Wright State University CORE Scholar

Browse all Theses and Dissertations Theses and Dissertations

2018

River structure and function in a landscape modified by agriculture: implications for primary consumers

Hannah M. Fazekas Wright State University

Follow this and additional works at: https://corescholar.libraries.wright.edu/etd_all

Part of the Environmental Sciences Commons

Repository Citation Fazekas, Hannah M., "River biofilm structure and function in a resource landscape modified yb agriculture: implications for primary consumers" (2018). Browse all Theses and Dissertations. 2244. https://corescholar.libraries.wright.edu/etd_all/2244

This Dissertation is brought to you for free and open access by the Theses and Dissertations at CORE Scholar. It has been accepted for inclusion in Browse all Theses and Dissertations by an authorized administrator of CORE Scholar. For more information, please contact [email protected]. RIVER BIOFILM STRUCTURE AND FUNCTION IN A RESOURCE LANDSCAPE MODIFIED BY AGRICULTURE: IMPLICATIONS FOR PRIMARY CONSUMERS

A Dissertation in partial fulfillment of the requirements for the degree of Doctor of Philosophy

by

HANNAH M. FAZEKAS B.S., Saginaw Valley State University, MI, 2012

2018 Wright State University

COPYRIGHT BY HANNAH M. FAZEKAS 2018

WRIGHT STATE UNIVERSITY GRADUATE SCHOOL July 18, 2018

I HEREBY RECOMMEND THAT THE DISSERTATION PREPARED UNDER MY SUPERVISION BY Hannah M Fazekas ENTITLED River biofilm structure and function in a resource landscape modified by agriculture: implications for primary consumers BE ACCEPTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF Doctor of Philosophy. ______Yvonne Vadeboncoeur, Ph.D. Dissertation Director

______Donald F. Cippolini, Ph.D. Director, Environmental Sciences Ph.D. Program

______Barry Milligan, Ph.D. Interim Dean of the Graduate School Committee on Final Examination:

______Lynn K. Hartzler, Ph.D.

______John O. Stireman III, Ph.D.

______Katie Hossler, Ph.D.

______David L. Strayer, Ph.D.

ABSTRACT

Fazekas, Hannah M. PhD., Environmental Sciences PhD Program, Wright State University, 2018. River biofilm structure and function in a resource landscape modified by agriculture: implications for primary consumers.

Anthropogenic alterations to , , and phosphorus bioavailability have increased the flux of these resources into the and altered stream function. Streams modify the transport of these resources to receiving through uptake, transformation, and mineralization. Understanding how streams process carbon, nitrogen, and phosphorus can provide insight about how stream ecosystems function in landscapes where human modification is inescapable. I investigated how land use in agricultural regions affect resource availability to primary producers and consumers and the subsequent impact on stream processes. I surveyed headwater streams in three Lake Erie watersheds to determine spatiotemporal limitation of attached . I found that low-order streams exhibit phosphorus limitation and the severity of phosphorus limitation was greatest post-fertilizer application when the imbalance between nitrogen: phosphorus concentrations was greatest.

These results suggest that biofilm nutrient uptake responded to landscape level influences and attached algae actively sequestered phosphorus from the water column. Agriculture alters the availability of carbon through modification of riparian vegetation. I used genomic techniques to describe longitudinal changes in microbial

iii

composition along a stream with headwaters that lacked riparian vegetation due to row crop agriculture but the width of the forested riparian area increased downstream. The relative of the most abundant microbial phyla varied along physical and chemical (light, phosphorus concentration) gradients. Land use affected physical- chemical characteristics of the river, which in turn, influenced sediment microbial community composition. The removal of riparian forested vegetation in agricultural streams leads to an increased availability of light to attached algae. I investigated the effect of attached algal on consumers across an experimental gradient in light intensity. Attached algal productivity and production were coupled across the light gradient. I also studied how land use influenced carbon resource use by common macroinvertebrate functional feeding groups in Midwestern streams. I found that invertebrates consistently used attached algal carbon. This reliance was not affected by riparian vegetation nor the percent of the watershed dedicated to agriculture. Futhermore, structure remained similar across the gradient in land use. This work demonstrates that attached assemblages in streams respond to landscape level processes that propagate to consumers.

iv

TABLE OF CONTENTS

Chapter One ...... 1 Introduction ...... 1 Literature Cited ...... 5 Chapter Two...... 10 SPATIOTEMPORAL VARIATION IN NUTRIENT LIMITATION OF EPILITHIC IN LOW-ORDER STREAMS IN THE LAKE ERIE WATERSHED ...... 10 Introduction ...... 10 Methods...... 14 Site Description ...... 14 Physicochemical characteristics ...... 16 Nutrient Concentrations ...... 16 Biofilm ...... 17 Biofilm Stoichiometry ...... 18 Eco-enzyme activities ...... 18 Statistical Analysis ...... 19 Nutrient limitation ...... 20 Results ...... 20 Physicochemical Characteristics ...... 20 Biofilm composition ...... 21 Periphyton Stoichiometry and Eco-enzyme Activity ...... 22 Drivers of Biofilm Biomass and Enzyme Activities ...... 22 Nutrient limitation ...... 23 Longitudinal changes in nutrient concentrations ...... 24 Discussion ...... 25 Literature Cited ...... 31 Chapter Three...... 55

v

SEDIMENT MICROBIAL ASSEMBLAGE COMPOSITION IN A HETEROGENEOUS RESOURCE LANDSCAPE: LONGITUDINAL TRANSITIONS IN BIOFILM COMMUNITIES ...... 55 Introduction ...... 55 Methods...... 59 Study site ...... 59 Physicochemical Characteristics ...... 60 Sediment Size Distribution ...... 61 Organic Component of Sediment ...... 62 Sediment Energy Content ...... 62 Attached algal biomass and total organic matter content ...... 63 Sediment stoichiometry ...... 63 Microbial community composition ...... 63 Microbial community biomass ...... 64 Statistical Analyses ...... 65 Results ...... 66 Sediment heterogeneity and environmental conditions ...... 66 Variability of organic matter and biofilm biomass ...... 67 Microbial diversity ...... 67 Microbial community composition ...... 68 Discussion ...... 70 Literature Cited ...... 76 Chapter Four ...... 98 LIGHT AVAILABILITY DETERMINES ATTACHED ALGAL PRODUCTIVY AND BENTHIC CONSUMER GROWTH AND PRODUCTION IN EXPERIMENTAL MESOCOSMS ...... 98 Introduction ...... 98 Methods...... 100 Overview ...... 100 Experimental Setup ...... 101

vi

Organic matter characterization ...... 102 Primary Productivity ...... 103 Secondary Production ...... 103 Statistical Analysis ...... 106 Results ...... 106 Biofilms on sediments ...... 106 Light intensity and ...... 107 Chironomid density and larval production ...... 108 Discussion ...... 109 Literature Cited ...... 115 Chapter Five ...... 130 PRIMARY CONSUMERS IN AGRICULTURAL STREAMS RELY ON CARBON FIXED BY ATTACHED ALGAE, INDEPENDENT OF THE DENSITY OF RIPARIAN VEGETATION ...... 130 Introduction ...... 130 Methods...... 134 Sample Collection and Preparation ...... 134 Metazoans ...... 135 Water Quality ...... 136 Periphyton ...... 136 Stable-isotopes, stoichiometry, and fatty acids ...... 137 Site selection and land use characterization ...... 138 Statistical Analysis ...... 139 Results ...... 142 Land use and water chemistry ...... 142 Periphyton biomass and stoichiometry ...... 142 Stable isotopes of primary resources and consumers ...... 143 Total lipids of periphyton and consumers ...... 144 Discussion ...... 145

