00 l 0.00 - E . E r E -0.08 _ —~ -_ PLFA PC4 (4.5%) f -O.16 “‘ JJJJJIJJJJJJJ C°\m°> use 3°°°tx°00 Figure 3.4. Plant community composition effects on the structure of the soil microbial community, measured with Phospholipid Fatty Acid profiles of 1998 soil samples. Principal component axis 4 accounted for 4.5% of the overall variation and reflected the amount of 15:0 anteiso in the PLFA profiles. Plant community identification codes are from Table 3.1B. No-plant control plots are represented by a blank circle. The reference plots were not included in this analysis. Monocultures are indicated by circles and mixtures by triangles. 57 ...1 r..- -. 3.99“ p <0, micrt Discu 1 EXP: sigm I" comm rather range effects 3.9%) was significantly related to plant community composition (Figure 3.4, Table 3.2, p <0.05). PC4 reflected the amount of 15:0 anteiso in the PLFA profiles. As with microbial biomass and CLPP, this relationship was independent of plant biomass. Discussion I expected to find that the composition and diversity of the plant community would significantly affect soil microbial diversity and productivity. I found that plant community diversity and composition affected soil microbial community structure rather than processes (Table 3.3A,B). Although I expanded the number of plots and range of treatments sampled in year 4, I detected fewer diversity and composition effects than in year 3 and the significant variables were inconsistent across years (Table 3.3A,B). In most cases, the initially detected effects of plant species number or number of functional groups on the soil microbial community were reduced or became non- significant when above-ground plant biomass was included as a covariate in these analyses. Plant diversity effects above and beyond those effects on the microbial community that could be accounted for by productivity were detected for CLPP and PLFA profiles in year 3. However, CLPP profile differences seem to be driven primarily by the no plant control plots. Only year 3 did I detect a shift in the soil microbial community (PLF A PC2) in response to plant species diversity (Table 3.3A). 58 Plant Diversity Effects on the Soil Microbial Community A number of studies have suggested that there should be a relationship between plant diversity and soil microbial diversity (Ohtonen et al. 1997, Wardle and Giller 1996, Schlapfer and Schmid 1999). In most field studies of plant community effects on soil microbial communities several explanatory variables (e.g. plant diversity, plant productivity, and plant community composition) are confounded. For example, J.C. Zak et al. (1994) detected differences in the structure of the soil microbial community (using Biolog) along an elevational and moisture gradient in the Chihuahuan Desert at the Jomada Long-Tenn Ecological Research site. Similarly, Goodfriend (1998) used CLPP patterns to distinguish among the communities at eight sites representing a variety of wetland communities. Broughton and Gross (2000) examined characteristics of the soil microbial community composition along a natural topographic, productivity and diversity gradient at a site in southwestern Michigan and found a correlation between the productivity of the plant and soil microbial communities, but no relationship between plant diversity and soil microbial community composition (results in Chapter 2). However, in all of these studies, the plant communities sampled were from different sites in which there were likely concomitant changes in soil characteristics, so the influence of plant community composition differences could not be assessed independently of differences in edaphic characteristics. In contrast, Wardle et al. (1999) showed plant effects on the soil microbial community, which were not confounded by either soil or management effects. They removed 59 subsets of the plant community (that varied in number and functional group composition) from a New Zealand perennial grassland. PLFA patterns distinguished among soils from the plant removal treatments, suggesting plant community composition effects on soil microbial community structure (Wardle et al. 1999). The BIODEPTH experiment provides a unique opportunity to examine both species diversity (through number and functional group) and composition effects on the soil microbial community. Because the Silwood Park site preparation included fumigation with methyl bromide after tillage to destroy the seed bank, the soil microbial community was “standardized” before the initiation of treatments. This allowed us to control for the effects of soil factors and focus solely on the manipulated plant diversity and composition treatments as explanatory factors for the soil microbial community. A recent paper by Stephan et al. (2000) from the Swiss BIODEPTH site reported a relationship between plant species richness and functional diversity of the culturable soil microbial community as measured by CLPP. They found that increased plant species richness and plant functional diversity increased the overall catabolic activity diversity in CLPP. However, Stephan et al. (2000) did not include a measure of plant biomass as a covariate in these analyses, so it is not clear the extent to which the plant diversity effect on the culturable soil bacteria is due to a correlated plant productivity effect on the microorganisms. Hector et al. (1999) reported a strong relationship between plant diversity and plant biomass at the Swiss site. At higher plant diversities, aboveground plant biomass is greater, likely making labile carbon available to the soil 6O microorganisms. Unlike the Silwood Park site, soils at the Swiss site were not fumigated prior to the establishment of the diversity treatments (Hector et al. 1999, Spehn et al. 2000a). These two sites also differed in the range of species diversity used in the experiment: the Silwood site had a maximum species richness of l 1, whereas the Swiss site had a maximum of 32 (Hector et al. 1999, Spehn et al. 2000b). These differences in range of species diversity examined at the two sites is reflective of the natural diversity at these sites. This may also affect the ability to detect plant species diversity effects on the soil microbial community at these two sites. CLPP catabolic activity has been found to be strongly related to innoculum density (Garland and Mills 1991, Haack et al. 1994) and, thus, can be an approximate indicator of overall bacterial number. I did not detect a significant relationship between CLPP overall catabolic activity and plant diversity at the Silwood site. However, the increase in culturable bacteria with both increased plant species diversity (Figure 3.2A) and increased plant functional group diversity (Figure 3.2B) reflects the underlying relationship between plant diversity and plant biomass. The number of culturable bacteria in the soil is clearly correlated with the overall aboveground plant biomass of the plot (Figure 3.2C). Plant Community Composition Effects on the Soil Microbial Community In year 3 (but not year 4), I found plant community composition effects on some soil and microbial parameters (Table 3.3A). Some previous studies have shown plant 61 COII‘. soil eff: Tn,- Cor community composition to be more important than species number or number of fimctional groups in influencing ecosystem processes. Hooper and Vitousek (1998) concluded that plant community composition better explained variation in nutrient cycling processes in a Californian serpentine grassland than number of functional groups. In a comprehensive study using plant removals, Wardle et al. (1999) found significant effects of plant community composition on several different trophic levels, including the soil microbial community, and ecosystem properties. These results suggest that individual plant species may influence communities and processes independent of any diversity or productivity effects. I also found that plant community composition significantly affected some soil and microbial parameters at the Silwood site, but these effects varied across years. The presence of legumes was positively correlated with CLPP PC4 (year 3), Total N (year 4), and N-mineralization rates (year 4; Table 3.3B). As legumes are symbiotic with nitrogen-fixing bacteria in their roots, it is not surprising to find an effect of legumes on soil microbial processes. However, I were not able to determine whether the legume effects were due to a particular legume species, because both L. corniculatus and Trifolium repens were present in all mixtures containing legumes. However, L. corniculatus monocultures differed from the other monocultures and mixtures in both pH and CLPP profiles. At the Swiss BIODEPTH site, Stephan et al. (2000) found that legumes had positive effect on overall CLPP catabolic activity, and Spehn et al. (2000a) reported a positive effect of legumes on microbial biomass. In contrast to this study, Stephan et al. (2000) were able to detect the effect of the presence of a specific 62 met. SM: duh bye I990 Lego SUQ. refl suc Mr fin: legume, T. repens, on the soil microbial community. The presence of T. repens was positively correlated with CLPP catabolic activity and number of carbon sources metabolized at the Swiss site (Stephan et al. 2000). Although both the Silwood and Swiss sites had L. corniculatus and T. repens grown in monocultures and mixtures; differences between these sites in the specific effects of these species suggest a species by environment interaction as seen for aboveground biomass (Table 3.3, Hector et a1. 1999). Legacy Effects on the Soil Microbial Community N-mineralization rates did not differ significantly among mixtures; however, the reference plots consistently had lower N-mineralization rates than the treatment plots. N-mineralization rates typically are higher in earlier successional sites and decline over time (Schlesinger 1997). These differences in N-mineralization rates may reflect the successional status of the treatment and reference plots. This temporal change may reflect the immobilization of nitrogen by the soil microbial community in later successional plots and a more mature soil microbial community (Schlesinger 1997). Microbial biomass measurements from the reference communities at Silwood indicate that the treatment plots still had not recovered from fumigation afler 4 years. Because microbial respiration and microbial biomass may be correlated with aboveground plant biomass (Broughton and Gross 2000), and plant diversity is correlated with aboveground plant biomass, I would expect to see higher microbial activities and 63 biomasses at higher plant diversity levels. Spehn et al. (2000a) observed a positive relationship between soil microbial biomass and plant species diversity at the Swiss site where the soil was not fumigated. The fact that I did not see a relationship between plant diversity and soil microbial biomass at the Silwood site suggests that the soil microbial community is still recovering from the severe disturbance of methyl bromide application. Alternatively, it may be that different relationships between plant diversity and soil microbial biomass emerge under different local conditions. Studies at other BIODEPTH sites may help to resolve this issue. Disturbance effects on the soil microbial community may persist for decades and make it difficult to detect current plant species or diversity effects on the soil microbial community. Buckley and Schmidt (2001) found that there was little difference between the soil microbial communities of a continuously tilled agricultural site and a companion successional site (abandoned for 12 years) in southwestern Michigan. A nearby reference field (never- tilled) had a distinct soil microbial community (detected using rRNA) from either the tilled or successional fields (Buckley and Schmidt 2001). The results from this study provide some evidence that there is an overall plant diversity effect on the soil microbial community. However, both the productivity and the composition of the plant community can have effects on the biomass and structure of the soil microbial community. In addition, disturbance effects on the soil microbial community may persist for quite some time. Longer-term studies that can separate soil microbial community function from structure are necessary to better investigate the role 64 that plants play in how the soil microbial community mediates the flow of carbon and nitrogen through the ecosystem. 65 REFERENCES Bligh, E.G. and W. J. Dyer. 1959. A rapid method of total lipid extraction and purification. Canadian Journal of Biochemistry and Physiology 37 : 911-917. Broughton, LC. and K. L. Gross. 2000. Patterns of diversity in plant and soil microbial communities along a productivity gradient in a Michigan old-field. Oecologia 125: 420-427. Buckley, DH. and T. Schmidt. 2001. The structure of microbial communities in soil and the lasting impacts of cultivation. Microbial Ecology 42: 11-21. Cavigelli, M.A., G. P. Robertson, and M. J. Klug. 1995. Fatty acid methyl ester (FAME) profiles as measures of soil microbial community structure. Plant and Soil 170: 99-113. Garland, J.L. and A. L. Mills. 1991. Classification and characterization of heterotrophic microbial communities on the basis of patterns of community-level sole-carbon- source utilization. Applied and Environmental Microbiology 57: 2351-2359. Goodfriend, W.L. 1998. Microbial community patterns of potential substrate utilization: a comparison of salt marsh, sand dune, and seawater-irrigated agronomic systems. Soil Biology and Biochemistry 30: 1169-1176. Haack, S.K., H. Garchow, D.A. Odelson, L.J. Fomey, and M.J. Klug. 1994. Accuracy, reproducibility, and interpretation of fatty acid methyl ester profiles of model bacterial communities. Applied and Environmental Microbiology 60: 2483- 2493. Hector, A., A. J. Beale, A. Minns, S. J. Otway, and J. H. Lawton. 2000. Consequences of the reduction of plant diversity for litter decomposition: effects through litter quality and microenvironment. OIKOS 90: 357-371. Hector, A., B. Schmid, C. Beierkuhnlein, M. C. Caldeira, M. Diemer, P. G. Dimitrakopoulos, J. A. Finn, H. Freitas, P. S. Giller, J. Good, R. Harris, P. Hogberg, K. Huss-Danell, J. Joshi, A. Jumpponen, C. Komer, P. W. Leadley, M. Loreau, A. Minns, C. P. H. Mulder, G. O’Donovan, S. J. Otway, J. S. Pereira, A. Prinz, D. J. Read, M. Scherer-Lorenzen, E-D. Schulze, A-S. D. Siamantziouras, E. M. Spehn, A. C. Terry, A. Y. Troumbis, F. I. Woodward, S. Yachi, and J. H. Lawton. 1999. Plant diversity and productivity experiments in European grasslands. Science 286: 1123-1127. Hooper, D. U. and P. M. Vitousek. 1998. Effects of plant composition and diversity on nutrient cycling. Ecological Monographs 68: 121-149. Huston, MA. 1997. Hidden treatments in ecological experiments: re-evaluating the ecosystem function of biodiversity. Oecologia 110:449-460. J onasson, S., A. Michesen, I. K. Schmidt, and E. V. Nielsen. 1999. Responses in microbes and plants to changed temperature, nutrient, and light regimes in the arctic. Ecology 80: 1828-1843. 66 Nelson, D.W. and LE. Sommers. 1982. Total carbon, organic carbon, and organic matter, pp. 539-580, In_ Methods of Soil Analysis Part 2 - Chemical and Microbiological Properties. ASA Special Publ. No. 9. ASA, SSSA, Madison, WI. Ohtonen, R., S. Aikio, H. Véire. 1997. Ecological theories in soil biology. Soil Biology and Biochemistry 29: 1613-1619 Paul, EA. and FE. Clark. 1996. Soil Microbiology and Biochemistry. 2"d ed. San Diego, CA: Academic Press, Inc. Paul, E.A., D. Harris, H. P. Collins, U. Schulthess, and G. P. Robertson. 1999a. Evolution of C02 and soil carbon dynamics in biologically manage, row-crop agroecosystems. Applied Soil Ecology 11: 53-65. Paul, E.A., D. Harris, M. J. Klug, and R. W. Ruess. 1999b. The determination of microbial biomass, pp. 291-317, In Standard Soil Methods for Long-Term Ecological Research (eds G.P. Robertson, D.C. Coleman, C.S. Bledsoe, & P. Sollins) Oxford University Press, New York. Rosenzweig, ML. 1995. Species Diversity in Space and Time. Cambridge University Press, Cambridge. Schlapfer, F. and B. Schmid. 1999. Ecosystem effects of biodiversity: a classification of hypotheses and exploration of empirical results. Ecological Applications 9: 893-912. Schmidt, I.K., A. Michelsen, and S. Jonasson. 1997. Effects of labile soil carbon on nutrient partitioning between an arctic graminoid and microbes. Oecologia 112: 557-565. Shaver, GR. and F. S. Chapin, III. 1980. Response to fertilization by various plant growth forms in an Alaskan tundra: nutrient accumulation and growth. Ecology 61: 662-675. Schlesinger, W.H. 1997. Biogeochemistry An Analysis of Global Change. Academic Press. San Diego, CA. Spehn, E.M., J. Joshi, B. Schmid, J. Alphei, and C. Komer. 2000a. Plant diversity effects on soil heterotrophic activity in experimental grassland ecosystems. Plant and Soil 224: 217-230. Spehn, E.M., J. Joshi, B. Schmid , M. Diemer, and C. Komer. 2000b. Aboveground resource use increases with plant species richness in experimental grassland ecosystems. Functional Ecology 14: 326-337. Stephan, A., A. H. Meyer, and B. Schmid. 2000. Plant diversity affects culturable soil bacteria in experimental grassland communities. Journal of Ecology 88: 988— 998. Tiedje, J. M. 1995. Approaches to the comprehensive evaluation of prokaryote diversity of a habitat, pp. 73-87, In Microbial Diversity and Ecosystem Function (eds D. Allisopp, R.R. Colwell, and D.L. Hawksworth). CAB International, Wallingford, UK. 67 Wardle, DA. and K. E. Giller. 1996. The quest for a contemporary ecological dimension to soil biology. Soil Biology and Biochemistry 28: 1549-1554. Wardle, D.A., K.I. Bonner, G.M. Barker, G.W. Yeates, K.S. Nicholson, R.D. Bardgett, R.N. Watson, and A. Ghani. 1999. Plant removals in perennial grassland: vegetation dynamics, decomposers, soil biodiversity, and ecosystem properties. Ecological Monographs 69: 535-568. Zak, J. C., M. R. Willig, D. L. Moorhead, and H. G. Wildman. 1994. Functional diversity of microbial communities: A quantitative approach. Soil Biology and Biochemistry 26: 1101-1108. 68 CHAPTER 4 PLANT-MEDIATED EFFECTS OF SOIL ORIGIN ON THE COMPOSITION AND FUNCTION OF SOIL MICROBIAL COMMUNITIES Introduction While considerable attention has been paid to factors that affect the composition and function of communities of macroorganisms, very little is known about the factors that affect the structure and function of soil microbial communities (Ohtonen et al. 1997, Tiedje 1995). To understand how changes in the structure of the soil microbial community affect ecosystem functions, I must first investigate what factors influence soil microbial community structure and function. Plant community composition can be stable for long time periods, but can also vary depending on factors such as disturbance history and successional status. In contrast, soil characteristics change much more slowly on average than the plant community (e.g. soil quality). Consequently, soil characteristics may have a more consistent effect on the soil microbial community than plants. While the soil has a large reserve of relatively recalcitrant carbon that is less available to microorganisms, much of the labile carbon available to the soil microbial community is derived from recent plant production (Paul and Clark 1996). Because the plant community is dynamic and the main source of carbon for the soil microbial community, the current plant community should have a big effect On structure and function. To date, many studies have investigated the role of edaphic factors in structuring the soil microbial community, while fewer have addressed the effects of plants (Metting 1993, Schlapper and Schmid 1999, Hooper et al. 2000). Additionally, 69 very few have sought to distinguish between the effects of soil and the effects of plants on the soil microbial community. Several studies that have reported plant community effects on soil microbial communities confound direct plant-mediated effects with soil effects. For example, J .C. Zak et al. (1994) detected differences in the structure of the soil microbial community (using Biolog) from plant communities that occur along an elevational and moisture gradient in the Chihuahuan Desert at the Jomada Long-Term Ecological Research site. Similarly, Goodfi'iend (1998) used Biolog to distinguish among the soil bacterial communities at 8 sites representing a variety of wetlands across a salinity gradient. However, in both these studies, the plant communities sampled were from different sites in which there were likely concomitant changes in soil characteristics, so the influence of plant community composition differences could not be assessed independently of differences in edaphic characteristics. Other studies in single sites have found little change in microbial communities in soils sampled from different plant communities (e.g. Buckley and Schmidt 2001). After 10 years of plant community divergence resulting from vastly different agricultural management, Buckley and Schmidt (2001) could detect no differences among rRNA patterns of the soil microbial communities among treatments at the Kellogg Biological Station’s Long Term Ecological Research site, suggesting that the soil microbial community structure was still dominated by the influence of the past land use and soil quality or, perhaps, that the various plant forms contributed similarly to the sustenance of the microbial assemblage despite our perception that they might differ in this respect. The plant communities 70 sampled ranged from successional fields to poplar plantations to conventional com. All had been under similar management (conventional corn) prior to treatment implementation (Robertson et al. 1993). In this study, I am interested in distinguishing the relative importance of variability in soil characteristics and the current plant community in controlling soil microbial community structure and composition (see Figure 1.1). Specifically, I used a manipulative greenhouse experiment to investigate whether (1) soils from different plant communities that differ in fertility vary in the composition and function of the soil microbial community and (2) plants can mediate these effects. Methods Site Descriptions I selected six successional old-fields at the W. K. Kellogg Biological Station of Michigan State University in southwestern Michigan to compare the soil microbial communities of sites with different plant communities. The sites varied in fertility, species richness, and dominant plant type, but all were located on Kalamazoo sandy loam soil. I determined plant species composition at each site in six 0.5 m x 2 m plots in August 1998. To estimate above-ground net primary productivity (ANPP), I clipped aboveground biomass at ground level from a 0.5 m x 0.5 m plot located within the plots used to assess species diversity and composition. 71 The six sites also differed in past land use and ranged in time since abandonment from 20 to 50 years. McKay field was abandoned from agriculture in 1973; a section was plowed once in 1981 and then re-abandoned (Burbank et al. 1992). Both the Upper and Lower Louden fields were abandoned from agriculture in 1951 (Burbank et al. 1992). The Bailey field site was farmed until ten years prior to this sampling (K.L. Gross, personal communication). The Pond Lab Orchard and Field K sites had been abandoned for at least twenty years (Foster 1996). Experimental Design To determine if site differences in ANPP and species composition had detectable effects on the soil microbial community I incubated soils from each site in the greenhouse and evaluated the soil microbial communities 12 to 16 weeks later. To determine if plants could mediate these differences, I sowed half the pots with Andropogon gerardi, a C4 grass native to Michigan prairies. I collected approximately 10 kg of soil in June 1998 from the top 15 cm of each field in the same area from which species diversity and plant biomass were sampled. The soil was sieved to 4 mm and thoroughly mixed. Soil was stored at room temperature until the experiment was established in the greenhouse (less than 2 weeks). I used a randomized complete block design for the greenhouse experiment to test for the effects of soil origin and plant effects on the soil microbial community: 6 soils x 2 72 treatments, with 8 replicates of each. The two treatments were control (no-plant) and plant (Andropogon gerardi). I chose A. gerardi because it is a native C4 grass that can grow in all of these fields, although it was not present in our soil collection sites and is rare in these communities because it is out-competed by naturalized C3 grasses (Foster 1996). This allowed us to measure the effects of a relatively novel plant on the soil microbial communities present in each site. Seeds of A. gerardi were collected from local fields in autumn 1997, and stored at room temperature in the laboratory until used for these experiments the following summer. Soils were placed in 5 cm diameter x 20 cm deep pots and kept well-watered with de- ionized water to avoid adding nutrients or contaminants. Temperature in the greenhouse ranged from 25 to 40 °C; light availability was controlled through a 12 h light/ 12 h dark cycle. Treatments were randomly assigned within replicates. Andropogon gerardi was added as 2 week old seedlings; all seedlings were germinated in a sterile sand medium in 3 grth chamber and were less than 2 cm in height when transplanted. I estimated initial biomass by drying a representative subset of the seedlings at the time of transplantation. The experiment ran for a total of 16 weeks. I harvested the experiment in two segments because of the number of samples and the time required to process each sample: 4 replicates were harvested at week 12, and the remaining 4 replicates were harvested at week 16. Thus, time was an additional factor in the ANOVA. 73 Data Collection Differences among sites and treatments in soil fertility were assessed by (1) grth of Andropogon gerardi, (2) inorganic nitrogen pools and N—mineralization rate, and (3) soil organic matter. Shoot and root biomass of Andropogon gerardi were harvested separately and dried at 60°C for 48 h. I separated root biomass from the soil during sieving; roots were rinsed thoroughly in de-ionized water before drying. I used the change in total plant biomass to estimate the relative growth rate (RGR) as [In (total plant biomass) — In (initial plant biomass)] / number of days between harvest and planting. I sieved the soil through a 2 mm sieve, and stored it in sealed plastic bags at 4°C until analysis. All analyses were done within 3 days of sampling, except PLFA. Soil for PLFA analyses was kept at —80°C until the fatty acids were extracted. I determined gravimetric soil moisture for each sample by drying 10 g soil at 105°C for 48 hours (Nelson and Sommers 1982). A subsample of the dried soil was ashed at 500°C for 4 hours to determine organic matter content (Nelson and Sommers 1982). For nitrogen analyses, I extracted 20 g of fresh soil in 100 m1 1M KCl. The samples were shaken for l min and allowed to settle for 24 h at room temperature. The supernatant mixture was filtered through a l-um Gelman glass filter and N03 and NH4+ concentrations were measured using Alpkem auto-analyzer. To determine potential N-mineralization and nitrification rates, a companion 20 g sample was 74 incubated for 21 days at 25°C and 15% humidity and then extracted using the same methods as above. The remaining soil was used to characterize the soil microbial community. I assessed differences in soil microbial community production among the sites and treatments by (1) microbial biomass C, (2) microbial respiration, and (3) plate counts (number of colony-fonning units). I determined microbial biomass using the chloroform fumigation incubation method (Paul et al. 1999). Two 25 g soil samples were pre-incubated for 5 days then one sample was fumigated with chloroform for 24 hours to kill the microorganisms. After a vacuum was created and the chloroform evaporated, 0.5 g of original soil was added to both samples. I measured initial headspace C02 and accumulated C02 after 10 days on an ADC series EGA infrared CO; gas analyzer (The Analytical Development Co. Ltd., Hoddesdon, Herts., UK). I calculated microbial biomass as [ 1.73 * (10 day accumulated COz-C — initial COz-C for the fumigated samples) - 0.56 "‘ (10 day accumulated COz-C — initial COz-C for the control samples)] (Paul et al. 1999). To determine microbial respiration I used a separate set of 10g soil samples that were pre-incubated 5 days in a 160 ml glass qorpak bottle. I measured initial headspace C02 and accumulated CO; afier 1 and 5 days. I determined the number of colony-forming units by mixing 5 g of fresh, sieved soil into 1% phosphate buffer to reach a final dilution of 10'6 g soil/ ml. I plated this solution on minimal media (R2A agar plates) and incubated the plates at 25 °C and then counted the number of colony-forming units after 24 h and 48 h. 75 I assessed soil microbial community structure differences among sites and treatments by (1) Community-level physiological profiles (CLPP) and (2) phospholipid fatty acid (PLFA) profiles. Community-level physiological profiles (CLPP) were determined using Biolog GN plates (Biolog, Inc., Hayward, Calif, USA) and reflect the range and amount of carbon sources or resources that can be metabolized by the community (Konopka et al. 1998). For the assay, 1 g of fresh, sieved soil was shaken with 99 ml 1% phosphate buffer solution for 20 min. 150 pl of the mixture was transferred to each well of the microtiter plate (GN Biolog, 95 Carbon sources + l non-Carbon control). The plates were incubated in the dark at 25 °C and optical densities were measured at 24 h intervals from 0 h to 96 h using an Emax precision microplate reader (Molecular Devices Corp., Menlo Park, Calif, USA). Because the 5 incubation times gave consistent results, I present here data only from the 96 h Biolog measurements. Optical densities (intensity of resource use) were used in the RDA analysis. For the PLFA analysis, I extracted lipids from 6 g whole soil samples for 2 h using a mixture of dichloromethane (DCM): methanol: phosphate buffer (1:2:0.8 v/v/v), following a modified Bligh-Dyer procedure (Bligh and Dyer 1959). Phase separation was achieved by adding DCM and saturated sodium bromide solution (1:4 v/v). I isolated the phospholipid fatty acids from the dried lipid extracts by solid phase extraction. The lipid material was added to a polar column consisting of 100 mg silica (Varian Bond Elut LRC Columns, Product # 1211-3010). Lipids of low or intermediate polarity were eluted with chloroform and acetone and discarded. Subsequently, 76 phospholipid fatty acids were eluted with 1.5 ml methanol for preparation of fatty acid methyl esters. I saponified the samples using 1 ml NaOH (15% w/v) in methanol (50% v/v) at 100 °C for 30 min and methylated the sample with 2 ml 6M HCl in methanol at 80 °C for 10 min. I extracted the fatty acid methyl esters into 1.25 ml (1:1 v/v) methyl- tert—butyl etherhexane for 10 min and washed the extract with 3 ml 1.2% NaOH. Phospholipid amounts were measured using a HP 5890 series II gas chromatograph (Hewlett Packard Co., Palo Alto, Calif, USA) equipped with a 7673 autosampler and flame ionization detector (Microbial ID Inc., Newark, Del., USA). Peaks were identified by comparison with an external standard. For analysis, I included only those phospholipid fatty acids that were present in greater than 50% of samples and reported their abundance as the square root of the proportion of the total phospholipid fatty acid amount in each sample (Hellinger transformation). Of the 70 lipids detected, 30 phospholipid fatty acids met this criterion (Table 4.3). Statistical Analyses I used a randomized complete block design (ANOVA) model to test the effects of soil origin, presence/absence of plant, and time on the following response variables: soil moisture, soil organic matter, plant biomass, relative growth rate, total N, N- mineralization rate, nitrification rate, number of culturable bacteria, microbial respiration, and microbial biomass. 