vii

Literature Cited ...... 153

viii

LIST OF FIGURES

Figure 2.1. Map of sampling locations for three watersheds within the Lake Erie Watershed drainage basin and land use from the National Land Cover Database (2011). Maumee River Watershed: Blue Creek, Eagle Creek, Flat Rock Creek, Lye Creek, and Ottawa Creek. Portage River Watershed: Middle Branch Portage River and North Branch Portage River. Grand River Watershed: Baughman Creek, Center Creek, Phelps Creek, and Swine Creek. Land use categories: crop agriculture (brown), pasture (yellow), forest (green), developed (red), water (blue)...... 49 Figure 2.2. Mean algal biomass as chlorophyll-a (mg/m2) and periphyton organic matter content (AFDM in g/m2) for each watershed post-fertilizer application, pre-harvest, and post-harvest. Horizontal lines represent median values and boxes represent inter- quartile ranges (25-75% percentiles). Whiskers show the range and points represent outlier values...... 50 Figure 2.3. Mean site periphyton algal biomass (chlorophyll-a in mg/m2) and sediment content (in g/m2) for post-fertilization, pre-harvest, and post-harvest in log-log scale...... 51 Figure 2.4. Canonical correspondence analysis of biofilms for the Maumee (squares), Portage (triangles), and Grand (circles) River watersheds post-fertilizer (black), pre- harvest (dark grey), and post-harvest (light grey). Response variables are biofilm chlorophyll-a (mg/m2), ash free dry mass (g/m2), percent organic matter, sediment content (g/m2), and P content (g/m2) and the predictor variables soluble reactive phosphorus (SRP in µg/L), nitrate (NO3, in µg/L), nitrite (NO2, in µg/L), water column N:P (NP), and total phosphorus (µg/L)...... 52 Figure 2.5. Canonical correspondence analysis of biofilms for the Maumee (squares), Portage (triangles), and Grand (circles) River watersheds post-fertilizer (black), pre- harvest (dark grey), and post-harvest (light grey). Response variables are biofilm extracellular enzyme activities (phosphatase, leucine aminopeptidase, β-glucosidase, and β-xylosidase in nmol/g C/h) and enzyme ratios (PHOS:LAMP and BG:BX) and the predictor variables are water column SRP and NO3, and biofilm chlorophyll-a, AFDM, sediment content (Ash), and percent organic matter (POM)...... 53 Figure 2.6. The difference in water column soluble reactive phosphorus concentrations (SRP, µg/L) between upstream sites in this study and the main stem rivers for the Portage and Maumee River. Stream abbreviations: Maumee River Watershed: Blue Creek = Blue, Eagle Creek = Eagle, Flat Rock Creek = Flat, Lye Creek = Lye, Ottawa Creek = Ottawa, Portage River Watershed: Middle Branch Portage River = Middle, North Branch Portage River = North...... 54

ix

Figure 3.1. Map of sampling locations along the Little Miami River and land use from the National Land Cover Database (2011). Land use categories: crop agriculture (brown), pasture (yellow), forest (green), developed (red), water (blue)...... 91 Figure 3.2. Photographs of each sampling location. Upstream site (LAt), next site downstream (laT), site 3 (lat) and the furthest site downstream (LAT)...... 92 Figure 3.3. Soluble reactive phosphorus concentration in µg/L (a, SRP), sediment organic matter content in mg/m2 (b), and chlorophyll-a in mg/m2 (c) across percent light. Reported p-values in the text are for the overall ANOVA. Within a single graph, boxes with the same letter above them are not significantly different (p < 0.05, Tukey’s HSD test). Horizontal lines represent median values and boxes represent inter-quartile ranges (25-75% percentiles) and whiskers show the range...... 93 Figure 3.4. Sediment fatty acid content among stream sites. Quantitative fatty acids are expressed as µg fatty acid per mg dry mass of sediment, not organic matter. Algal biomass includes . Reported p-values in the text are for the overall ANOVA. Within a single graph, boxes with the same letter above them are not significantly different (p < 0.05, Tukey’s HSD test). Horizontal lines represent median values and boxes represent inter-quartile ranges (25-75% percentiles) and whiskers show the range...... 94 Figure 3.5. Rank-abundance curve of the microbial communities across the four sites (a) using relative abundance data. Community composition of the most abundant phyla or classes (for Proteobacteria) of the stream sediments expressed as relative abundances (b)...... 95 Figure 3.6. The first two axes of the principal component analysis (PC1, PC2) synthesizing sediment microbial community composition. Abbreviations: Cyano = Cyanobacteria, Bacter = Bacteroidetes, Plancto = Planctomycetes, Unk = Unassigned, Chloro = Chloroflexi, and Alpha, Beta, Delta, Epsilon, and Gamma = Proteobacteria classes...... 96 Figure 3.7. CCA of microbial relative abundances with environmental predictors soluble reactive phosphorus (SRP) and light availability (Light) for each site, LAt (closed circle), LAT (closed triangle), lat (closed square), and laT (open triangle)...... 97 Figure 4.1. Mean organic matter content (a), algal biomass (b), biofilm thickness (c), biofilms Chl:C ratio (d), and primary productivity (e) as a function of light intensity. Mean ETRmax over time (f). Circle and square symbols represent aquaria with and without Chironomus dilutus larvae, respectively. In graph (e), closed symbols represent gross primary production and open symbols represent net ecosystem

x

production (NEP). Bars represent standard error of the mean. Regression statistics in Table 4.2...... 126 Figure 4.2. Larval biomass (a), specific growth rate (b), larval density (c), days to peak (d), number of adults emerged (e), and total adult biomass (f) as a function of benthic net ecosystem production. Error bars represent standard error of the mean. Regression statistics are in Table 4.2...... 128 Figure 4.3. Secondary production of Chironomus dilutus as a function of net ecosystem production (a). Production efficiency across the light gradient (b). Estimates of larval production for various mortality rate scenarios: Estimate = production estimate using calculated mortality rates in this paper; Estimate + and – 10% = estimated mortality rate ± 10%; Estimate varied = production estimated using mortality rates that varied over time based on survivorship curves from Jónasson (1979); Initial -0.99 = my calculated mortality rate with the mortality rate as -0.99 for day 0-1. Error bars represent standard error of the mean. Regression statistics in Table 4.2...... 129 Figure 5.1. Map of sampling locations and land use from the National Land Cover Database (2011). Land use categories: crop agriculture (brown), pasture (yellow), forest (green), developed (red), water (blue). Stream abbreviations: TC = Twin Creek, PR = Middle Branch Portage River, NC = Nimishillen Creek, EFWOC = East Fork White Oak Creek, BR = Blanchard River, LDC = Little Darby Creek, SBC = Scioto Brush Creek, PC = Paint Creek, LBC = Little Beaver Creek, EFLMR = East Fork Little Miami River, LMR = Little Miami River, VR = Vermilion River...... 171 Figure 5.2. The first two axes of the principal component analysis (PC1, PC2) synthesizing periphyton characteristics for the 12 streams sampled. Each point represents a study stream and are scaled by agricultural land use. Abbreviations: Chl = chlorophyll a in mg/m2, AFDM = ash-free dry mass in g/m2, N = percent nitrogen, P = percent phosphorus, NP = nitrogen: phosphorus ratio...... 172 Figure 5.3. Mean basal resource δ15N as a function of watershed agriculture (%, open symbols = periphyton, closed symbols = leaf), (b) mean leaf and periphyton δ15N signature for the 12 study streams. Bars represent standard error of the mean...... 173 Figure 5.4. Consumer δ15N as a function of primary producer (leaf and periphyton) δ15N (a-d). ANCOVA statistics in Table 5.7...... 174 Figure 5.5. The first two axes of the principal component analysis (PC1, PC2) synthesizing a subset of the periphyton total lipid profiles for the 12 streams sampled. Each point represents a study stream and are scaled by agricultural land use (%). Abbreviations for biomarkers: SAFA = saturated fatty acids, MUFA = monounsaturated fatty acids, HUFA = highly unsaturated fatty acids, PUFA =