77 I used a modified redundancy analysis (RDA, Legendre & Anderson 1999) to determine the relationship between the environmental factors (soil origin, presence/absence of plant, time of harvest) and the two measures of soil microbial community structure, CLPP and PLFA. This is a relatively new, powerful technique for multivariate analysis. RDA is a multiple regression technique that reduces the number of variables necessary to explain the variation in a data set by creating composite variables. In addition, RDA compares a second matrix that describes the environment in which the original variables were measured. This new technique also uses permutations to allow for statistical tests of how these composite variables vary with the explanatory variables to determine the strengths of the significance of any environmental correlations with measures of the soil microbial community. Because the CLPP and PLFA data matrices have many zeros, I transformed the CLPP and PLFA data using a Hellinger transformation (a square root transformation of relative abundance, Legendre & Gallagher, in press). The RDA procedure involved: (1) the creation of a matrix of dummy variables corresponding to the randomized complete block design (modeled from the experimental design: soil origin, presence/absence of Andropogon gerardi, time at harvest), (2) redundancy analysis of the relationship between the principal coordinates (matrix of optical density or phospholipid fatty acid data) and the environmental variables (matrix of dummy variables in (1)), and (3) implementation of a Monte Carlo permutation test to estimate the statistical relationship between the two matrices (Legendre & Anderson 1999). This analysis allows us to test which factors from the experimental design (soil origin, 78 presence/absence of Andropogon gerardi, time at harvest) are significantly related to the variation in the CLPP and PLF A patterns. The modified RDA is a better statistical technique than regular ordination techniques because it allows for significance testing. Results Plant Communities The six field sites varied in ANPP, species diversity, and soil organic matter (Table 4.1). McKay Field (MK) had the lowest ANPP and a low species diversity and was dominated by Agropyron repens, a C4 perennial grass. Both Bailey (Ba) and Upper Louden (UL) had moderate ANPP and high species diversities and were dominated by diverse forb communities. In contrast, Lower Louden (LL) and the Pond Lab Orchard (PL) field had moderate ANPP and species diversities and were dominated by graminoids. Lower Louden was dominated by Bromus inermz's, a C3 perennial grass, although perennial forbs such as Solidago canadensis, Daucus carota, T araxacum oflicionale, and Hieracium sp. contributed significant biomass to total ANPP. The Pond Lab Orchard site was dominated by several C3 species: Bromus inermis, Agropyron repens, and Poa pratensis. Field K (FK) had the highest ANPP and lowest species diversity and was dominated by Bromus inermis (Table 4.1) 79 Table 4.1. Plant productivity and diversity of abandoned fields from which soil was collected for the greenhouse experiment. Values for Annual Net Primary Productivity, species richness, and mean percent organic matter are expressed as mean i standard deviations. Values that are not significantly different for a given variable based on Fisher’s LSD test have the same letter. Site Dominant Peak Plant Biomass Species diversity Soil Organic Plant Form (standing + litter, g/mz) (#/m2) Matter (%) MK Grass 188 i 16 a 2.2 i 0.4 a 2.40 i 0.43 a UL Forb 3203223 b 15.8i1.0e 3.17:0.23b Ba Forb 424i52c 11.3i1.0d 3.03:0.17b LL Grass 432 i 27 c 8.5 i 0.4 c 3.84 :t 0.32 0 PL Grass 480 i 48 c 5.7 i 0.8 b 3.63 i 0.28 c FK Grass 592 i 22 d 1.3 i 0.2 a 3.84 i 0.20 c Effects of Soil Origin on Soil and Microbial Processes The differences among sites in plant community productivity were reflected in the growth of Andropogon gerardi in the greenhouse (p < 0.001, F = 40.3, Table 4.2A, Figure 4.1A). A. gerardi grown in soils from more productive sites had greater total biomass (Table 4.2A, Figure 4.1). The positive relationship between A. gerardi production and 1998 field above-ground plant biomass (Figure 4.1) suggests that the ranking of sites based on plant productivity also reflected differences in fertility. Interestingly, time had no effect on any of the measured plant growth or soil or 80 Table 4.2. Effect of soil origin and presence of A. gerardi on (A) plant and (B) soil and microbial variables as detected by Analysis of Variance. NS = not significant, p > 0.05. The time factor investigates the results of harvesting half the experiment at 12 weeks, the other half at 16 weeks. (A) Plant Variable BLOCK SOIL TIME SOIL*TIME Total Biomass (g) NS < 0.001 NS NS Root Biomass (g) 0.045 < 0.001 NS NS Shoot Biomass (g) NS < 0.001 NS NS RGR (g/day) NS < 0.001 < 0.001 NS Plant Height (cm) NS < 0.001 NS NS (B) Soil or Microbial Variable BLOCK SOIL PLANT TIME SOIL*PLANT Percent Organic Matter < 0.01 < 0.001 NS -- NS Soil Moisture NS < 0.01 < 0.05 NS < 0.05 Total Inorganic Nitrogen NS < 0.001 < 0.001 NS < 0.001 N-mineralization Rate NS NS < 0.001 NS < 0.05 Nitrification Rate NS NS < 0.001 NS < 0.01 Microbial respiration NS < 0.001 < 0.001 -- < 0.05 Culturable Bacteria (CFU’s) NS < 0.001 NS -- < 0.05 Microbial Biomass < 0.01 NS NS NS NS 81 10 Han7est ' T ' A 9 12 wks FK :9 8 C l6wks T 3 LL PL i3 e 6- r i "U 8 Q4 1 :5 4" MK UL ; “ is 3 i Ba 8 be 2 -— - V’ l l l l 0 l 00 200 300 400 500 600 Field Above-Ground Biomass (g/mz) Figure 4.1. Total plant biomass of Andmpogon gerardi produced at 12 and 16 weeks in relation to variation among sites in 1998 field above-ground plant biomass. Soils are coded as in Table 4.1. Values are mean i standard error, n = 8. Significance values from the ANOVA are listed in Table 4.2A. 82 microbial variables measured, indicating that the 4 weeks difference in harvesting replicates had no discemable effect on the results (Table 4.2A&B). Consequently I combined data from the two sampling intervals for the subsequent analyses. Soil origin significantly affected total inorganic nitrogen (p < 0.001, F = 14.9, Table 4.2B, Figure 4.2A), microbial respiration (p < 0.001, F = 17.8, Table 4.2B, Figure 4.2D), and the number of colony-forming units (p < 0.001, F = 12.2, Table 4.2B, Figure 4.2E), but did not influence N-mineralization rate (Table 4.23, Figure 4.28), nitrification rate (Table 4.2B, Figure 4.2C), or microbial biomass (Table 4.2B, Figure 4.2F). In general, sites with higher fertility soils had higher soil microbial respiration, and higher nitrogen pools in the absence of plants. Effects of Andropogon gerardi on Soil and Microbial Processes The presence of Andropogon gerardi significantly affected several soil and microbial characteristics and processes (Table 4.28). The presence of Andropogon gerardi decreased soil moisture (p < 0.05, F = 4.5, Table 4.2B), total inorganic nitrogen (p < 0.001, F = 643.6, Table 4.2B, Figure 4.2A), N-mineralization (p < 0.001, F = 14.9, Table 4.2B, Figure 4.2B) and nitrification rates (p < 0.001, F = 18.6, Table 4.2B, Figure 4.2C), and increased soil microbial respiration (p < 0.001, F = 322.7, Table 4.2B, Figure 4.2D). The most dramatic effect was on total inorganic nitrogen; the presence of Andropogon gerardi reduced nitrogen to similar low levels in all soils (Figure 4.2A). 83 .mmé 03:: :_ 08m: 0:: <>OZ< 0:: 80.: m03:> 0080585 .v n : 0>:: 2033 8:888: 33088: :8: $883 03838: :0 8080: 808:0 =: :0: w n 2 .:0:8 080:8: H ::08 0:: m0:_:> .2008»: :003 30:88 8:30: 0:: 8:08 888% zomomofifix 5:5 3:08:08 :0038: 005,383.06 0:: 838:0. 63:8? 8:0 :0... .3: 03:... 8 00%: m: 53:02:08 818088 .«0 80:0 8 888:: 0:: 8:5 8:805 83088 0: 0:: 888:: 0383.30 :0 808:: m: 803888: 83088 R: .0::: 83:05:08 ADV .08: :0_::N__:8:_8-Z Amv .:0w0::: 08:80:: .80: 03 8 00:0. w:08: :0_::_::> .Né 0:33: 0:5 0:5 03m v7.— 15 s: :m ADM—2 vi .5 x: :m ADM—2 v.“— 15 s: :m JD 22 — q n 4 - _ o u A a u q a o N 8 1fi — _ fl 4 o ) .n ) . a m .. w .n 2.: l. fir Au 18N “W M T w «W. LON) m. T e - 1% 7m M.J. F 0m . DJ: 0 6 3 «J m. . + . .2: z. m. .H. n y. e - .w m... . mm - .0va r + -Epu 8 8 r. o e o m m r 180 D. m. nwr m. s .m 0 . ,u w ( m o m. m C m m # .00 m. r o w + -2 N. w .l 160” ”h s ..V m- u . ( W M _ . : L : GOA: : : _ _ _ _ Ow : h : : r : ON 25 : 25 m 35 : vi 4: .3 :m 43% . vi .5 ‘3 :m .5 v.2 .2 vi .5 ‘5 :m :5 v=2 : _ i : : 4 mo mm + l 4 : q : vo N f O f O O O \l N w w. W: - La 1 o o o 430% w I o o .0: Wm e we; .0 Q0 0 . p D.W. _l IOV m. W. 6 v :w ..M a :W J S u m: u. l C L . : ++12mm - hon: H 8m. mm mm . + c:: A: be .(A\ m T A: .9800 0.. m ( 8:3 0 p _ _ : _ h o._ : r : _ _ _ o._ . _ : : 50850:. oo— U m < 84 uaSon! N oruefiioul [moi Similarly, the presence of A. gerardi decreased N-mineralization and nitrification rates to similar low levels in all soils (Figure 4.ZB&C). In contrast, soil microbial respiration increased in the presence of A. gerardi, but the magnitude of this effect decreased with fertility (Figure 4.2D). Although the number of culturable bacteria varied across sites (Table 4.2B), and there was a significant plant x site interaction, there was no consistent effect of A. gerardi on this variable across sites. The presence of A. gerardi also did not have a consistent effect on either the number of culturable bacterial colonies (Table 2B, Figure 2B) or microbial biomass (Table 2B, Figure 2F). Plant and Soil Effects on Soil and Microbial Processes For some variates, the soil x plant treatment interaction (i.e. the magnitude of the A. gerardi effect) varied across the sites and appeared to be related to soil fertility. To evaluate this relationship, I estimated the magnitude of the relative “plant effect” on these variables by calculating the relative difference in the variable in the plant versus no-plant treatments ((plant — control)/ control). The magnitude of the effects of A. gerardi on soil and microbial variables is illustrated in Figure 4.3. For total nitrogen, the magnitude of the effect of A. gerardi varied with site fertility and was inversely related to plant biomass (Figure 4.3A, sites are ranked by fertility as per Figure 4.1, R2 = 0.18, p < 0.01). The effect of A. gerardi on soil microbial respiration also varied with site fertility and was inversely related to plant 85 .w u : 0>:: 0:033 :03::38: 33808 3:: :30:0:3 038330 :0 838:: 308:0 w u 2 .880 8:88: H :88 0:: 8:_:> 8:803 33808 E: 3:: £3083 038830 :0 838:: Am: 603838: 33828 5: .0::: 83:03:52 6: .0::: :03:N:::0:_8-Z Amv .:0w0::_: 0_::w:0:_ 3:0: A 0 UL O 0 Ba 0 0 LL "'7 V PL it ir F K :L A Figure 4.4. Soil origin and plant effects on the structure of the soil microbial community as measured by CLPP. CLPP patterns are distinguished between soil microbial communities from the A. gerardi (solid symbols) and no plant treatments (open symbols). Soil microbial communities in soils from different sites are indicated by symbols: circles, BA; triangles, FK; upside-down triangles, LL; diamonds, MK; stars, PL; pentagons, UL. Significance values from the RDA are listed in Table 4.4. 89 There were also some significant interactions between the presence of A. gerardi and soil origin (Table 4.3A), indicating that the presence of a plant did not have uniform effects on the CLPP profiles across all soils. In general, CLPP patterns in the presence of A. gerardi loaded lower on canonical PC] and canonical PC2; however, CLPP patterns from UL soil were markedly different from all other soils regardless of the presence of A. gerardi (Figure 4.4). The RDA for the PLF A profiles of the soil microbial community also revealed variation among the six sites and detected an effect of the presence of A. gerardi (Table 4.5, Figure 4.5). Axis 1 accounted for 20.8% of the variance in phospholipid fatty acid data, 47.2% of the variance in the phospholipid fatty acid-environment relationship and had a phospholipid fatty acid-environment correlation of 0.879 (Table 4.5B). Axis 2 accounted for 9.2% of the variance in phospholipid fatty acid data, 21.2% of the variance in the phospholipid fatty acid-environment relationship and had a phospholipid fatty acid-environment correlation of 0.890 (Table 4.5B). Axis 3 accounted for 5.1% of the variance in phospholipid fatty acid data, 11.6% of the variance in the phospholipid fatty acid—environment relationship and had a phospholipid fatty acid -environment correlation of 0.790 (Table 4.5B). Axis 4 accounted for 3.0% of the variance in phospholipid fatty acid data, 6.7% of the variance in the phospholipid fatty acid-environment relationship and had a phospholipid fatty acid-environment correlation of 0.630 (Table 4.5B). The presence of A. gerardi (Trt) was significantly related to the ordination of the PLFA profiles, as were all levels of soil origin (MK, BA, LL, PL, and FK), and the time at harvest (Table 4.5A). 90 Table 4.4. Fatty acids used in principal components analysis of PLFA profiles. I describe fatty acids using standard nomenclature where the total number of carbon atoms appears before the colon and the total number of C-C bonds appears after it. Cyclo-propane analogs are indicated by "cyclo," and the location of the epoxy bond is indicated by a "c" followed by two numbers. If the cis or trans configuration is unknown, the word "at" is used. The number following "cis," "trans," or "at" indicates the location of the double bond in relation to the carboxyl end of the molecule. Fatty acids with the same retention time are grouped as "sum in feature" and given a unique number designation. Phospholipid Fatty Acids C9 Dicarboxylic acid 14:0 iso 14:0 15:0 iso 15:0 anteiso 15:0 16:0 iso 16:1 cis 7 16:1 cis 9 16:1 cis 11 16:0 iso 17:1 G 17:0 iso 17:0 anteiso 17:0 cyclo 16:1 20H 18:1 9 trans alcohol 18:2 cis 12 18:1 cis 9 18:1 cis 13 18:0 19:1 at 11 alcohol 19:0 cyclo c11-12 19:0 cyclo 11-12 20H 22:0 22:0 20H 24:0 Coprostane Unknown 25.339 Summed feature 8: 18:1 trans 9 91 Table 4.5. Effect of soil origin, time of harvest, and presence of A. gerardi on the soil microbial community as detected by Distance-Based Redundancy Analysis of PLFA profiles. (A) Significance values for the permutation tests on the environmental factors of the RDA. (B) Variance explained by species data and species-environment correlations for the RDA. (A) Factor Lambda F -stat p-value % variance Plant 0.06 9.189 0.0010 6.4 Time 0.01 1.859 0.0540 1.3 Soil 0.31 9.037 0.0010 31.3 Plant*Time 0.09 4.351 0.0010 9.0 Soil*Plant 0.42 5.607 0.0010 42.1 Soil*Time 0.36 4.799 0.0010 36.4 All 0.51 3.285 0.0010 51.2 (B) Axis Eigenvalue Species-Environment Cum%Variance Cum%Variance Correlation of Species Data of Species- Environment 0.208 0.879 20.8 47.2 0.092 0.890 30.0 68.0 0.051 0.790 35.1 79.6 0.030 0.630 38.1 86.3 92 Canonical PC2, 9.2% I ;=‘¢ l D l 1’ o 0 v 0 0’ 90 <> 0 ,3 v K K ’ ; ’ . fl iii}; if? 4; O -2 -1 0 1 Canonical PCl, 20.8% 2 Androgogon S_0i|f 0 + MK 0 6 UL O 0 Ba 0 0 LL 7 V PL it? * FK it A Figure 4.5. Distance-based Redundancy Analysis of PLFA profiles, investigating the effects of soil origin, plant, and time effects on the structure of the soil microbial community. The PLFA profiles are distinguished between soil microbial communities from the A. gerardi (solid symbols) and no plant treatments (open symbols). Soil microbial communities in soils from different sites are indicated by symbols: circles, BA; triangles, FK; upside-down triangles, LL; diamonds, MK; stars, PL; pentagons, UL. Labels for phospholipid fatty acids are listed in Table 4.3. Significance values from the RDA are listed in Table 4.5. 93 Figure 4.5 shows the separation of samples by environmental factors and fatty acids. There was a significant interaction between the presence of A. gerardi and time at harvest (Table 4.5A, Plant*Time). Soils with plants tended to cluster higher on canonical PC1 (Figure 4.5). This was related to higher amounts of high carbon chain phospholipid fatty acids (22:0, 24:0, 22:0 20H, and unknown 25.339), indicating more eukaryotes were present in soils with plants. Additionally, there was also a significant interaction between the presence of A. gerardi and soil origin (Table 4.5A: Soil*Plant), indicating that the presence of a plant did not have uniform effects on the PLFA profiles across all soils. The higher fertility soils (Ba, LL, PL, and FK) had similar PLFA patterns and responded the same way to the presence of A. gerardi (an increase in canonical PC1, Figure 4.5); however, the PLFA patterns of the soils from the two low productivity sites did not change in response to the presence of A. gerardi and were different from the patterns of the high fertility sites (Figure 4.5). The UL site was different from all other sites in the amounts of some monounsaturated fatty acids on canonical PC2 (UL site had higher amounts of 18:1 cis 13 and lower amounts of summed in feature 8). Soils from the MK site showed a smaller elevation along canonical PC2 and increased numbers of eukaryotes (higher canonical PC 1 ). 94 Discussion Growing A. gerardi in these soils provided an independent assay of the potential productivity of each of these sites (A. gerardi production) and a direct test of the “plant” effect on soil microbial community structure and processes. This study suggests the origin of the soil and the presence of a plant both influence the structure and functioning of the soil microbial community. Most previous studies have been unable to distinguish between the effects of plants and the effects of soil origin on the structure of the soil microbial community. For example, Zelles et al. (1992) distinguished among the soil communities of grassland and agricultural fields using PLF A; however, both the soils and the plant communities differed among sites. Zelles et al. (1992) were able to distinguish among management regimes, but it was not possible to determine the relative effects of the soil versus the plants on these differences. In addition, those studies which have attempted to distinguish between soil and plant effects on the soil microbial community often did not measure the soil effect independent of any plant influence. Grayston and Campbell (1996) used CLPP patterns to differentiate between the soil microbial communities from the rhizospheres of hybrid larch (Larix eurolepis) and Sitka spruce (Picea sitchensis) trees in woodland and forest sites. However, the study does not estimate the magnitude of the plant effect on the CLPP patterns because there was no independent measure of the CLPP patterns of the soil microbial community in the absence of plants in these sites (Grayston and Campbell 1996). 95 Site Fertility Effects on Soil Microbial Processes In this study, soil origin had a significant influence on soil properties controlled by the soil microbial community and on the structure of the soil microbial community itself. This suggests that historical factors of the soil can have persistent effects on the soil microbial community, while the extant plant community is a major source of labile carbon and can influence the structure of the soil microbial community and, consequently, ecosystem functioning through the soil microbial community. I have presented site fertility as the driving factor explaining the relationships between soil origin and the soil and microbial properties that I measured in this study. However, other factors besides site fertility differed among these sites (Table 4.1). There were some differences in percent soil organic matter across sites, and species diversity varied dramatically among sites. Bossio et al. (1998) have shown that enrichment of organic matter through agricultural management produces recognizable differences in the PLFA patterns from the soil microbial communities from various management regimes (organic, low-input, and conventional farming). The results from this study do not change if I rank the sites by soil organic matter rather than field above-ground biomass (data not shown). Both soil organic matter and above-ground biomass are surrogates for site fertility. Bossio et al. (1998) suggested that higher soil organic matter should lead to greater soil microbial biomass, but this study does not support this assertion. Increased organic matter inputs generally occur in agricultural systems that are being managed organically; perhaps natural gradients in soil organic matter should not be 96 expected to reflect the same pattern because the systems have already had time to reach an equilibrium in soil organic matter turnover. Historical Plant Diversity E fleas on Soil Microbial Processes Another major difference among these field sites was the current plant species composition and diversity. There appears to be a unimodal relationship between above- ground plant biomass and species diversity across these six sites with low diversity, grass-dominated communities at both the lowest (MK) and highest (FK) fertility sites (Table 4.1). However, the soil and microbial properties of field MK were consistently more similar to the other low fertility site (UL) rather than FK, the other low species diversity site (Figure 4.2). This suggest that it is fertility more than diversity or composition that influences the soil microbial community. To better understand the impact of the global decline in species diversity due to human activities, many researchers have been investigating the relationship between species diversity and ecosystem function. The ‘rivet hypothesis’ proposes that each species contributes something unique to ecosystem function, and so ecosystem function declines as biodiversity declines (Ehrlich and Ehrlich 1981, Lawton 1984). A contrasting hypothesis suggests that species are redundant and that ecosystem function only declines when functional groups are missing from an ecosystem (Walker 1992, Lawton 1994). Finally, Lawton (1994) proposed that ecosystem fimction changes when a species is lost, but the direction and amount of that change are not predictable. 97 Observational studies and manipulative experiments investigating these theories have provided mixed results. My work in Chapter 2 (Broughton and Gross 2000) showed a significant effect of plant species diversity on the respiration or biomass of the soil microbial community, but these results were confounded with plant productivity and edaphic changes. My work at the Silwood, England BIODEPTH site (Chapter 3) found a positive relationship between plant diversity and two measures of microbial community structure (CLPP and PLFA). The Swiss BIODEPTH experiment has shown positive relationships between plant diversity and plant biomass (Spehn et al. 2000a, Spehn et al. 2000b), soil microbial respiration and functional diversity (Stephan et al. 2000), microbial biomass (Spehn et al. 2000a), and earthworm population density (Spehn et al. 2000a). In contrast, Wardle has consistently shown no relationship between diversity and ecosystem function in a series of plant removal experiments in New Zealand perennial grasslands (Wardle et al. 1999, Wardle et al. 2000, Wardle and Nicholson 1996). In a plant removal study in a North American grassland, Symstad et al. (1998) showed a positive relationship between plant species diversity and productivity, but no relationship between plant diversity and other ecosystem functions. Mikola and Setala (1998) found unpredictable ecosystem functioning responses to changes in species diversity when studying the phenomenon in a simple (three trophic level) decomposer food web from the soil of a pine forest in Finland. In this experiment, I found no relationship between species diversity and soil microbial processes, regardless of the presence of Andropogon gerardi. Soils from communities 98 with high plant species diversity did not have higher or more efficient ecosystem processes or higher soil microbial activities or biomass, as suggested by some recent theory (Loreau 2000). These results are consistent with Symstad et al.’s (1998) plant species removal study in which they detected the effect of declining species richness on productivity, but not nitrogen retention, suggesting that the relationship between species diversity and ecosystem processes is not necessarily consistent or predictable. In this study, the one exception was microbial biomass. The pots without A. gerardi showed a positive relationship between the microbial biomass and the plant species diversity of the communities from the soils were taken. This relationship disappeared when A. gerardi was grown in the soil, however, suggesting that the present plant community can have a large effect on the present soil microbial community. Several studies have shown plant composition (rather than plant diversity) effects on ecosystem functioning. The researchers reason that the quality of the carbon available to the microbial community is important and so the identity of the plant species providing that carbon should influence how ecosystem functions change (Paul and Clark 1996). Wardle et al. (1999) saw plant composition effects on PLFA patterns in the soils from a plant removal experiment in New Zealand grasslands, while Symstad et a1. (1998) found plant composition effects on productivity and nitrogen retention. Hector et al. (2000) detected a relationship between the species composition of litter and the decomposition rate at the Silwood Park, England BIODEPTH site. Hooper and Vitousek (1998) investigated the relationship between plant composition and nutrient cycling in experimental plots on serpentine soil in California. They determined that 99 plant community composition accounted for much more of the variation in nutrient cycling processes than did plant functional group diversity alone (Hooper and Vitousek 1998). Knowing the identity of the plant species involved allowed for a much better explanation of changes in inorganic N pools, soil moisture, microbial biomass, and microbial immobilization as a result of the experimental manipulations (Hooper and Vitousek 1998). I found some evidence to support this view in the relationships between plant composition and microbial respiration and soil PLFA patterns at the Silwood Park, England BIODEPTH site (Chapter 3). Historical plant composition effects may be contributing some of the variation in this study in the effects of soil origin on soil and microbial processes, but I are unable to test this assertion due to the design of the study. The ‘Plant’ Effect on Soil Microbial Processes Not only was soil origin important in structuring the soil microbial community in this experiment, the presence of a plant also had important effects on the structure and function of the soil microbial community. The presence of A. gerardi drove down nitrogen pools, but increased microbial respiration without affecting microbial biomass, suggesting an increase in the turnover rate of the soil microbial community and faster nutrient cycling. Other studies have shown plant effects on soil and microbial processes. Bachmann and Kinzel (1992) detected differences in the amounts of amino acids and sugars and the 100 rates of C02 evolution and some enzyme activities in the rhizosphere soils of six different plant species grown in four different soils. As in our study, Bachmann and Kinzel (1992) were able to detect strong plant effects regardless of soil origin and strong plant-soil interactions, although the magnitude and direction of change for enzyme activities and resource amounts were not consistent for different plant species. However, Bachmann and Kinzel (1992) did not measure nitrogen transformation rates, nitrogen pools, or microbial respiration or biomass. Groffrnan et al. (1996) investigated the relative roles of plant versus soil effects on the soil microbial community by measuring microbial biomass and activity and nitrogen transformation rates of soil taken from a range of old—field sites. The two experiments used various combinations of 10 plant species in monoculture and 4 soils that had been established for 4 years. Groffman et al. (1996) concluded that the main driver for microbial biomass and activity was soil type rather than plant species, although they suggested that plant effects might become more important afier a longer period of time. In this study, the magnitude of the plant effect on microbial respiration and total inorganic Nitrogen pools was larger in higher fertility soils than low fertility soils. One possible explanation for this result is that in lower fertility soils, there is a greater possibility that the resources supporting the soil microbial community are coming from the extant plant community. Consequently, the presence of a plant constantly providing resources to the soil microbial community could allow for a more active microbial community. This is in contrast to the possibility of increasing the soil microbial 101 biomass using the influx of new resources, which Bossio et al. (1998) suggest is often the result of increasing organic matter inputs. Plant and Soil Effects on Soil Microbial Community Structure Both soil origin and the presence of Andropogon gerardi had significant effects on potential metabolic diversity (measured with CLPP patterns) and phenotypic diversity (measured with PLFA profiles) of the soil microbial community in this study. Communities from field UL soils had strikingly different PLFA and CLPP patterns from communities grown in other soils; additionally, the presence of A. gerardi had no effects on CLPP and PLFA patterns for UL soils. The UL field is a highly diverse, low productivity site and the composition of the soil microbial community does not seem to respond quickly to changes in the plant community. The soil microbial communities detected in MK soil (another low fertility site) had PLFA patterns similar to field UL, but had similar CLPP patterns similar to the higher productivity sites. This indicates that structurally different microbial communities (from fields MK and UL) are capable of consuming the same resources. Other studies have found that community function may not change when community structure does. Buyer and Drinkwater (1997) detected differences in PLFA patterns between replicates of manipulations of different management treatments involving different crop residues but saw no differences in CLPP patterns. Similarly, IbekI and Kennedy (1998) 102 determined that PLFA profiles were more sensitive than CLPP patterns to differences in the soil microbial community grown in two different soils under six plant treatments. Like field UL, field MK is a low productivity field; however, MK also has low organic matter is dominated by a grass, Agropyron repens. If the long chain phospholipid fatty acids indicate mycorrhizae, the historical presence of A. repens plants in MK may explain why microbial communities incubated in MK soils have a higher proportion of eukaryotes than microbial communities grown in other soils, regardless of the presence of Andropogon gerardi. Although, the long chain phospholipid fatty acids may only be indicative of more plant material present in the soil. A central goal in ecology is to determine the factors controlling the abundance and distribution of species. Microorganisms in their natural habitats are only recently being studied in ecology. This study suggests that both the history of the soil and the plants presently growing in a community affect the structure and function of the soil microbial community. Further studies should eludicate the relative importance of these two factors in influencing the soil microbial community. 103 REFERENCES Bachmann, G. and H. Kinzel. 1992. Physiological and ecological aspects of the interactions between plant roots and rhizosphere soil. Soil Biology and Biochemistry 24: 543-552. Bligh E. G., W. J. Dyer. 1959. A rapid method of total lipid extraction and purification. Canadian Journal of Biochemistry and Physiology 37 : 911-917 Bossio, D. A. and K. M. Scow. 1998. Impacts of carbon and flooding on soil microbial communities: Phospholipid fatty acid profiles and substrate utilization patterns. Microbial Ecology 35: 265-278. Bossio, D. A., K. M. Scow, N. Gunapala, and K. J. Graham. 1998. Determinants of soil microbial communities: effects of agricultural management, season, and soil type on phospholipid fatty acid profiles. Microbial Ecology 36, 1-12. Broughton, LC. and K. L. Gross. 2000. Patterns of diversity in plant and soil microbial communities along a productivity gradient in a Michigan old-field. Oecologia 125: 420-427. Buckley, DH. and T. Schmidt. 2001. The structure of microbial communities in soil and the lasting impacts of cultivation. Microbial Ecology 42: 11-21. Burbank, D. H., K. S. Pregitzer, and K. L. Gross. 1992. Vegetation of the W. K. Kellogg Biological Station. Michigan State University Agricultural Experiment Station Research Report Number 510. Buyer, J. S. and L. E. Drinkwater. 1997. Comparison of substrate utilization assay and fatty acid analysis of soil microbial communities. Journal of Microbiological Methods 30: 3-11. Ehrlich, P. R. and A. H. Ehrlich. 