xi

polyunsaturated fatty acids, EPA = eicosapentaenoic acid, EFA = essential fatty acids, w6 = omega-6 fatty acids, BAFA = bacterial fatty acids, and Fungi...... 175 Figure 5.6. Ratio of diatom fatty acid biomarkers (16:1ω7 and 20:5ω3) to green and cyanobacteria fatty acid biomarkers (18:2ω6 and 18:3ω3) as a function of agricultural land use. Bars represent standard error of the mean...... 176 Figure 5.7. First two axes of the non-metric multidimensional scaling (NMDS) ordination of fatty acid compositions of food-web components in stream ecosystems (two- dimensional stress = 0.11). Abbreviations: PUFA = polyunsaturated fatty acids, BAFA = bacterial fatty acids, MUFA = monounsaturated fatty acids, SAFA = saturated fatty acids, ω3 = ω3 fatty acids, ω6 = ω6 fatty acids, ω3_ω6, ω3:ω6 ratio...... 177 Figure 5.8. Periphyton-corrected NMDS scores as a function of periphyton δ15N...... 178

xii

LIST OF TABLES

Table 2.1. Land use and physical characteristics of the Grand, Maumee, and Portage river watersheds. Other = land use categories of Barren, Shrub, and Grassland...... 43 Table 2.2. Mean (± SE) water column nutrient concentrations for each stream during the three sampling periods. NO2-N + NO3-N are nitrogen, SRP is soluble reactive phosphorus, and TP is total phosphorus in µg/L. N:P is the molar ratio of nitrogen to phosphorus in the water column...... 44 Table 2.3. Mean periphyton nutrient stoichiometry, δ15N, and eco-enzyme activities for each stream during the three sampling periods: post-fertilizer application (May 2017), pre-harvest (September 2017), and post-harvest (November 2016). Ecoenzyme activities are normalized to nmol/g C/h. Values in parentheses are standard error of the mean. Abbreviations: N:P = molar N:P of periphyton, LAMP = leucine amino peptidase, PHOS = phosphatase, BG = β-glucosidase, BX = β-xylosidase. Missing values are denoted with --...... 45 Table 2.4. Spearman correlations (r) of stream water chemistry, and biofilm biomass, stoichiometry and eco-enzyme activities for post-fertilizer application (May 2017), pre-harvest (September 2017), and post-harvest (November 2016) sampling events. Correlations in bold are significant at p < 0.05. Abbreviations: Chl = biofilm chlorophyll-a, AFDM = ash free dry mass, N:P = molar N:P ratio in periphyton, N:Pw = molar N:P in water column, LAMP = leucine amino peptidase activity, PHOS = phosphatase activity, BG = β-glucosidase activity, BX = β-xylosidase activity...... 46 Table 2.5. N-, P-, or co-limitation of stream sites based on molar stoichiometry of water and sediments, and on the relative activities of phosphorus- to nitrogen-acquiring eco- enzymes (PHOS:LAMP) for each watershed during post-fertilizer application (May 2017), pre-harvest (September 2017), and post-harvest (November 2016) sampling events...... 47 Table 2.6. Discharge and water column metrics for main stem rivers averaged from the month prior to sampling upstream sites. Nitrate (NO3), nitrite (NO2), soluble reactive phosphorus (SRP), and total phosphorus (TP) are reported in µg/L. Total suspended solids is in mg/L and Discharge is in L/s. Data were downloaded from the National Center for Water Quality Research...... 48 Table 3.1. Dominant microbial community members and their relevant ecological functions and ...... 88

xiii

Table 3.2. Physical characteristics, sediment structure, and energy content of sediments at the stream sites sampled. Sediment particle size metrics include the coefficient of gradation (Cc) and the uniformity coefficient (Cu). Energy content is reported in joules/g organic matter. Values in parentheses are SE. Site abbreviations: LAt = Dolly Varden, laT = Selma Pike, lat = Garlough, LAT = Grinnell...... 89 Table 3.3. Diversity and evenness metrics of microbial communities analyzed from stream sediments...... 90 Table 4.1. Mean and standard errors (in parentheses) of measured variables: light intensity (μmol/m2/s), water temperature (°C), , sediment P (g/m2), sediment percent P (%), percent organic matter, and the density of eggs added (no./m2) of each treatment...... 123 Table 4.2. Simple linear regression statistics for light, production, and biomass parameters: light intensity (μmol/m2/s), chlorophyll-a (Chl-a, mg/m2), sediment organic matter (AFDM, g/m2), sediment P (g/m2), sediment percent P (%), gross primary production (GPP, mg C/m2/d), net primary production (NEP, mg C/m2/d), larval biomass density (mg/m2), and larval production (mg C/m2/d)...... 124 Table 5.1. Calculated watershed land use and mean measured water column characteristics for the study streams. Land use categories include agriculture (hay, pasture, and cultivated crops), developed land, forest, and (%). Total phosphorus (TP), soluble reactive phosphorus (SRP) and water column chlorophyll-a (Chl-a) are in µg/L. Total suspended solids (TSS) are in mg/L. Riffles refers to the number of riffles sampled in each stream. Values in parentheses represent standard errors...... 164 Table 5.2. Classification of organisms into functional feeding groups based on Merritt and Cummins (1996) and Cummins and Klug (1979). Organisms were used in stable isotope and total lipids analyses. a denotes functional feeding groups analyzed for stable-isotopes and stoichiometry only. Stream abbreviations as in Table 1...... 165 Table 5.3. Mean (SE) of periphyton and terrestrial metrics measured in 12 streams. Abbreviations: Chl-a = chlorophyll-a in mg/m2, AFDM = ash free dry mass in g/m2, Ash = ash content in g/m2, P = phosphorus content in g/m2, P% = percent phosphorus, N% = percent nitrogen, C% = percent carbon. Streams are ordered by increasing agricultural land use and abbreviations listed in Table 5.1. A “--“ indicates no data...... 166 Table 5.4. Top model regression statistics for water column total phosphorus (TP, µg/L), 15 15 leaf δ N and periphyton δ N. TP data were log10 -transformed. Agriculture and developed land use percentage data were arcsine square-root transformed...... 167

xiv

Table 5.5. Mean (SE) δ13C isotope ratios of consumer groups for each study stream. Stream are ordered by increasing agricultural land use and abbreviations are in Table 5.1...... 168 Table 5.6. Mean (SD) δ 15N isotope ratios of consumer groups for each study stream. Streams are ordered by increasing agricultural land use and abbreviations are in Table 5.1...... 169 Table 5.7. ANCOVA results of periphyton and leaf δ15N as predictors of consumer δ15N. Both models were significant at 0.05 level. Significant p-values in bold...... 170

xv

Acknowledgements

I thank my advisor, Yvonne Vadeboncoeur, for her commitment to my research and intellectual advancement. I am a better scientist for having had the opportunity to witness the way she approaches science and asks just the right questions. I cannot thank her enough for her patience, guidance, and encouragement throughout the Ph.D. process.

I thank my dissertation committee members, David Strayer, Lynn Hartzler, Katie

Hossler, and John Stireman, for their encouragement, reviews, and comments.

Many graduate and undergraduate students have come and gone from the

Vadeboncoeur lab during my time and all of them played a part in helping improve and complete this dissertation work, whether through help with field and lab work, or participating in lengthy discussions of science. Thanks to Leon Katona, Erica Hile,

Blythe Hazellief, Trey Godbolt, and numerous undergraduate students for help in the lab and in the field. They provided needed friendship, support, and understanding in the ways only lab-mates can.

Finally, I must thank my family, whose tireless support throughout this process has been a constant source of comfort. And to Kuyer, your unyielding love, encouragement, and friendship is a blessing and I could not have finished my dissertation without your support. You were always there to provide a boost of confidence when necessary and have been an enthusiastic sounding board for my ideas and frustrations.

Your passion for people, life, and learning remains a constant inspiration.

xvi

Chapter One

Introduction Streams drain the continents, forming channel networks that permeate the landscape. As such, stream ecosystems are physically, chemically, and biologically dynamic components of the landscape (Hynes 1974). Streams link watersheds to lakes and oceans by physically transporting , carbon, and sediments, but during transport, these terrestrial inputs are transformed by stream biota. Agricultural activities increase the amount of nutrients and sediments entering streams, modify stream channels, and remove or alter riparian vegetation. The interplay between these modifications has compounding effects on ecosystem processes and food web dynamics. I examined how landscapes, biofilms, and food webs interact in stream ecosystems that drain agricultural watersheds.

Streams are connected to their watersheds through surface water, groundwater and soil water flow paths (Findlay 1995, Miller et al. 2017). Inorganic and organic nutrients and carbon enter streams through these hydrological flow paths where they interact with the biofilms that colonize the surface of the streambed (Fisher et al. 1998).

Biofilms comprised of a community of , , fungi, and algae sequester, store, and recycle inorganic and organic matter in streams. As a result, the attached biofilm assemblages contribute to and determine biogeochemical fluxes within stream ecosystems (Pearl and Pinckney 1996).

1

Agricultural practices are diverse and involve multiple physical and chemical alterations to the landscape. Thus, the effects of land use on stream ecosystems can be complex (Allan 2004, Riseng et al. 2011). Agricultural practices increase the input of sediments and nutrients into streams, cause deterioration of riparian and stream channel habitat, and alter natural flow regimes (Johnson and Host 2010). These physical and chemical modifications of streams alter resource availability for biofilm and consumer communities, causing shifts in their composition, function, and trophic structure (Allan

2004, Diana et al. 2006, Riseng et al. 2011).

Removal of canopy cover allows high levels of light to reach the streambed and increases water temperature (Burrell et al. 2014). Streams without riparian trees have reduced retention capacity of non-point nutrient loads (Sweeney et al. 2004).

Furthermore, unnatural channel widening due to bank erosion and downcutting occurs due to the loss of the stabilizing effect of woody root systems when forested vegetation is removed (Faustini et al. 2009, Larson et al. 2018). These alterations interact and combine to compromise instream ecosystem function and processing of nutrients. Biofilms are key sites of biological activity, including ecosystem respiration, and primary production

(Battin et al. 2008, Griffiths et al. 2013). Additionally, attached algal carbon and terrestrial inputs of organic carbon form the basis of stream food webs (Rosemond et al.

2015).

The heterotrophic and autotrophic members of biofilms assimilate nutrients from runoff. The processes by which inorganic nutrients are sequestered from the water column and transformed, can mitigate problems associated with by reducing nutrient delivery to downstream ecosystems (Alexander et al. 2000). High rates

2

of biological and biogeochemical activity occurs on stream bottoms where dissolved nutrients are sequestered from the overlying water column (Battin et al. 2016). Greater uptake of inorganic nutrients results in slower downstream transport and increases the possibility of its removal directly via denitrification, in the case of N, or indirectly through trophic transfer to terrestrial in the case of P and N. I used three metrics to assess biofilm nutrient limitation in 3 watersheds with high influence of row-crop agriculture that contribute to the nutrient loading to Lake Erie (Chapter 2).

Nutrient-diffusing substrates (NDS) are extensively to assess biofilm nutrient limitation across regions a (Francouer 2001). Most NDS studies have used algal biomass measured as chlorophyll-a as the response metric. Fewer investigators have examined functional response metrics such as primary production and ecosystem respiration

(Marcarelli et al. 2009, Reisinger et al. 2016). Ecoenzyme activity of N- and P-acquiring enzymes produced by microbial biofilms provide an additional metric of the functional response of biofilms to nutrient availability. Ecoenzyme activities are correlated with the available nutrients within an ecosystem and the ratios of nutrient-acquiring enzymes provide an alternative assessment of biofilm nutrient limitation (Hill et al. 2012).

Biofilms mediate carbon turnover and ecosystem metabolism in streams. In agricultural watersheds, reduced litter input can limit the substrate available for heterotrophic respiration (Young and Huryn 1999), while increased nutrient loading from nonpoint sources can stimulate algal growth (Rosemond et al. 1993) providing an alternative substrate for biofilm bacterial . Identifying the factors that constrain biofilm diversity and function is critical to understanding the potential role of small streams in biogeochemical processing. I assessed how spatial variation in organic

3

matter quantity and quality affected stream sediment microbial community composition and function at a relatively small spatial scale (< 25 km) in a longitudinal stretch of a low order stream (Chapter 3).

The production, quantity, and composition of the attached algal assemblage are linked to nutrient and light availability (Hill et al. 2010). The production of primary consumers, the crucial link between basal energy and higher trophic levels, is constrained by the quality, nutrient stoichiometry, and production of attached algae (Lau et al. 2014,

Guo et al. 2016). Stable points to a strong reliance of consumers on attached algae. However, a direct, energetic dependence of consumer production on attached algal production has rarely been quantified. I conducted a laboratory experiment in which I manipulated light availability to control attached algal productivity on sediments to investigate the subsequent effect of attached algal productivity on the production of a dominant constituent of aquatic macroinvertebrate communities in lakes and streams, chironomid midges (Chapter 4).