1981. Extinction: The Causes and Consequences of the Disappearances of Species. Random House, New York. Foster, B. L. 1996. Plant competition and diversity in relation to productivity in old- field plant communities. Ph.D. Dissertation. Michigan State University, East Lansing, MI, USA. Goodfriend, W.L. 1998. Microbial community patterns of potential substrate utilization: a comparison of salt marsh, sand dune, and seawater-irrigated agronomic systems. Soil Biology and Biochemistry 30: 1169-1176. Grayston, S. J. and C. D. Campbell. 1996. Functional biodiversity of microbial communities in the rhizospheres of hybrid larch (Larix eurolepis) and Sitka spruce (Picea sitchensis). Tree Physiology 16: 1031-8. Groffrnan, P. M., P. Eagan, W. M. Sullivan, and J. L. Lemony. 1996. Grass species and soil type effects on microbial biomass and activity. Plant and Soil 183: 61- 67. 104 Hamilton, E. W., HI, and D. A. Frank. 2001. Can plants stimulate soil microbes and their own nutrient supply? Evidence from a grazing tolerant grass. Ecology 82: 2397-2402. Hector, A., A. J. Beale, A. Minns, S. J. Otway, and J. H. Lawton. 2000. Consequences of the reduction of plant diversity for litter decomposition: effects through litter quality and microenvironment. OIKOS 90: 357-371. Hooper, D. U., D. E. Bagel, V. K. Brown, L. Brossard, J. M. Daingerfield, D. H. Wall, D. A. Wardle, D. C. Coleman, K. E. Giller, P. Labelle, W. H. van deer Patten, P. C. de Reiter, J. Ruse, W. L. Silver, J. M. Tiedje, and V. Walters. 2000. Interactions between aboveground and belowground biodiversity in terrestrial ecosystems: patterns, mechanisms, and feedbacks. Bioscience 50: 1049-1061. Hooper, D. U. and P. M. Vitousek. 1998. Effects of plant composition and diversity on nutrient cycling. Ecological Monographs 68: 121-149. Bikeway, A. M. and A. C. Kennedy. 1998. Phospholipid fatty acid profiles and carbon utilization patterns for analysis of microbial community structure under field and greenhouse conditions. FEMS Microbiology Ecology 26: 151-163. Konopka, A., L. Oliver, and RF. Turco, Jr. 1998. The use of carbon substrate utilization patterns in environmental and ecological microbiology. Microbial Ecology 35: 103-115. Lawton, J. H. 1994. What do species do in ecosystems? OIKOS 39: 391-422. Legendre, P. and M. J. Anderson. 1999. Distance-based redundancy analysis: testing multispecies responses in multifactorial ecological experiments. Ecological Monographs 69: 1 —24. Legendre, P. and E. D. Gallagher. 2001. Ecologically meaningful transformations for ordination of species data. Oecologia (in press). Loreau, M. 2000. Biodiversity and ecosystem functioning: recent theoretical advances. OIKOS 91: 3-17. Metting, B. F., Jr. 1993. Structure and physiological ecology of soil microbial communities, pp. 3-25, In Soil Microbial Ecology: Applications in Agricultural and Environmental Management (ed. F. B. Metting, Jr.). Marcek Dekker, Inc., New York. Mikola, J. and H. Setala. 1998. Relating species diversity to ecosystem functioning: mechanistic backgrounds and experimental approach with a decomposer food web. OIKOS 83: 180-194. Nelson, D.W. and LE. Sommers. 1982. Total carbon, organic carbon, and organic matter, pp. 539-580, In_ Methods of Soil Analysis Part 2 - Chemical and Microbiological Properties. ASA Special Publ. No. 9. ASA, SSSA, Madison, WI. Ohtonen, R., S. Aikio, H. Vare. 1997. Ecological theories in soil biology. Soil Biology and Biochemistry 29: 1613-1619 105 Paul, EA. and FE. Clark. 1996. Soil Microbiology and Biochemistry. 2nd ed. San Diego, CA: Academic Press, Inc. Paul, E.A., D. Harris, M. J. Klug, and R. W. Ruess. 1999. The determination of microbial biomass, pp. 291-317, In Standard Soil Methods for Long-Term Ecological Research (eds G.P. Robertson, D.C. Coleman, C.S. Bledsoe, & P. Sollins) Oxford University Press, New York. Robertson G. P., J. R. Crum, and B. G. Ellis. 1993. The spatial variability of soil resources following long-term disturbance. Oecologia 96: 451-456 Schléipfer, F. and B. Schmid. 1999. Ecosystem effects of biodiversity: a classification of hypotheses and exploration of empirical results. Ecological Applications 9: 893-912. Spehn, E.M., J. Joshi, B. Schmid, J. Alphei, and C. Korner. 2000a. Plant diversity effects on soil heterotrophic activity in experimental grassland ecosystems. Plant and Soil 224: 217-230. Spehn, E.M., J. Joshi, B. Schmid , M. Diemer, and C. Komer. 2000b. Aboveground resource use increases with plant species richness in experimental grassland ecosystems. Functional Ecology 14: 326-337. Stephan, A., A. H. Meyer, and B. Schmid. 2000. Plant diversity affects culturable soil bacteria in experimental grassland communities. Journal of Ecology 88: 988- 998. Symstad, A. J., D. Tilman, J. Wilson, and J. M. H. Knops. 1998. Species loss and ecosystem functioning: effects of species identity and community composition. OIKOS 81: 389-397. Tiedje, J. M. 1995. Approaches to the comprehensive evaluation of prokaryote diversity of a habitat, pp. 73—87, In Microbial Diversity and Ecosystem Function (eds D. Allisopp, R.R. Colwell, and D.L. Hawksworth). CAB International, Wallingford, UK. Walker, B. H. 1992. Biodiversity and ecological redundancy. Conservation Biology 6: 18-23. Wardle, D.A., K.I. Bonner, and GM. Barker. 2000. Stability of ecosystem properties in response to above-ground functional group richness and composition. OIKOS 89: 11-23. Wardle, D.A., K.I. Bonner, G.M. Barker, G.W. Yeates, K.S. Nicholson, R.D. Bardgett, RN. Watson, and A. Ghani. 1999. Plant removals in perennial grassland: vegetation dynamics, decomposers, soil biodiversity, and ecosystem properties. Ecological Monographs 69: 535-568. Wardle, D. A. and K. S. Nicholson. 1996. Synergistic effects of grassland plant species on soil microbial biomass and activity: implications for ecosystem-level effects of enriched plant diversity. Functional Ecology 10: 410-416. Zak, J. C., M. R. Willig, D. L. Moorhead, and H. G. Wildman. 1994. Functional diversity of microbial communities: A quantitative approach. Soil Biology and Biochemistry 26: 1101-1108. 106 Zelles, L., R. Rackwitz, Q. Y. Bai, T. Beck, and F. Beese. 1995. Discrimination of microbial diversity by fatty acid profiles of phospholipid fatty acids and lipopolysaccharides in differently cultivated soils. Plant and Soil 170: 115-22. Zelles, L., Q.Y. Bai, T. Beck, and F. Beese. 1992. Signature fatty acids in phospholipid fatty acids and lipopolysaccharides as indicators of microbial biomass and community structure in agricultural soils. Soil Biology and Biochemistry 24: 317-323. 107 CHAPTER 5 AN EXPERIMENTAL EVALUATION OF THE EFFECTS OF DIFFERENT PLANT SPECIES ON THE STRUCTURE AND FUNCTION OF SOIL MICROBIAL COMMUNITIES Introduction Most of the carbon and nitrogen entering the soil results from litterfall, root exudates, or root death of plants (Paul and Clark 1996). Because organic inputs from plants species can differ in quantity, timing, and biochemistry, plant species identity has the potential to affect microbial process rates through litter quality and root exudation effects. As a result, the composition and productivity of the plant community may influence the soil microbial community. Conversely, the productivity or diversity of the plant community may be affected by processes mediated by soil microorganisms (e. g. N-mineralization rates). Consequently, changes in the plant community and the accompanying change in the soil microbial community potentially affect ecosystem function. To date, many studies have investigated the role of edaphic factors in structuring the soil microbial community, while fewer have addressed the effect of the plant community. Additionally, very few have sought to distinguish between the effects of soil and the effects of plants on the soil microbial community. Most studies of plant community effects on soil microbial communities confound direct plant effects with soil effects. For example, J .C. Zak et al. (1994) detected differences in the structure of the soil microbial community (using Biolog) along an elevational and moisture gradient 108 in the Chihuahuan Desert at the Jomada Long-Tenn Ecological Research site. Similarly, Goodfriend (1998) used Biolog to distinguish among the microbial communities at eight sites representing a variety of wetlands. However, in both of these studies, the plant communities sampled were from different sites in which there were likely concomitant changes in soil characteristics Consequently, the influence of plant community composition differences could not be assessed independently of differences in edaphic characteristics. It remains unclear how the plant community affects the soil microbial community relative to effects of soil environment on the microbial community. In an earlier greenhouse experiment I used field soils from six sites to distinguish the relative importance of direct soil effects versus plant-mediated soil effects on the soil microbial community in a controlled environment. Using a single species, Andropogon gerardi, I found that microbial communities grown in the six soils differed in both structure and function and that A. gerardi mediated those differences. To follow up on those results and determine if plant species differed in their effects on the structure and function of the soil microbial community, I conducted a greenhouse experiment in which soils from two old fields that differed in fertility were planted with all combinations of three plant species common to local old fields. I hypothesized that (1) plant species have unique effects on the structure and fiinction of soil microbial communities, (2) the effects of different plant species on soil microbial community structure and function are non- additive, and (3) soil microbial communities close to the roots (rhizosphere) are different from the communities in the bulk soil. To test whether soil microbial 109 communities in rhizosphere soil differed from those in bulk soil, 1 excluded roots from a cylinder of soil in each pot. Methods Site and Species Descriptions Soil was collected in October 1999 and October 2000 from two of the six successional old-fields at the W. K. Kellogg Biological Station in southwestern Michigan sampled for the experiment in Chapter 4. The sites differed in fertility, species diversity, dominant plant type, and years since abandonment, but both were located on Kalamazoo sandy loam soil (Table 5.1). To estimate plant species diversity, I determined species composition in six 0.5 m x 2 m plots at each site in August 1998. To estimate annual net primary productivity (ANPP), I clipped aboveground biomass at ground level from a 0.5 m x 0.5 m plot located within the plots used to assess species diversity. The Upper Louden (UL) field had moderate ANPP, high species diversity and was dominated by diverse forb communities. In contrast, Field K (FK) had high ANPP, low species diversity, and was dominated by Bromus inermis (Table 5.1). The UL field had been abandoned from agriculture for over fifty years. Field K was once used as pasture, but had been unmanaged for over twenty-five years. 110 Table 5.1 Plant species by functional group and aboveground biomass at the two field sites from which soil was collected. Functional groups are coded by Grass (G), Forb (F), Legume (L), and Woody (W). Plant species used in this greenhouse experiment are in bold. Field Species Functional Biomass % of Total group (g/mz) Biomass F K Bromus inermis G 589.6 98.7 A gropyron repens G 8.0_, 1.3 Total 597.6 UL Poa compressa G 48.8 15.6 T rifolium pratense L 34.4 1 1.0 Andropogon virginicus G 30.6 9.8 Hieracium sp. F 30.5 9.8 Danthonia spicata G 26.3 8.4 Rudbeckia hirta F 20.6 6.6 Solidago nemoralis F 20.0 6.4 Solidago canadensis F 14.4 4.6 Aster sp. F 14.4 4.6 Antennaria plantaginifolia F 9.4 3.0 Rubus sp. W 9.3 3.0 Panicum sp. G 8.1 2.6 Achillea millifolium F 8.0 2.6 Centaurea maculosa F 7.0 2.3 Chrysanthemum Ieucanthemum F 5.5 1.8 Aster sp. F 4.4 1.4 Rumex acetosella F 3.8 1.2 Panicum sp. G 3.8 1.2 Solidago speciosa or juncea F 3.7 1.2 Other (5 species) 3F,2G 9.2 . 2.9 Total 312.3 To maximize the possible differences among the plant species, I selected three perennial species that are representative of different functional groups in old-field communities: Solidago canadensis (forb), T rifolium pratense (legume), and Bromus inermis (grass). Solidago canadensis is a native herbaceous perennial dicot that is 111 commonly dominant in higher productivity fields of southwestern Michigan (Foster 1996, Werner et al. 1980). Trifolium pratense is an herbaceous perennial legume naturalized from Eurasia that occurs in fields across a broad range of productivities (Scoggin 1978b). Bromus inermis is a C3 perennial grass dominant in high productivity fields in southwestern Michigan and is naturalized from Eurasia (Scoggin 1978a). Experimental Design To distinguish direct plant effects versus plant-mediated soil effects on the soil microbial community, I grew each species alone and in all combinations in soil from the two sites described above in a greenhouse experiment. I also included no-plant controls. There were root exclosures in each pot to separate bulk from rhizosphere soil. Soil for these experiments was collected from both field sites from the top 15 cm in the same area from which species diversity and plant biomass were sampled. Because of sample processing constraints, I conducted the experiment in three time blocks. Soil collected in December 1999 was used for the first two time blocks, while soil collected in October 2000 was used for the third time block. The soil was sieved to 4 mm and mixed to reduce variability. Soil was stored at room temperature until the experiment was established in the greenhouse (less than 2 months). I used a randomized complete block factorial design for the greenhouse experiment in which I varied soil (2 sources - Soil) and plant species (3 species in monoculture and all 112 combinations, including no plants), with 3 replicates of each treatment (Replicate) at 3 different times (Time). The plant treatments used are described in Table 5.2. Soils were placed in 5 cm diameter x 20 cm deep pots and kept well-watered with de-ionized water to avoid adding nutrients or contaminants. Treatments were randomly assigned to pots within replicate. To compare soil microbial communities between bulk and rhizosphere soil, I included one root exclosure tube in each pot. Root exclosures were sewn into 15 cm x 2 cm diameter cylinders from 20-micron mesh, and the seams were sealed with silicone sealant. The exclosure tube was filled with soil and placed in the center of the pot. All plant species were added as 1 month old seedlings; all seedlings were germinated in a sterile sand medium in a grth chamber and were less than 2 cm in height at transplantation. One plant per species was added to each pot; for mixtures, seedlings were planted at equal distances from each other with the root exclosure in the center. If the transplant was unsuccessful, the seedling was replaced for up to four weeks into the experiment. Each segment of the experiment ran for a total of 18 weeks to ensure that the roots had filled the pot. I ran the experiment in three time blocks because of the number of samples and the time required to process each sample. Thus, time was an additional 113 Table 5.2. Treatments used in the Randomized Complete Block Design to test for the effects of plant species on the soil microbial community. Each treatment was replicated three times within three Time blocks on two different soils. Species planted were G = Bromus inermis, F = Solidago canadensis, L = T rifolium pratense. 0 = absent, + = present, C = control (no plants). Plant Treatment Bromus inermis Solidago canadensis T rifolium pratense C O O 0 G + 0 0 F O + O L 0 0 + FL 0 + + GF + + 0 GL + O + GFL + + + 114 factor in the ANOVA. The first three replicates grew from December 1999 until May 2000, the second three replicates grew from February to June 2000, and the third set of three replicates grew from November 2000 until March 2001. Temperature in the greenhouse from 25 to 40 °C; light availability was controlled through a 12 h light/ 12 h dark cycle. Data Collection Differences among treatments were assessed by (1) plant growth, (2) the inorganic nitrogen pool, and (3) N-mineralization rate. At harvest, I determined the dry mass of above and belowground plant biomass. I separated root biomass from the soil (and other species’ roots) during sieving. To avoid confusing roots from different plant species, root systems were kept intact. Roots were rinsed thoroughly in de-ionized water before drying. The root and shoot biomass for each species were measured separately. All plant material was dried at 60°C for 48 h. Shoot biomass was dried separately from root biomass. After the roots were removed, soil from the entire pot (rhizosphere influenced) was passed through a 2 mm sieve and stored in sealed plastic bags at 4°C until analyzed. Soil from the exclosures was sieved and stored separately. All analyses were done within 3 days of sampling, except PLFA. Soil for PLFA analyses was kept at —80°C until the fatty acids were extracted. 