Riparian vegetation removal associated with modern agriculture alters the relative availability of within-stream carbon (attached algae) and terrestrial carbon (riparian vegetation subsidies) to stream macroinvertebrate communities (Wiley et al. 1990,

O’Brian et al. 2017). Land-use-induced modifications to the stream channel can alter macroinvertebrate assemblage structure (Sponseller et al. 2008). Most stream-dwelling macroinvertebrates are primary consumers thereby serving as the channel through which the effects of basal carbon resources are propagated to and terrestrial landscapes

(Wallace and Webster 1996). I used natural biochemical tracers (stable isotopes, phospholipid fatty acids) to assess how stream food webs respond to variation in carbon

4

resources as a result of the amount of agricultural land use in the watershed and streamside riparian vegetation structure (Chapter 5).

Literature Cited Alexander, R. B., Smith, R. A., & Schwarz, G. E. (2000). Effect of stream channel size

on the delivery of nitrogen to the Gulf of Mexico. Nature, 403(6771), 758–761.

Allan, J. D. (2004). Landscapes and Riverscapes: The influence of land use on stream

ecosystems. Annual Review of , , and Systematics, 35(1), 257–

284.

Battin, T. J., Besemer, K., Bengtsson, M. M., Romani, A. M., & Packmann, A. I. (2016).

The ecology and of stream biofilms. Nature Reviews

Microbiology, 14(4), 251–263.

Battin, T. J., Kaplan, L. A., Findlay, S., Hopkinson, C. S., Marti, E., Packman, A. I.,

Newbold, J. D., & Sabater, F. (2008). Biophysical controls on organic carbon

fluxes in fluvial networks. Nature Geoscience, 1(2), 95–100.

Burrell, T. K., O’Brien, J. M., Graham, S. E., Simon, K. S., Harding, J. S., & McIntosh,

A. R. (2014). Riparian shading mitigates stream eutrophication in agricultural

catchments. Freshwater Science, 33(1), 73–84.

Diana, M., Allan, J. D., & Infante, D. (2006). The influence of physical habitat and land

use on stream fish assemblages in southeastern Michigan. In American Fisheries

Society Symposium, 48, 359-374.

Dodds, W. K., Bouska, W. W., Eitzmann, J. L., Pilger, T. J., Pitts, K. L., Riley, A. J.,

Schloesser, J. T., & Thornbrugh, D. J. (2009). Eutrophication of U.S. freshwaters:

5

analysis of potential economic damages. Environmental Science & Technology,

43(1), 12–19.

Faustini, J. M., Kaufmann, P. R., & Herlihy, A. T. (2009). Downstream variation in

bankfull width of wadeable streams across the conterminous United States.

Geomorphology, 108(3), 292–311.

Findlay, S. (1995). Importance of surface-subsurface exchange in stream ecosystems:

The . and Oceanography, 40(1), 159–164.

Fisher, S. G., Grimm, N. B., Martí, E., Holmes, R. M., & Jeremy B. Jones, J. (1998).

Material spiraling in stream corridors: a telescoping .

Ecosystems, 1(1), 19–34.

Francoeur, S. N. (2001). Meta-analysis of lotic nutrient amendment experiments:

detecting and quantifying subtle responses. Journal of the North American

Benthological Society, 20(3), 358–368.

Griffiths, N. A., Tank, J. L., Royer, T. V., Roley, S. S., Rosi-Marshall, E. J., Whiles, M.

R., Beaulieu, J. J., & Johnson, L. T. (2013). Agricultural land use alters the

seasonality and magnitude of stream metabolism. Limnology and Oceanography,

58(4), 1513–1529.

Guo, F., Kainz, M. J., Sheldon, F., & Bunn, S. E. (2016). Effects of light and nutrients on

periphyton and the fatty acid composition and somatic growth of invertebrate

grazers in subtropical streams. Oecologia, 181(2), 449–462.

Hill, B. H., Elonen, C. M., Seifert, L. R., May, A. A., & Tarquinio, E. (2012). Microbial

enzyme stoichiometry and nutrient limitation in US streams and rivers. Ecological

Indicators, 18, 540–551.

6

Hill, W. R., Smith, J. G., & Stewart, A. J. (2010). Light, nutrients, and growth

in oligotrophic streams. Ecology, 91(2), 518–527.

Hynes, H. B. N. (1975). The stream and its valley. Internationale Vereinigung für

theoretische und angewandte Limnologie: Verhandlungen, 19(1), 1-15.

Johnson, L. B., & Host, G. E. (2010). Recent developments in landscape approaches for

the study of aquatic ecosystems. Journal of the North American Benthological

Society, 29(1), 41–66.

Larson, D. M., Dodds, W. K., & Veach, A. M. (2018). Removal of woody riparian

vegetation substantially altered a stream ecosystem in an otherwise undisturbed

grassland watershed. Ecosystems, 1–13.

Lau, D. C. P., Sundh, I., Vrede, T., Pickova, J., & Goedkoop, W. (2014). Autochthonous

resources are the main driver of consumer production in dystrophic boreal lakes.

Ecology, 95(6), 1506–1519.

Marcarelli, A. M., Bechtold, H. A., Rugenski, A. T., & Inouye, R. S. (2009). Nutrient

limitation of biofilm biomass and metabolism in the Upper Snake River basin,

southeast Idaho, USA. Hydrobiologia, 620(1), 63–76.

Miller, M. P., Tesoriero, A. J., Hood, K., Terziotti, S., & Wolock, D. M. (2017).

Estimating discharge and nonpoint source nitrate loading to streams from three

end-member pathways using high-frequency water quality data. Water Resources

Research, 53(12), 10201–10216.

O’Brien, J. M., Warburton, H. J., Graham, S. E., Franklin, H. M., Febria, C. M.,

Hogsden, K. L., Harding, J. S., & McIntosh, A. R. (2017). Leaf litter additions

7

enhance stream metabolism, denitrification, and restoration prospects for

agricultural catchments. Ecosphere, 8(11).

Paerl, H. W., & Pinckney, J. L. (1996). A mini-review of microbial consortia: Their roles

in aquatic production and biogeochemical cycling. Microbial Ecology, 31(3),

225–247.

Riseng, C. M., Wiley, M. J., & Stevenson, R. J. (2004). Hydrologic and

nutrient effects on benthic community structure in midwestern US streams: a

covariance structure analysis. Journal of the North American Benthological

Society, 23(2), 309–326.

Rosemond, A. D., Benstead, J. P., Bumpers, P. M., Gulis, V., Kominoski, J. S., Manning,

D. W. P., Suberkropp, K., & Wallace, J. B. (2015). Experimental nutrient

additions accelerate terrestrial carbon loss from stream ecosystems. Science,

347(6226), 1142–1145.

Rosemond, A. D., Mulholland, P. J., & Elwood, J. W. (1993). Top-down and bottom-up

control of stream periphyton: effects of nutrients and . Ecology, 74(4),

1264–1280.

Sponseller, R. A., Benfield, E. F., & Valett, H. M. (2008). Relationships between land

use, spatial scale and stream macroinvertebrate communities. ,

46(10), 1409–1424.

Sweeney, B. W., Bott, T. L., Jackson, J. K., Kaplan, L. A., Newbold, J. D., Standley, L.