115 I determined gravimetric soil moisture, total inorganic nitrogen, and N-mineralization and nitrification rates for each sample as described in Chapter 4. I assessed differences in soil microbial community production among treatments by (1) microbial biomass C and (2) microbial respiration. The method used to determine microbial respiration and the chloroform fumigation incubation method used to determine microbial biomass are described in Chapter 4 (see also Paul et al. 1999). I assessed soil microbial community structure differences among treatments by PLFA profiles as described in Chapter 4. For analysis, I included only those phospholipid fatty acids that were present in greater than 50% of samples and reported their abundance as the square root of the proportion of the total phospholipid fatty acid amount in each sample (Hellinger transformation). Of the 70 lipids detected, 31 phospholipid fatty acids met this criterion for the entire experiment (n = 144) and 34 phospholipid fatty acids met this criterion for the subset (n = 58) analyzed with root exclosures (Table 5.4). Statistical Analyses I used a randomized complete block design (ANOVA) model to test the effects of soil origin, presence/absence of each plant species, and time on the following response variables: soil moisture, plant biomass, relative growth rate, total N, N-mineralization rate, nitrification rate, microbial respiration, and microbial biomass (Table 5.3). The plant treatments were analyzed in three ways. (1) I included all 8 plant treatments as independent factors in the ANOVA (Table 5.3A) to assess the effect of each plant treatment on the soil and microbial characteristics. (2) Because in Chapter 4 I found 116 large effects of the presence of a plant on many soil and microbial variables, to determine the effects of the plants on soil and microbial characteristics, I ran the ANOVA with only the 7 plant treatments, excluding the no-plant control (Table 5.3B). (3) To determine the compositional effects of Bromus inermis, Solidago canadensis, and T rifolium pratense on soil and microbial characteristics, I used a 2 x 2 x 2 factorial design for the plant treatments in the ANOVA (Table 5.3C). As carbon should be limiting to the microorganisms, the size of the microbial community should depend on the input of carbon to the system. Because the plant material is the primary source of new carbon to microorganisms, then the microbial biomass should be proportional to the belowground plant biomass in any one spot. To determine if significant treatment effects on soil community structure or function were the indirect results of changes in plant biomass, I also used root, shoot, and total plant biomass as co-variates in all three analyses. 1 used a modified redundancy analysis (RDA, Legendre & Anderson 1999) to determine the relationship between the environmental factors (soil origin, presence/absence of each plant species, time of harvest) and the PLF A measure of soil microbial community structure. I transformed the PLFA data using a Hellinger transformation because this transformation does a good job of handling data matrices with many zero values (Legendre & Gallagher, in press). The RDA procedure involved: (1) the creation of a matrix of dummy variables corresponding to the randomized complete block design (modeled from the experimental design: soil origin, presence/absence of each plant, time at harvest), (2) redundancy analysis of the 117 relationship between the principal coordinates (matrix of optical density or phospholipid fatty acid data) and the environmental variables (matrix of dummy variables in (1)), and (3) implementation of a Monte Carlo permutation test to estimate the statistical relationship between the two matrices (Legendre & Anderson 1999). This analysis allows me to test which factors from the experimental design (soil origin, presence/absence of each plant, time at harvest) are significantly related to the variation in the PLFA patterns. The modified RDA is a better statistical technique than regular ordination techniques because it allows for significance testing. Results Plant Communities As presented in Chapter 4, the two field sites differed significantly in peak aboveground plant biomass (live and litter) and species composition (Table 5.1). The Upper Louden (UL) field had 320 i 23 g/m2 aboveground plant biomass and 15.8 i 1.0 species per square meter, while Field K (FK) had 592 i 22 g/m2 aboveground plant biomass and 1.3 i 0.2 species per square meter. Field K was dominated by Bromus inennis with very little Agropyron repens (Table 5.1). Upper Louden field supported a complex forb-dominated community, which included Poa compressa, T rifolium pratense, Andropogon virginicus, several Hieracium species, Danthonia spicata, Rudbeckia hirta, several Solidago species, Antennaria plantaginifolia, Achillea 118 millefolium, Centauria maculosa, Chrysanthemum leucanthemum, several Panicum species, and a species of Rubus (Table 5.1). Time Effects on Soil and Microbial Processes Time significantly affected most soil and microbial variables (Table 5.3A). The first two time groups used soil collected in 1999, while the last time group used soil collected in 2000. The last experimental group (time 3) had significantly higher total inorganic nitrogen pools (p < 0.001, F = 83.3), N-mineralization rates (p < 0.001, F = 43.5), and nitrification rates (p < 0.001, F = 39.4). Microbial respiration (p< 0.001, F = 8.0) and microbial biomass (p < 0.001, F = 17.8) also differed significantly among time blocks (Table 5.3A). The time effect seems to be the result of the two different soil collection times. Because patterns were consistent among time blocks and there were no significant interactions between Time and other factors, I did not investigate the Time effects in any further detail. Plant Species Responses to Soil Types Plants grown in soil from field FK, the higher fertility site, attained higher biomass than those grown in soil from field UL (Figure 5.1A-C, Table 5.3B). There were also significant differences among the seven plant treatments in total plant biomass (Figure 5.1A, p < 0.001, F=5.2), root biomass (Figure 5.2A, p < 0.001, F = 11.2), and shoot biomass (p< 0.05, F = 2.6). In FK soil, Bromus inermis had the highest biomass and 119 .OZ< 0:: 80:: 82:> 00::0::8w:m 8808 0:: 83:> AU: ::: Am: :0: ”:88 :::::::m a: ::08 0:: 82:> A3 :0: .33: :0::0»: 883 AU :8: Aviv v: 20:: Am 80:: :0: :0: 80880:: 8:8 :3 80:00:: 8:8 ::: ”Gm 03: :. 8 0:: 80:00 80880:: 80880:: 8:8 ::: 88:0 =0: 9:: :3 :0:00 0808:: 23:03:: :8: 88030800 0:038: 8220:: 880%: :0 8:803 8:3 :80 :1 ._.m 0:53: AD 0 0808:: 88:03:: 9 vi 0 8.8080500 0:00:20: 0 x=Om 8.8202: 880:: I 80880:. 8:7: 3:040 :0 A: A : O _ fl _ o o _ — d A d _ a E < l l a. 8 l _ - m w u e -. e .H. m. j .. N 8 N 8 r + 8 m. m. W m. w m. + w B B e r r m «w m m w r + w .m .0 .0. U m < _ L : _ : _ _ v v _ _ : : : : : 3:: 5:53 :25: 0:: v: 20:: 80880; 8:i 3:640 :0 A: A : O 80888:. 8:5 4:030 :0 A: A : O 120 <0: 2%,: a :22: 2: <>0z< 2:: 82: 328 0083583 48:08 0:: 838 A0: :8: Am: :0: ”:80 ::::8::: H 8:08 0:: m08_:> A43 :0 : .33 80:80: 08:3 :0: :8: CE: V: 20:: Am: 80:: :0: :0: 80880:: 8:8 :3 360:: 8:8 :8: ”N6 03: H 8: 0:: 8:00 80880:: 808::08 8:8 :8: 83:0 :0: £3 :3 :0:00 0808.8: 80:05:: :8: 88000800 0:00.00: 88:08.: 880$ :0 3:803 :00: .m.m 0:83: 0:80:08: 88:08:: 8 88000200 0:00.00: a 88:08 880$ I 80880; 8:7: 4:040 :0 A: A : 0 E: <<< r m<- m< _ _ _ h _ _ _ 3:: 80:84 :0:—:3 D O. ‘0. O. — o o (8) ssewoyg 1003 V! I— O. N wfiDSH—NDPH «GM—Am 3:0A0:0x:: 1: : 0 0:: v: 22.8 <3. V? O. '— O O (8) ssmuoyg 100)] V? v-! o. N ADO vi. £Om 80880; 8:5 3:040:0 A: H : 0 :_:__J. 86 0.0 md o.: W: o.m (8) ssmuogg 1003 121 Table 5.3. Summary of ANOVA results analyzing effect of soil origin and (A) individual plant treatments, (B) individual plant treatments without no-plant controls, or (C) factorial plant treatments on plant and soil and microbial variables. Species codes in 3C are: BROIN, Bromus inermis; SOOCA, Solidago canadensis; TRFPR, T rifolium pratense. NS = not significant, p > 0.05. ANCOVA results, using total plant biomass as the covariate, are shown in parentheses if the effect changed in significance. (A) Variable SOIL TRT TIME SOIL*TRT Total Biomass (g) < 0.001 < 0.001 < 0.05 < 0.001 Root Biomass (g) < 0.001 < 0.001 < 0.001 < 0.001 Shoot Biomass (g) < 0.001 < 0.001 NS < 0.001 Root to Shoot Ratio <0.01 < 0.001 NS < 0.05 Total Inorganic < 0.05 (NS) < 0.001 < 0.001 NS Nitrogen N-mineralization Rate < 0.01 NS < 0.001 NS (< 0.05) Nitrification Rate < 0.01 NS < 0.001 NS (< 0.05) Microbial respiration < 0.001 < 0.001 < 0.001 < 0.01 (< 0.01) (< 0.05) Microbial Biomass NS NS < 0.001 NS 122 (B) Variable SOIL TRT TIME SOIL*TRT Total Biomass (g) < 0.001 < 0.001 < 0.05 NS Root Biomass (g) < 0.001 < 0.00] < 0.00] < 0.05 Shoot Biomass (g) < 0.001 < 0.05 NS < 0.001 Root to Shoot Ratio <0.01 < 0.001 < 0.05 NS (NS) Total Inorganic Nitrogen < 0.05 NS < 0.001 NS N-mineralization Rate < 0.01 NS < 0.001 NS (< 0.05) Nitrification Rate < 0.01 NS < 0.001 NS (< 0.05) Microbial respiration < 0.001 NS < 0.01 < 0.05 (< 0.01) Microbial Biomass NS NS < 0.001 NS 123 (C) Variable SOIL BROIN SOOCA TRIPR TIME Total Biomass (g) < 0.001 < 0.001 < 0.05 < 0.001 < 0.05 Root Biomass (g) < 0.001 < 0.001 NS < 0.01 < 0.001 Shoot Biomass (g) < 0.001 < 0.01 < 0.001 < 0.001 NS Root to Shoot Ratio < 0.001 < 0.001 < 0.05 < 0.05 < 0.05 Total Inorganic NS < 0.001 < 0.001 < 0.01 < 0.001 Nitrogen (< 0.05) (< 0.01) (< 0.05) N-mineralization < 0.001 NS NS < 0.05 < 0.001 Rate (NS) (NS) Nitrification Rate < 0.001 NS NS < 0.05 < 0.001 (NS) (NS) Microbial respiration < 0.001 < 0.01 < 0.001 < 0.01 < 0.001 (< 0.05) (NS) (< 0.01) (NS) (< 0.05) Microbial Biomass NS NS NS NS < 0.001 124 T rifolium pratense had the next highest biomass (Figure 5.1A-C). In UL soil, Bromus inermis and Solidago canadensis had higher biomass than T rifolium pratense (Figure 5.1A-C). However, the diversity effect on productivity was non-additive: mixtures did not have higher biomass than monocultures (Figure 5.1A). Bromus inermis and T rifolium pratense dominated mixtures in FK soil (Figire 5.1B), while Bromus inermis tended to dominate the mixtures in UL soil (Figure 5.1C). Plants grown in the higher fertility soil (FK) had higher root biomass than those grown in soil from field UL (Figure 5.2A-C, p < 0.001, F = 21.9). Bromus inermis had the highest root biomass in both PK and UL soils (Figure 5.2B,C, Table 5.3C). Bromus inermis and T rifolium pratense grown in the lower fertility UL soil had higher root to shoot ratios than plants grown in FK soil (Table 5.4). Bromus inermis, the grass, had two times higher root to shoot ratios than both Solidago canadensis and T rifolium pratense, regardless of soil origin (p < 0.001, F = 42.5). Plant species did not appear to change their root to shoot ratios in response to the presence of competitors (Table 5.4). Effects of Soil Origin on Soil and Microbial Processes Soils from the FK site had significantly higher total inorganic nitrogen pools (p < 0.05, F = 5.7, Figure 5.3A), N-mineralization rates (p < 0.01, F = 9.0, Figure 5.38), nitrification rates, (p < 0.01, F = 8.5, Figure 5.3C), and microbial respiration (p < 0.001, 125 cc.— 2.— id 4&0 No.0 H Sod Smo H two @od H co; imp—U 8.. 558 . 83on m cote Emucfim H 26H ocd m_.onm.N 4+ No.0 H ovd wm.oH3.N 4+ SEE 8m mo:_m> mmd H «5.— hmd H mud m+ mod H end Omd H ow.— m+ 8.5332 86on N m_.ova._ :dH hwd omd H mwd deofio m_.onm._ nod H cod 3.0 H 05m 9:51-96:62 cod H nod 006 H 05.0 :deo; 2.5—:35: 3V 3.2295. sea: EV flutmhomoo Sagan 6v 2::on §Soxm SE 5.53 .5an 3V 3.28on naught CV nwwtmnomoo assay. 6V .3:me 3 Beam 0:: v. 22m .Um.m 038i E 3%: 2m <>OZ< 2: Eofi mos—g oogocEflm .EoEEu: 28 :8 some 5 $6on :83 some .59 8:8 80% 8 Hood .vw 2an 126 .OZ< 2: Bot 82? 353$in :88 Emvcfim H SEE 2m 82a> Ame—”Eu 30:2: 15 98 @386 3:5 vi ”8% oz: 2: 50253 uozflzwcumfi 2m 8353 diet? some com .3285 3582:. Am: 95 death—v.2 3580:: R: .28 :ocaomtzc ADV £88 :EENmEoEEi Amv .coon: oanSE .88 2V E 355235 :83 wcoEa c252; .m.m 2sz H5860; Ema E258; EmE iEO 40 LO :5 A m 0 U .50 40 m0 ‘5 i— n— O U u — _ q A d 1 — O — d a N A m ) \.l 10 WHO i . 18. cm W i W i 3 M. D m. i . Av 1w :0 m . O m + + 3 m.. i r 0 -SN z. m.. r 6 a / e 3 m... AV co .1 / .l 10— p H L. % Av S 8 .M 9 i a icon Pm. is L. m. w L. m w i . + + is mm as. m. x 5 o + + i: ../i m. r .. (U. vi 0 I * w. u m 58 .. . oi: M _ _ p _ _ _ p r cam _ _ _ _ _ _ p F m— EoEEoci Ema E0823; EmE 25:58:. EmE .5040 LO 4... i— n. O U iEOi—O m0 i: i— m D U iEOi—O LO 4“— i. m 0 U . _ d O ai 9 O _ 4 _ S — a O Q% ) C H i: e 6 10— l 1 Aw * 1N aim 1. +% IN ml .N I . + fl ION) m9 + N W co W. I v V 10m“ H... v 0 m ..v N u no u 1 i s/ U i 3 m r Low N m co 0 D. mi I now /c.o T P ic P m. i + S M m. cm. w ,M m m n... n a room m. i in mun T iwmm i 13mm. ../.. m. D. 1 an H n H i P a .m e 1 3(aw l lO—rw l lO—( m r. 1 9 i . 52 u U r m < . L _ _ ~ _ _ _ N~ _ _ _ _ _ _ _ N_ _ _ _ L _ L _ . : 127 F = 22.6, Figure 5.3D) than those from the UL sites. However, soil origin did not affect soil microbial biomass (Table 5.3A, Figure 5.3B). The differences in soil and microbial variables between the two sites were partly due to differences in plant biomass produced on the two soils (Figure 5.1A-C, Table 5.3A,C). When the analyses were re-run with total plant biomass as a covariate, total inorganic Nitrogen was no longer affected by soil origin (Table 5.3A) and N-mineralization and nitrification rates were not significantly influenced by soil origin (Table 5.3C). In contrast, when root biomass was the covariate, total inorganic Nitrogen (p < 0.05, F = 4.5) and N-mineralizaton (p < 0.01, F = 8.6) and nitrification rates (p < 0.01, F = 8.3) were still significantly affected by soil origin. Plant Effects on Soil and Microbial Processes The presence of any plant significantly reduced total inorganic Nitrogen (Figure 5.3A, p < 0.001, F = 13.0) and significantly increased microbial respiration (Figure 5.3D, p < 0.001, F = 8.7) on both soils. However, the composition of the plant community had no effect on either of these variables and there were no significant plant treatment effects when the no-plant controls were excluded from the analysis (Table 5.3B). This experimental design allowed me to test for the effects of the presence of specific plant species on these processes (Table 5.3C). The presence of T rifolium pratense was significantly correlated with higher N-mineralization rates (Figure 5.3B, Figure 5.4B,E, 128 p < 0.05, F = 5.1) and nitrification rates (Figure 5.3C, p < 0.05, F = 5.1) and this effect was still detectable when root biomass was used as the covariate (p < 0.05, F = 4.5; p < 0.05, F = 4.5, respectively). However, increasing the number of plant species (or functional groups, the same in this design) did not have additive effects on any of the soil or microbial processes. Mixtures did not have detectably higher nitrogen pools or process rates than monocultures (Figures 5.3A-E). Interactive E fleets of Plant Species and Soil Origin on Soil and Microbial Processes In this study there were few significant interactions between soil origin and plant treatments that affected soil and microbial processes (Table 5.3A-B). Plant biomass was significantly affected by the interaction between soil origin and plant treatments (Figure 5.1A, p < 0.001, F = 28.3). For example, Solidago canadensis had the lowest total biomass in monoculture in FK soil, while T rifolium pratense had the lowest total biomass in monoculture in UL soil (Figure 5.1A). Additionally, microbial respiration was significantly affected by the interaction between soil origin and plant treatment (Table 5.3A, p < 0.01, F = 3.1), primarily because the control, Solidago canadensis monoculture, and three species mixture (GFL) did not differ significantly in microbial respiration between the two soils, PK and UL (Figure 5.3D). Although there was no significant interaction between soil origin and plant treatment for N-mineralization and nitrification rates (Table 5.3A,B), in mixtures in the FK soil both tended to decrease in the presence of Solidago canadensis and increase in the presence of T rifolium pratense, 129 but were less responsive in the UL soil (Figure 5.3B-C). Total plant biomass and root biomass had no effect on this relationship. Plant and Soil Effects on Soil Microbial Community Structure PLF A profiles of the soil microbial community varied between the soils from the two sites and also responded to the presence of the three plant species (Table 5.6A). Soil origin accounted for 12% of the variation in PLF A profiles (p < 0.001, F = 24.4), while Time accounted for 10% (p < 0.001, F = 20.4) and Replicate an additional 5% of the variation (p < 0.001, F = 10.6). The three plant species had small but significant effects PLF A profiles, each accounting for only 1% of the variation (Table 5.6A). Axis I from the Redundancy Analysis accounted for 20.7% of the variance in phospholipid fatty acid data, 41.1% of the variance in the phospholipid fatty acid -environment relationship and had a phospholipid fatty acid -environment correlation of 0.853 (Table 5.58). Axis 2 accounted for 12.4% of the variance in phospholipid fatty acid data, 25.2% of the variance in the phospholipid fatty acid -environment relationship and had a phospholipid fatty acid -environment correlation of 0.852 (Table 5.6B). Axis 3 and Axis 4 accounted for 5.1% and 2.5% of the variance in phospholipid fatty acid data, respectively (Table 5.68). Figure 5.4 shows the separation of samples by environmental factors and fatty acids as related to the axes from the Redundancy Analysis. PLFA profiles fi‘om UL soils loaded higher on PC2 than profiles from FK soils (Figure 5.4). UL soils had higher amounts 130 Table 5.5. Fatty acids used in principal components analysis of PLFA profiles. I describe fatty acids using standard nomenclature where the total number of carbon atoms appears before the colon and the total number of CC bonds appears after it. Cyclo-propane analogs are indicated by "cyclo," and the location of the epoxy bond is indicated by a "c" followed by two numbers. If the cis or trans configuration is unknown, the word "at" is used. The number following "cis," "trans," or "at" indicates the location of the double bond in relation to the carboxyl end of the molecule. Fatty acids with the same retention time are grouped as "sum in feature" and given a unique number designation. Phospholipid Fatty Acids Phospholipid Fatty Acids (Full Experiment) (Exclosure subset) C9 Dicarboxylic acid C9 Dicarboxylic acid 14:0 14:0 15:0 iso 15:0 iso 15:0 anteiso 15:0 anteiso 15:0 15:0 16:0 iso 16:0 iso 16:1 cis7 16:1cis7 16:1cis9 16:1cis9 16:1 cis 11 16:1 cis 11 16:0 16:0 iso 17:1 G iso 17:1 G 17:0 iso 17:0 iso 17:0 anteiso 17:0 anteiso 17:0 cyclo 17:0 cyclo 16:1 2OH 17:0 18:1 trans 9 alcohol 16:1 20H 18:2 cis 12 18:2 cis 12 18:1cis9 18:1cis9 18:1cis 13 18:1 cis 13 18:0 18:0 19:0 cyclo c11-12 19:1 trans 11 20:0 19:0 cyclo cl 1-12 19:0 cyclo 11-12 20H 20:4 cis 14 22:0 20:0 22:0 2OH 19:0 cyclo 11-12 20H 24:0 22:0 23:0 20H 22:0 20H Coprostane 24:0 Unknown 25.339 23:0 20H Cholesteryl-palmitate Coprostane summed feature 8: 18:1 trans 9 Unknown 25.339 Cholesteryl-palmitate summed feature 8: 18:1 trans 9 - summed feature 12: 131 2 c,\" ‘1: 1— §0_il. 2 Plant FK UL N" C o o 8 G A 2.. Te 0" F v .9. L 4 <1 8 FL .1 O 5 CF I D U -1" GL O O GFL * H _2 1 l 1 -2 -1 O 1 2 Canonical PC1, 20.7% Figure 5.4. Distance-based Redundancy Analysis of Phospholipid Fatty Acid profiles, investigating the effects of soil origin, plant, and time effects on the structure of the soil microbial community. Canonical principal component plot for the full RDA. The PLFA profiles are distinguished between soil microbial communities from FK soil (solid symbols) and UL soil (open symbols). Soil microbial communities in soils with different plant communities are indicated by symbols: circles. controls; triangles, monocultures; four or five-sided polygons, mixtures. Labels for phospholipids are listed in Table 5.4. Significance values from the RDA are listed in Table 5.5A. 132 . 2 mo 3 £88 9.35.3386 omonmom. Amv .30 8:3th be: can £3253: Eatobfi Ea omehaom 6:30:98 .Ewto =8 53> 2 90 C mos MESS 359$ $3 :8an «mm-E mo marge 5:32:68 o5 Bow Bigot moo—Q Eoconfioo RENEE. 303058 :2th 6-5 .m.m oSwE owtfioxm ESE-bus flatmhoaoo omohzom $5on 353%. + o . + o . + o . H H MOI H H MOI ‘ H “O: r ivd- i + ind- Lio- - $ -3- m - -3- m - -3- i imd- m. i iod m. i imd- 1 -_.o- m i L6 m i L6- 1 iod M i 1N6 M i iod - -3 .0 - -3 .0 - -3 i Lmd 1. r Lac .0 i imd - -3 m .. -3 WW - -3 r + oil. 1 m 19o i <-v.o _ _ md _ L 5o _ _ md %Z'l ‘ I Dd 19311101193 133 Table 5.6. Effect of soil origin, time, replicate, and presence of Bromus inermis (BROIN), Solidago canadensis (SOOCA), and T rifolium pratense (TRFPR) on the soil microbial community as detected by Distance-Based Redundancy Analysis (db-RDA) of PLFA profiles. (A) Significance values for the permutation tests on the environmental factors of the RDA. (B) Variance explained by species data and species- environment correlations for the RDA. (A) Factor Lambda F—stat p-value % variance Soil 0.12 24.4 0.001 12.2 BROIN 0.01 2.3 0.017 1.1 SOOCA 0.01 1.9 0.050 0.9 TRFPR 0.01 2.5 0.010 1.2 Time 0.10 20.4 0.001 10.3 Rep 0.05 10.6 0.001 5.3 Sum 0.50 2.7 0.001 (B) Axis Eigenvalue Species-Environment Cum % Variance Cum % Variance of Correlation of Species Data Species-Environment 1 0.207 0.853 20.7 41.1 2 0.124 0.852 33.1 66.3 3 0.051 0.761 38.2 76.5 4 0.025 0.589 40.6 81.4 134 of 18:1 cis 13, 16:1 cis 11, and 18:2 cis 12, indicating a higher fungal to bacterial ratio and more gram-negative bacteria (Harwood and Russell 1984). Partial canonical principal components derived from the redundancy analysis of PLFA profiles revealed that soil microbial communities grown under Bromus inermis had higher amounts of 16:1 cis 9, 16:1 cis 11, 18:1 cis 9, and summed in feature 8 (Figure 5.5A). Communities grown in the presence of Solidago canadensis had higher 18:1 cis 13, 16:1 cis 7, 16:1 cis 11, and summed in feature 8 (Figure 5.5B), while communities grown in the presence of T rifolium pratense had higher 18:1 cis 13, summed in feature 8, 16:0 iso, iso 17:1 G, 16:1 cis 9, 17:0 cyclo, and 18:1 cis 9 (Figure 5.5C). This indicates that the microbial communities under all three plants had a higher proportion of gram-negative bacteria than microbial communities in the no-plant controls. In addition, microbial communities grown in the presence of T rifolium pratense had a higher proportion of gram-positive bacteria than soil microbial communities not exposed to T rzfolium pratense. Overall there were few differences in PLFA profiles among replicates (within time) or time. A Comparison of the Microbial Communities within and outside the Root Exclosures I used root exclosures to exclude rhizosphere effects and create “bulk” soil in these pots. I compared the structure of the soil microbial community of the “bulk” soil to that of the rhizosphere soil using PLFA profiles. The RDA of the PLFA profiles showed no 135 Canonical PC2, 10.1% 3 I I <5 2 L— _ O . Exclosure ' o + 1 — w .- 301! 0 Q UL O a. ““ PK 0 A F» r? 0 H :2" ‘ L13: ;- _ Er ( (1 -1 l — -2 -l 2 Canonical PC1, 28.9% Figure 5.6. Distance-based Redundancy Analysis of Phospholipid Fatty Acid profiles, investigating the effects of exclosure, soil origin, plant, and time effects on the structure of the soil microbial community. Canonical principal component plot for the full RDA. The PLFA profiles are distinguished between soil microbial communities from FK soil (solid symbols) and UL soil (open symbols). 136 Table 5.7. Effect of exclosure, soil origin, time, and presence of Bromus inermis (BROIN), Solidago canadensis (SOOCA), and T rifolium pratense (TRF PR) on the soil microbial community as detected by Distance-Based Redundancy Analysis (db-RDA) of PLF A profiles. (A) Significance values for the permutation tests on the environmental factors of the RDA. (B) Variance explained by species data and species- environment correlations for the RDA. (A) Factor Lambda F-stat p-value % variance Exclosure 0.02 1.4 0.110 1.8 Soil 0.25 20.0 0.001 24.8 _ BROIN 0.02 1.8 0.060 2.2 ‘ SOOCA 0.01 0.9 0.520 1.1 ‘i TRFPR 0.03 2.6 0.002 3.3 Time 0.02 1.6 0.060 2.0 - Sum 0.79 2.0 0.001 (B) Axis Eigenvalue Species-Environment Cum%Variance Cum%Variance Correlation of Species Data of Species- Environment 1 0.289 0.976 28.9 36.8 2 0.101 0.920 39.0 49.6 3 0.085 0.898 47.5 60.4 4 0.066 0.907 54.1 68.8 137 effect of root exclosure on the structure of the soil microbial community. However, like RDA of the PLFA profiles for the whole experiment (above), the RDA for the PLFA profiles of the soil microbial communities within and outside the root exclosures also revealed variation between sites and detected an effect of the presence of T rifolium pratense on the soil microbial community (Table 5.7, Figure 5.6). Axis I accounted for 28.9% of the variance in phospholipid fatty acid data, while Axis 2 accounted for 10.1% of the variance in phospholipid fatty acid data (Table 5.68). Soil origin accounted for 25% of the variation in PLFA profiles (Figure 5.6, p < 0.001, F = 20.0), while Trifolium pratense accounted for 3% of the variation (p < 0.01, F = 2.6). However, there was no effect of root exclosure on PLFA patterns (Figure 5.6), indicating that resources that had influenced the microbial community in these pots were able to travel through the soil and away from the roots. Differences detected among plant species were not due to direct contact with the roots. Discussion 1 expected to find that plant species had unique effects on the structure and function of soil microbial communities and that the effects of different plant species on soil microbial community structure and function are non-additive. I found that plant species had unique effects on soil microbial community structure (as indicated by PLFA) and some soil processes like total inorganic Nitrogen pools, N-mineralization and nitrification rates, and microbial respiration. However, I did not find additive effects of plant diversity, and soil effects were much stronger than individual plant effects on the 138 soil microbial community. This study suggests the origin of the soil and the presence of a plant both independently influence the structure and functioning of the soil microbial community. Most previous studies have been unable to distinguish between the effects of plants and the effects of soil origin on the structure of the soil microbial community. As I saw in the previous chapter, soil process rates were higher and nitrogen pools were lower in the higher fertility soil (FK). However, because of the longer time frame during which these experiments were conducted (4 months in Chapter 4 versus 16 months in this study), Time did significantly affect soil and microbial process rates. The presence of a plant significantly affected soil and microbial processes and soil microbial community structure. In addition, the effects of different plant species on soil microbial community structure and function were non-additive. Process rates from mixtures could not be determined by summing process rates from monocultures of the plant species included in the mixtures. Finally, contrary to our hypothesis, soil microbial communities from the root exclosures did not differ structurally (as measured by PLF A) from soil microbial communities in soil closely associated with plant roots. The Effects of Soil Origin on Soil and Microbial Processes and Soil Microbial Community Structure In this study, soil origin had a significant influence on soil properties controlled by the soil microbial community and on the structure of the soil microbial community itself. 139 In fact, soil origin explained more variation in soil microbial community structure (PLFA profiles) than any of the other explanatory variables. This suggests that the history of the soil and plant community plays a major role in structuring the soil microbial community and, consequently, influences ecosystem functioning through the soil microbial community. Many studies have shown site or soil characteristics to be an important influence on soil microbial processes and structure. Zelles et al. (1992) differentiated among eight agricultural management treatments using PLFA. Zelles et a1. (1995) distinguished among three different soils in farmland and grassland using PLFA. Bossio et al. (1998) determined that soil type was the most important environmental factor in structuring the soil microbial communities of sustainable agriculture systems in California. Groffrnan et al. (1996) measured microbial biomass and activity and nitrogen transformation rates of soil taken from a range of old-field sites. As with our study, Groffrnan et al. (1996) concluded that the main driver for microbial biomass and activity was soil type. Plant Effects on Soil and Microbial Processes and Soil Microbial Community Structure As I saw in Chapter 2 (Broughton and Gross 2000), Chapter 3 (Broughton and Gross 2001), and Chapter 4, in this study, plants can have significant effects on soil and microbial processes and soil microbial community structure. Plant effects are complex and include effects that will be mediated through diversity, species composition, and individual plant species effects. This study cannot distinguish among the effects of 140 species diversity, functional group diversity, and plant community composition as aspects of the overall plant effect because each functional group is represented by only one species. However, I can address whether species differ in their effects on soil and microbial characteristics, whether plant species effects are detectable in mixtures, and whether these effects are enhanced by diversity. Which component or components (diversity, composition, individual plant species) are most important influencing the soil microbial community can be addressed by other studies like the BIODEPTH experiment in Chapter 3. The results from this study do not support the hypothesis that increased plant species diversity leads to increased soil and microbial processes because monocultures differed, but mixtures did not. The mixtures did not have higher process rates than the monocultures. Similarly, Wardle (Wardle et al. 1999, Wardle et al. 2000, Wardle and Nicholson 1996) has consistently shown no relationship between diversity and ecosystem fiinction in a series of plant removal experiments in New Zealand perennial grasslands. Symstad et al. (1998) also found no relationship between plant diversity and ecosystem functions (other than productivity) in a plant removal study in a North American grassland. However, other experiments investigating the relationship between plant diversity and ecosystem function have shown variation in the relationship between plant diversity and ecosystem function across habitats. In Chapter 2, Broughton and Gross (2000) found that there was a significant effect of plant species diversity on the respiration or 141 biomass of the soil microbial community. However, in that study, changes in diversity were correlated with changes in edaphic variables. My work at the Silwood, England BIODEPTH site (Chapter 3) also showed a significant relationship between plant diversity and two measures of microbial community structure (CLPP and PLFA). Similarly, results from the Swiss BIODEPTH experiment have shown positive relationships between plant diversity and plant biomass (Spehn et al. 2000a, Spehn et al. 2000b), soil microbial respiration and functional diversity (Stephan et al. 2000), microbial biomass (Spehn et al. 2000a), and earthworm population density (Spehn et al. 2000a) In Chapter 4, I saw that soil effects alone can influence soil and microbial processes, but these effects could be mediated by plants. In this study, the plant effect on soil and microbial processes seemed to be limited only to the presence of a plant: the identity of the plant mattered very little for soil processes in this study. The exception was T rzfolium pratense, which did affect inorganic nitrogen pools and nitrogen process rates. In addition, I did find small differences among the plant treatments in their effects on soil microbial community structure. At the Swiss BIODEPTH site, Stephan et al. (2000) also detected a legume effect on soil processes. T rifolium repens significantly increased diversity and activity of catabolic profiles of the soil microbial community (Stephan et al. 2000). Many studies have shown plant composition effects on ecosystem functioning. Symstad et al. (1998) found plant composition effects on productivity and nitrogen retention, and Wardle et al. (1999) observed plant composition effects on PLFA patterns in the soils fiom a plant removal experiment in 142 New Zealand grasslands. Hector et al. (2000) detected a relationship between the species composition of litter and the decomposition rate at the Silwood Park, England BIODEPTH site. Hooper and Vitousek (1998) determined that plant community composition accounted for much more of the variation in nutrient cycling processes on serpentine soil in California than just plant functional group diversity. Broughton and Gross found some evidence to support this view in the relationships between plant composition and microbial respiration and soil PLFA patterns at the Silwood Park, England BIODEPTH site (Chapter 3). Plant community composition should influence the soil microbial community through inputs of carbon into the soil. The quality and/or quantity of the carbon available to the microorganisms should influence which microorganisms thrive in a particular environment: therefore, the identity of the plant species providing that carbon should influence how ecosystem functions change (Paul and Clark 1996). Plants provide carbon to the soil in two ways: (1) litter and (2) root exudation. Hector et al. (2000) found large effects of plant litter composition on litter chemistry and decomposition rate at the Silwood Park, BIODEPTH site. However, this study did not last long enough to test litter effects. All differences in soil and microbial processes and soil microbial community structure among plant treatments must have been driven by differences in root exudation and root turnover. A recent study by Hamilton and Frank (2001) showed that herbivory could cause plants to stimulate rhizospheric microbial communities to increase nitrogen cycling and make nitrogen more available to the plants. 