J., Hessian W. C., & Horwitz, R. J. (2004). Riparian deforestation, stream

narrowing, and loss of stream ecosystem services. Proceedings of the National

Academy of Sciences, 101(39), 14132–14137.

8

Wallace, J. B., & Webster, J. R. (1996). The role of macroinvertebrates in stream

ecosystem function. Annual Review of Entomology, 41(1), 115–139.

Wiley, M. J., Osborne, L. L., & Larimore, R. W. (1990). Longitudinal structure of an

agricultural prairie river system and its relationship to current stream ecosystem

theory. Canadian Journal of Fisheries and Aquatic Sciences, 47(2), 373–384.

Young, R. G., & Huryn, A. D. (1999). Effects of land use on stream metabolism and

organic matter turnover. Ecological Applications, 9(4), 1359–1376.

9

Chapter Two

SPATIOTEMPORAL VARIATION IN NUTRIENT LIMITATION OF EPILITHIC BIOFILMS IN LOW-ORDER STREAMS IN THE LAKE ERIE WATERSHED Introduction Biofilms on streambeds fix and mineralize carbon, and sequester and mineralize macronutrients (Romani et al. 2004, Battin et al. 2016). The autotrophic and heterotrophic component of benthic biofilms are linked through the exchange of labile sugars produced by algae and carbon dioxide respired by bacteria (Espeland et al. 2001,

Battin et al. 2003a). The spatial proximity of biofilm constituents facilitates these interactions largely through hydrolytic enzyme activity, the proximate agents of organic matter decomposition and nutrient acquisition (Sinsabaugh et al. 2009, Ylla et al. 2009).

The growth of the mixed autotrophic-heterotrophic communities can be limited by scouring events and light availability to the stream bottom (Francoeur and Jensen 1999,

Battin et al. 2003b, Ryder et al. 2006). Additionally, heterotrophic members can be limited by substrate (carbon, C) availability and quality. Both algae and heterotrophic bacteria within biofilms can be limited by nutrient availability (Francoeur and Jensen

1999, Rier and Stevenson 2002). Row crop agriculture and livestock farming increase nutrient loading to stream ecosystems in agricultural landscapes, ultimately affecting the structure and function of biofilms through accumulation of biomass and stimulation of primary and heterotrophic productivity (Carpenter et al. 1998, Hill et al. 2012).

10

The natural biogeochemical cycles of N and P have been altered by anthropogenic activities which have amplified the bioavailability of these nutrients by 100% and 400%, respectively, over the past hundred years (Galloway et al. 2004). Human activities related to food and energy production have increased the mobilization of these elements through mining of P and fixation of N2 through the Haber-Bosch process (Smil 2001, Cordell et al. 2009). Agricultural land use increases N and P concentrations in receiving bodies through runoff of excess fertilizer applied to the landscape and from inputs from waste generated by livestock farming (Howarth et al. 1996, Johnson et al. 2009). Application rates of N fertilizers have increased 30% since the 1960’s (Zhang et al. 2015) and over half of applied N is lost to the environment (Ciampitti and Vyn, 2014). Global P inputs have tripled since the 1960’s even though there has been considerable effort to manage and reduce P fertilizer application (Schindler et al. 2008, Lu and Tian 2017).

High nutrient loading of N and P can have synergistic effects, increasing biomass and productivity in most environments (Elser et al. 2007). High productivity and biomass of primary producers in aquatic ecosystems lead to eutrophication and degradation of water quality (Biggs et al. 2000). Elevated inputs of N and P have also been implicated in reductions in biological diversity (Smith et al.1999). Furthermore, excess nutrients stimulate the growth of nuisance algae, especially filamentous species that macroinvertebrate consumers find difficult to ingest (Miltner and Rankin 1998, Dodds

2007). Although excess nutrients stimulate attached algal growth, biofilms eventually become saturated, which reduces the efficiency of uptake of water column nutrients.

Excess nutrients travel downstream and contribute to eutrophication of receiving bodies

(Davis and Minshall 1999, Price and Carrick, 2016).

11

Not only does agricultural runoff increase the magnitude of nutrients entering receiving bodies, but also alters the ratios of those nutrients. Heterotrophic- and autotrophic-driven processes control the uptake and removal of nutrients in streams and they can be limited by nitrogen (N) and phosphorus (P). In pristine watersheds co- limitation of algae by N and P is common (Tank and Webster 1998, Francoeur et al.

1999). However, when enrichment by one element exceeds the other, it can quickly induce limitation by the alternative nutrient. Under low-nutrient conditions in the open ocean, algal cells have a molar N:P ratio of 16:1 (Redfield 1958, Falkowski 2000).

Departures from this ratio suggest algal metabolism is strongly limited by either P (N:P >

22) or N (N:P< 12 ) (Sterner and Elser, 2002). Strong N or P limitation of algal biofilms can affect ecosystem processes through reducing biomass accrual or algal productivity

(Vitousek et al. 1997, Elser et al. 2007). The large increase in bioavailable N relative to P through fertilizer production and application has caused an associated shift in N∶P supply ratios towards P limitation in agricultural streams (McDowell et al. 2009). This shift may intensify P limitation of benthic microbial communities (Liess et al. 2009).

Small streams transport water, particulate matter, and nutrients to larger streams and rivers. Despite their relatively small size, these streams play a disproportionately large role in nutrient uptake and transformation due to high benthic surface area to water volume and the relatively wide spatial extent, constituting up to 85% of total stream length within a drainage network (Alexander et al. 2000, Meyer et al. 2007). These streams collect most of the water, N, and P from adjacent terrestrial ecosystems. The high surface to volume ratios and shallow depths characteristic of low-order streams contribute to the rapid movement of nutrients from the water column into biofilm

12

biomass (Peterson et al. 2001). Thus, small streams have an immense capacity to sequester and transform nutrients ultimately regulating the downstream transport of nutrients (Peterson et al. 2001, Alexander et al. 2007). This capacity to store and transport nutrients is rarely studied in highly impacted agricultural streams. The nutrient retention capacity is likely controlled by algal and bacterial metabolism, which is regulated by nutrient availability, temperature, and light availability, disturbance, and grazing by consumers (Hall and Tank 2003).

Lake Erie has experienced dramatic changes in external nutrient loading. High nutrient inputs from agriculture cause seasonal harmful algal blooms and anoxic dead zones during summer in the western basin of the lake (Watson et al. 2016). The majority of nutrients that end up in the lake originate from the agricultural fields drained by low- order streams. P loading from tributary streams has been directly linked to the harmful algal blooms in Lake Erie (Kane et al. 2014) and resulted in efforts to regulate and reduce

P inputs (Schindler et al. 2008). This effort hinges on the assumption that P limitation occurs almost universally in freshwater ecosystems; however, there are periods during the year that Lake Erie is limited by N (Chaffin et al. 2011). Therefore, by observing changes in environmental parameters and N, P dynamics within the Lake Erie watershed, we can begin to understand how different factors combine to control nutrient processing within hydrological flow paths.

I investigated variation in dissolved and particulate N and P content in biofilms along low-order streams across three watersheds draining into Lake Erie. Phosphates from fertilizer are often sorbed to soil particles or incorporated into soil organic matter.