143 Several studies have detected differences among the rhizosphere soil microbial communities from different plant species. Grayston and Campbell (1996) distinguished between the rhizosphere of hybrid larch (Larix eurolepis) and Sitka spruce (Picea sitchensis) trees using CLPP at a woodland site and two plantations. Garland (1996) and Westover et al. (1997) also were able to differentiate the rhizosphere communities of several herbaceous plant species both in the field and greenhouse using Biolog. Miethling et al. (2000) used CLPP and PLFA profiles to distinguish among soil microbial communities from different fields planted with both alfalfa (Medicago sativa) and rye (Secale cereale). Miethling et al. (2000) determined that plant species identity was the most important factor in determining microbial community characteristics in the rhizosphere. Bachmann and Kinzel (1992) showed that plant species exude different amounts of organic metabolites, indicating that individual plants might have quite different effects on the soil microbial community. These studies suggest that plant species differences in root exudates may be important in distinguishing the composition of the rhizosphere communities. The results from my study also indicate that it is possible to distinguish among soil microbial communities grown in the presence of different plant species, as I could distinguish among the plant treatments with PLFA profiles. Plant species effects on soil microbial processes are ofien detected in greenhouse studies; however, these effects are much more difficult to find in the field. Buckley and Schmidt (2001) found that there was little difference between the soil microbial 144 communities of a continuously tilled agricultural site and a companion successional site in southwestern Michigan that had different plant communities for 12 years. Plant species effects on soil microbial processes may take a long time to manifest themselves (Buckley and Schmidt 2001, Broughton and Gross 2000), despite the possibility that plant species may need only to change the process rates of a small proportion of the soil organic matter to have large effects on soil processes (Wedin and Pastor 1993). Implications of the Similarity of Soil Microbial Communities in the Root Exclosures and the Rhizosphere Soil In this experiment, all of the soil in the pot was available to the plant roots, except the soil in the exclosure in the center. Root production was high in the experiment. Most of the plants in the experiment were root-bound by harvest time. The 20—micron mesh of the exclosure allowed nutrients, water, microorganisms, and mycorrhizae to pass through, but not plant roots. Consequently, there was no direct “contact” effect of plants on the soil and the soil microbial communities inside the root exclosure. In effect, the soil outside the exclosure was entirely rhizosphere soil, while the soil inside the exclosure was the equivalent of bulk soil. However, I could not detect any differences (in PLFA profiles) in the soil microbial communities in the two different types of soil (rhizosphere and bulk). This implies that the soil microbial communities in the rhizosphere and bulk soils were not significantly different in structure. This suggests that root exudates that can influence soil microbial community structure moved freely from the soil near the plant roots to the soil within the root exclosure. 145 This effect implies that in field settings I often do not find plant effects in the bulk soil because the resources (exudates) are used by rhizosphere microorganisms before the resources have a chance to migrate away from the roots. The soil microbial community plays a crucial role in the flow of nutrients and energy through the ecosystem. Changes in ecosystem function are intimately tied to the composition and activity of the soil microbial community. This study has provided evidence that soil is the most important factor influencing soil microbial communities, but the extant plant community composition can also influence soil microbial community structure. Better understanding the role microorganisms play in the overall functioning of the ecosystem can only improve our ability to predict how these ecosystems will change in the future under the influences of changing vegetation cover, agricultural practices, invasive species, and other dynamic drivers of the plant-soil ecosystem. 146 Wm‘ REFERENCES Bachmann, G. and H. Kinzel. 1992. Physiological and ecological aspects of the interactions between plant roots and rhizosphere soil. Soil Biology and Biochemistry 24: 543-552. Bossio, D. A. and K. M. Scow. 1998. Impacts of carbon and flooding on soil microbial communities: Phospholipid fatty acid profiles and substrate utilization patterns. Microbial Ecology 35: 265-278. Bossio, D. A., K. M. Scow, N. Gunapala, and K. J. Graham. 1998. Determinants of soil microbial communities: effects of agricultural management, season, and soil type on phospholipid fatty acid profiles. Microbial Ecology 36, 1-12. Burbank, D. H., K. S. Pregitzer, and K. L. Gross. 1992. Vegetation of the W. K. Kellogg Biological Station. Michigan State University Agricultural Experiment Station Research Report Number 510. Buyer, J. S. and L. E. Drinkwater. 1997. Comparison of substrate utilization assay and fatty acid analysis of soil microbial communities. Journal of Microbiological Methods 30: 3-11. Foster, B. L. 1996. Plant competition and diversity in relation to productivity in old- field plant communities. Ph.D. Dissertation. Michigan State University, East Lansing, MI, USA. Goodfriend, W.L. 1998. Microbial community patterns of potential substrate utilization: a comparison of salt marsh, sand dune, and seawater-irrigated agronomic systems. Soil Biology and Biochemistry 30: 1169-1176. Grayston, S. J. and C. D. Campbell. 1996. Functional biodiversity of microbial communities in the rhizospheres of hybrid larch (Larix eurolepis) and Sitka spruce (Picea sitchensis). Tree Physiology 16: 1031-8. Groffman, P. M., P. Eagan, W. M. Sullivan, and J. L. Lemunyon. 1996. Grass species and soil type effects on microbial biomass and activity. Plant and Soil 183: 61- 67. Haack, S.K., H. Garchow, D.A. Odelson, L.J. Fomey, and M.J. Klug. 1994. Accuracy, reproducibility, and interpretation of fatty acid methyl ester profiles of model bacterial communities. Applied and Environmental Microbiology 60: 2483- 2493. 147 Hector, A., A. J. Beale, A. Minns, S. J. Otway, and J. H. Lawton. 2000. Consequences of the reduction of plant diversity for litter decomposition: effects through litter quality and microenvironment. OIKOS 90: 357-371. Hooper, D. U. and P. M. Vitousek. 1998. Effects of plant composition and diversity on nutrient cycling. Ecological Monographs 68: 121-149. Legendre, P. and M. J. Anderson. 1999. Distance-based redundancy analysis: testing multispecies responses in multifactorial ecological experiments. Ecological Monographs 69: l —24. Legendre, P. and E. D. Gallagher. 2001. Ecologically meaningful transformations for ordination of species data. Oecologia (in press). Miethling, R., G. Wieland, H. Backhaus, and C. C. Tebbe. 2000. Variation of microbial rhizosphere communities in response to crop species, soil origin, and inoculation with Sinorhizobiom meliloti L33. Microbial Ecology 41: 43-56. Nelson, D.W. and LE. Sommers. 1982. Total carbon, organic carbon, and organic matter, pp. 539-580, In_ Methods of Soil Analysis Part 2 - Chemical and Microbiological Properties. ASA Special Publ. No. 9. ASA, SSSA, Madison, WI. Ohtonen, R., S. Aikio, H. Wire. 1997. Ecological theories in soil biology. Soil Biology and Biochemistry 29: 1613-1619 Paul, EA. and FE. Clark. 1996. Soil Microbiology and Biochemistry. 2nd ed. San Diego, CA: Academic Press, Inc. Paul, E.A., Harris, D., Klug, M.J., and Ruess, R.W. (1999) The determination of microbial biomass, pp. 291-317, In Standard Soil Methods for Long-Term Ecological Research (eds G.P. Robertson, D.C. Coleman, C.S. Bledsoe, & P. Sollins) Oxford University Press, New York. Scoggin, H. J. 1978a. The Flora of Canada. Pt 2: Pteridophyta, Gymnospermae, Monocyotyledoneae. National Museum of Canada. Crown Copyrights, Ottawa, Canada. Scoggin, H. J. 1978b. The Flora of Canada. Pt 3: Dicotyledoneae (Sauruaceae to Violaceae). National Museum of Canada. Crown Copyrights, Ottawa, Canada. Spehn, E.M., Joshi, J., Schmid, B., Alphei, J., & Korner, C. (2000) Plant diversity effects on soil heterotrophic activity in experimental grassland ecosystems. Plant and Soil 224, 217-230. 148 Spehn, E.M., Joshi, J., Schmid, B., Diemer, M., & Komer, C. (2000) Aboveground resource use increases with plant species richness in experimental grassland ecosystems. Functional Ecology 14, 326-337. Stephan, A., A. H. Meyer, and B. Schmid. 2000. Plant diversity affects culturable soil bacteria in experimental grassland communities. Journal of Ecology 88: 988- 998. Symstad, A. J., D. Tilman, J. Wilson, and J. M. H. Kn0ps. 1998. Species loss and ecosystem functioning: effects of species identity and community composition. OIKOS 81: 389-397. Tunlid, A., and DC. White. 1992. Biochemical analysis of biomass, community structure, nutritional status, and metabolic activity of microbial communities in soil, pp. 229-262, In Soil Biochemistry, vol. 7 (eds G. Stotsky and J.M. Bollag). Wardle, D.A., K.I. Bonner, and GM. Barker. 2000. Stability of ecosystem properties in response to above-ground functional group richness and composition. OIKOS 89: 11-23. Wardle, D.A., K.I. Bonner, G.M. Barker, G.W. Yeates, K.S. Nicholson, R.D. Bardgett, R.N. Watson, and A. Ghani. 1999. Plant removals in perennial grassland: vegetation dynamics, decomposers, soil biodiversity, and ecosystem properties. Ecological Monographs 69: 535-568. Wardle, D. A. and K. S. Nicholson. 1996. Synergistic effects of grassland plant species on soil microbial biomass and activity: implications for ecosystem-level effects of enriched plant diversity. Functional Ecology 10: 410-416. Wedin, D. A. and J. Pastor. 1993. Nitrogen mineralization dynamics in grass monocultures. Oecologia 96: 186-192. Werner, P. A., I. K. Bradbury, and R. S. Gross. 1980. The biology of Canadian weeds. 54. Solidago canadensis L. Canadian Journal of Plant Science 60: 1393-1409. Westover, K. M., A. C. Kennedy, and S. E. Kelley. 1997. Patterns of rhizosphere microbial community structure associated with co-occurring plant species. Journal of Ecology 85: 863-873. Zak, J. C., M. R. Willig, D. L. Moorhead, and H. G. Wildman. 1994. Functional diversity of microbial communities: A quantitative approach. Soil Biol. Biochem. 26: 1101-1108. Zelles, L., R. Rackwitz, Q. Y. Bai, T. Beck, and F. Beese. 1995. Discrimination of microbial diversity by fatty acid profiles of phospholipid fatty acids and lipopolysaccharides in differently cultivated soils. Plant and Soil 170: 115-22. 149 Zelles, L., Q.Y. Bai, T. Beck, and F. Beese. 1992. Signature fatty acids in phospholipid fatty acids and lipopolysaccharides as indicators of microbial biomass and community structure in agricultural soils. Soil Biology and Biochemistry 24: 317-323. I. r 150 CHAPTER 6 SUMMARY From the collection of field surveys and manipulative experiments in the previous chapters I can make several conclusions. This dissertation suggests the origin of the soil and the presence of a plant both influence the structure and functioning of the soil microbial community. However, plant effects were often not as strong as I had expected and there were other factors influencing the structure and function of the soil microbial community. First, I found that legacy effects can last for years. The disturbance caused during site preparation at the Silwood Park BIODEPTH site was still influencing the soil microbial community four years later (Chapter 3). Plant effects on the soil microbial community were often the result of plant biomass. In Chapter 2, plant productivity had significant effects on soil microbial respiration. In Chapter 3 many plant effects on soil microbial community processes were associated with larger amounts of plant matter rather than the diversity of the plant community. As I saw in Chapter 2 (Broughton and Gross 2000), Chapter 3 (Broughton and Gross 2001), Chapter 4, and Chapter 5, plants can have significant effects on soil and microbial processes and soil microbial community structure. Plant effects are complex and include effects that will be mediated through productivity (Chapters 2 and 3), 151 diversity (Chapter 3), species composition (Chapters 3 and 5), and individual plant species effects (Chapters 4 and 5). In Chapter 4, I saw that soil effects alone can influence soil and microbial processes, but these effects could be mediated by plants. In Chapter 5, the plant effect on soil and microbial processes seemed to be limited only to the presence of a plant: the identity of the plant mattered very little soil processes in this study. The exception was T rifolium pratense, which did affect inorganic nitrogen pools and nitrogen process rates. In addition, I did find small differences among the plant treatments in their effects on soil microbial community structure. However, these unique effects of plant species did not seem to be additive in mixtures. Finally, root exclosures had no effect on the soil microbial community structure, indicating plant resources can migrate away from the roots. The large number of findings in which plant effects are limited to rhizosphere soils in the field imply that these plant resources are used before they have the opportunity to migrate away from the roots in the field. The soil microbial community is a complex part of the old-field ecosystem that controls many important ecological processes. A multitude of factors influence the structure and firnction of this community. Understanding the roles each aspect of the community plays in structuring the soil microbial community will allow better understanding of the effects of global change on these important processes. 152