Therefore, P moves through the soil more slowly than N (Holten et al.1988). However,

13

the increase in implementation of subsurface tile drainage, as much as 40% since 1974, can result in the loss of as much as 50% of the soluble phosphorus from watersheds (King et al. 2015, USDA NASS 2016). The most common fertilizers applied in this region are inorganic blends for phosphorus and anhydrous ammonia or urea, and organic manure

(Pearl et al. 2016).

I employed eco-enzyme analysis of epilithic biofilms to assess spatiotemporal variation in nutrient limitation in relation to environmental drivers such as water column nutrient concentrations. I assessed temporal variation in nutrient limitation during periods relating to agricultural land use practices: post-fertilization, pre-harvest, and post-harvest of adjacent agricultural fields. I compared stream nutrient limitation based on eco- enzyme activity with biofilm stoichiometry and water column N:P ratios.

Methods Site Description I sampled 49 sites in 11 streams (n = 3-6 per stream) in 3 watersheds (n = 2-5 streams) within the Lake Erie drainage basin. To investigate temporal variation in biofilm nutrient limitation, I sampled during three time periods relevant to agricultural land use practices: post-fertilization (June 2-9th 2017), pre-harvest (September 4-10th 2017), and post-harvest (November 6-13th 2016).

The Maumee River watershed in northwestern Ohio has the largest drainage area of any Great Lakes river (~ 21500 km2, Table 2.1). The watershed is predominantly comprised of cultivated cropland. In the 1987 Great Lakes Water Quality Agreement, the

Maumee River was designated as an Area of Concern primarily due to sediment contamination and agricultural nutrient input (GLWQA 2012). The high concentrations

14

of phosphorus entering the river led to cultural eutrophication in Lake Erie (Kane et al.

2014).

The Portage River watershed is located directly adjacent to the Maumee River watershed and is considerably smaller, draining only ~1500 km2. The watershed is predominantly comprised of cultivated cropland (78%) and has a small area of wetlands along Lake Erie. Like the Maumee River watershed, the Portage River has excessive amounts of fine sediment and nutrient enrichment from cropland and animal farming.

Both the Maumee and Portage River watersheds have been highly modified from historic land uses for agriculture.

The Grand River watershed in northeastern Ohio is similar in size to the Portage

River watershed, draining ~ 1800 km2. The watershed is a mixture of forest, cultivated crops, pasture, and woody wetlands. The condition of biological communities in the upper Grand River basin is driven primarily by post-glacial physiography which has resulted in three classes of streams: lowland streams, upland headwaters, and the non- wadeable Grand River main stem. Many of the streams in this region are considered high quality and cold-water streams. Nevertheless, livestock and agriculture are influential land uses (33%), especially in the headwaters.

While at each site, I visually assessed the dominant land cover adjacent to the stream and riparian vegetation along each bank. Land cover was primarily row crop agriculture but a few sites were forest or animal pasture, particularly in the Grand River watershed.

Riparian vegetation was categorized based on woody vegetation presence: no woody vegetation, young woody vegetation (brush and immature plants), or mature woody vegetation (trees) for each side of the stream.

15

Physicochemical characteristics I measured dissolved oxygen, temperature (YSI ProODO, YSI, Yellow Springs,

Ohio), pH and conductivity (Oakton pH/Con 10 Series, Eutech Instruments, Singapore), near the edge of each stream. I collected water samples in triplicate for soluble reactive phosphorus (SRP), total phosphorus (TP), nitrate (NO3) and nitrite (NO2) analyses. I collected cobbles for periphyton samples from within the typical run habitat of each stream. Due to the high degree of channelization in these streams, run habitat was often the only habitat available. All samples were stored on ice and transferred to the laboratory for processing the same or next day.

To assess the relationship between low-order stream nutrient concentrations and concentrations in larger, downstream reaches of each river, I downloaded discharge data from US Geological Survey (USGS, https://waterdata.usgs.gov/nwis) gauging stations and nutrient concentrations from the National Center for Water Quality Research at

Heidelberg University (https://ncwqr.org/) for each of the main stems of the Portage and

Maumee Rivers. Data were not available for the Grand River watershed. I averaged SRP concentrations (µg/L) for the month prior to my sampling event to compare to sample day concentrations.

Nutrient Concentrations

Samples for nitrogen analyses (NO3 and NO2) were filtered immediately upon return to the lab with 0.2 µm nylon syringe filters (GE Millipore) into 20 mL screw-cap bottles (Wheaton) and stored frozen at −20 °C. Nitrogen samples were analyzed on a

Lachat 8500 nutrient analyzer according to manufacturer directions.

16

I filtered SRP samples using pre-ashed PALL AE glass fiber filters (1 µm nominal pore size) on the day samples were collected. TP and filtered SRP samples were frozen at -20°C until analyzed. I measured TP and SRP spectrophotometrically using the acid-molybdate method (UV-1700 Spectrophotometer, Shimadzu, Japan, Stainton et al.

1977).

Biofilm Biomass I collected cobbles (n = 5) in each run habitat within each site along a transect perpendicular to stream flow. In the laboratory, a periphyton sample was isolated by placing 2 plastic caps (diameter depended on size of cobble, 2.5 – 5.8 cm) on a flat section of each rock and scraping all uncovered algae off of the rock with a wire brush

(Flecker 1996). I then removed one of the caps, scraped and collected the periphyton from the defined area. I repeated this for the other cap. One sample was used for eco- enzyme analyses and the other for biomass and stoichiometric analyses. Eco-enzyme samples were frozen at -20°C until analyses. Biomass samples were frozen, freeze-dried, weighed, and ground into a fine powder (Hansson 1988). I subsampled periphyton for biomass estimates using the proxies chlorophyll-a and ash free dry mass (AFDM).

Additional subsamples were used for stable-isotope (SIA) and stoichiometric analyses (C,

N, and P).

I quantified chlorophyll-a spectrophotometrically (UV-1700 Spectrophotometer,

Shimadzu, Japan) and corrected for pheophytin with an acidification step (Arar and

Collins 1997). I quantified AFDM and biofilm sediment content as loss on ignition (LOI) at 500°C for 1 hour (APHA 2005).

17

Biofilm Stoichiometry Subsamples of periphyton were acidified prior to stable isotope, %C, %N, and %P analysis by adding single drops of 1N-HCl until the sample stopped emitting CO2 gas.

Periphyton samples were combined for each riffle. Stable carbon (δ13C) and nitrogen

(δ15N) isotopes, %C and %N were analyzed using an isotope-ratio mass spectrometer

(Finnigan MAT Delta Plus) coupled to an elemental analyzer (Carlo Erba NC2500) at the

Cornell University Stable Isotope Laboratory. Replicates were run on 8% of the samples to determine analytical error. I measured periphyton phosphorus content spectrophotometrically after burning 1-3 mg at 500°C for 1 hour and using the acid- molybdate method (UV-1700 Spectrophotometer, Shimadzu, Japan) (Stainton et al.

1977).

Eco-enzyme activities I analyzed eco-enzyme activity of carbon- and nutrient-acquiring enzymes in lotic ecosystems. These enzymes include: phosphatase (PHOS), and leucine aminopeptidase

(LAMP). The ratio of PHOS∶LAMP enzyme activities shift in response to metabolic limitation by the availability of inorganic P or N (Sala et al. 2001). To assess the eco- enzyme activity of carbon-acquiring enzymes, I analyzed the activity of β-glucosidase

(BG) and β-xylosidase (BX). The ratios of BG:BX characterize organic C processing in biofilms and indicate the contribution of autochthonous:allochthonous sources to C processing dynamics (Romani and Sabater 2000).

I diluted thawed periphyton samples to ~ 35 mL with filtered stream water from the sample site. I homogenized each sample using a mortar and pestle for 1 minute or until samples appeared uniform. I quantified eco-enzymes fluorometrically in 96-well,

18

black polystyrene microplates with 300-µL wells using methylumbelliferone (MUB)- linked (for PHOS, BG, and BX) and leucine 7-amido-4-methylcoumarin (for LAMP) model substrates. The emission wavelength was 455nm and the excitation wavelength was 365nm (Smucker et al. 2009). I normalized ecoenzyme activity to substrate accumulated per unit biofilm over time for sample organic C content (nmol/g C/h).

Statistical Analysis All statistical analyses were performed using R software (Version 3.3.2, R Core

Team, 2013). Prior to statistical analysis, data were checked for normality using the

Shapiro–Wilk normality test. Replicates were averaged at the site level. To explore potential environmental drivers of nutrient limitation, biofilm biomass and biofilm stoichiometry, I use general linear mixed effect models with a nested design. I ran models using the random effects of time, site within stream, and stream within watershed.

Stepwise models were run using the nlme package in R. The best-fitting model was selected based on the minimum Akaike’s information criteria (AIC; Akaike, 1974). To normalize data for parametric analysis, all non-normally distributed variables were log(x

+ 1)-transformed before running the model. In models with nutrient concentrations as predictor variables, site within a stream was usually not a significant random effect.

Therefore, site was considered a replicate within a stream for further analyses. I evaluated the relationships among biofilm biomass, stoichiometry, nutrient limitation and environmental variables using Spearman rank (r) correlation to avoid problems associated with non-normal data distribution. Analysis of variance (ANOVA) was used to assess variation in biofilm and water quality metrics across time and among watersheds. I used canonical correspondence analysis to assess potential environmental drivers of biofilm

19

structural components (AFDM, Chl-a, %AFDM, P content, and sediment content) and biofilm functional metrics (PHOS, LAMP, BG, BX activities, and PHOS:LAMP, BG:BX ratios) on untransformed data.

Nutrient limitation I used three approaches to assess relative N or P limitation. I used water column N and P concentrations and the biofilm ratio of N:P < 12 and > 22 for N and P limitation, respectively (Redfield 1958, Sterner and Elser 2002). For eco-enzyme activities of P- to

N- acquiring enzyme activities, I used deviations from the > 1:4 (P-limited) and < 1:10

(N-limited) threshold ratios of phosphatase to peptidase (Hill et al. 2006, Sinsabaugh et al. 2009). I classified streams as N or P limited based on their deviation from these expected N:P ratios. Values within these thresholds were classified as co-limited.

Results Physicochemical Characteristics In general, sites within a stream did not significantly vary in terms of water chemistry assessed using general linear mixed effects models. The exception to this pattern occurred for Blue Creek within the Maumee River watershed which has a concentrated animal feeding operation (CAFO). Downstream of the CAFO, Blue Creek

NO2-N and NO3-N concentrations were substantially elevated.

Across watersheds and time periods, streams encompassed a wide range of water nutrient concentrations that spanned an order of magnitude within each variable (Table

2.2). NO3-N + NO2-N concentrations ranged between 9.1 and 21100 µg N/L, with significantly higher concentrations post-harvest than post-fertilizer application and pre- harvest (ANOVA, F2,210 = 12.24, p < 0.001). Additionally, NO3-N + NO2-N concentrations were significantly lower in the Grand River watershed compared to 20

Maumee and Portage River watersheds (ANOVA, F2,210 = 20.55, p < 0.001). In contrast,

SRP concentrations did not significantly vary across time (ANOVA, F2,263 = 0.9097, p >

0.05). However, SRP was significantly lower in the Grand River watershed than in the

Portage and Maumee River watershed (ANOVA, F2,263 = 7.164, p < 0.01) and ranged between 3.3 and 161 µg/L. NO3-N concentrations were 1-2 orders of magnitude greater than SRP concentrations post-fertilizer application and post-harvest. TP concentrations ranged between 5.1 and 720 µg/L, with significantly higher concentrations post-fertilizer application than pre- and post-harvest (ANOVA, F2,263 = 54.97, p < 0.001). TP did not vary significantly across watersheds (ANOVA, F2,263 = 2.123, p > 0.05).

Biofilm composition Attached algal biomass, measured as chlorophyll-a, ranged from 11.9 to 180 mg/m2 (Figure 2.2a) and was significantly higher post-harvest (November) and post- fertilizer (June) than pre-harvest (September) (ANOVA, F2,151 = 19.26, p < 0.001). Algal biomass was significantly lower in the Grand River watershed than in the Maumee and

Portage River watersheds (ANOVA, F2,151 = 6.632, p < 0.01). Total organic matter content (AFDM) ranged between 2.3 and 149 g/m2 across streams (Figure 2.2b). AFDM was significantly lower pre-harvest than post-fertilizer and post-harvest (ANOVA, F2,151

= 28.41, p < 0.001), and was significantly higher in the Portage River watershed than the

Grand and Maumee river watersheds (ANOVA, F2,151 = 7.744, p < 0.001). Water column

15 nutrient concentrations (NO3-N, TP, SRP concentrations), periphyton δ N, and riparian vegetation, did not predict biofilm algal biomass or organic matter content (generalized linear mixed effect model, p > 0.05). In contrast, biofilm sediment content and algal biomass were positively correlated across watersheds and sampling periods (Figure 2.3).

21

Periphyton Stoichiometry and Eco-enzyme Activity The molar ratio of biofilm N:P ranged from 26.1 to 57.8 among streams (Table

2.3). Biofilm biomass as chlorophyll-a and AFDM, and water column nutrient concentrations were not correlated with biofilm nutrient ratios (Table 2.4).

Biomass-specific hydrolytic enzyme activity for carbon-acquiring enzymes were often an order of magnitude greater for BG than for BX (Table 2.3). The activity of nutrient-acquiring enzymes was greatest during pre-harvest. Carbon-acquiring enzyme activity was greater pre-harvest compared to post- harvest and fertilizer application. BG activity was positively correlated with AFDM (g/m2) for all sampling periods (Table 2.4).

PHOS activity was positively correlated with algal biomass (chlorophyll-a) post-fertilizer application and negatively with AFDM post-harvest. BG activity was positively correlated with LAMP activity while BG:BX was negatively correlated with SRP and

AFDM pre-harvest. Post-harvest, LAMP and PHOS activities positively correlated with water column SRP. Ratios of BG:BX were weakly positively correlated with AFDM post-harvest and negatively correlated with AFDM pre-harvest and post-fertilizer.

PHOS:LAMP ratios were correlated with NO3 and water column N:P post-harvest. In general, biofilm eco-enzyme activity was correlated with water chemistry rather than with biofilm nutrient content, although we have biofilm N and P data only for the post- harvest sample.

Drivers of Biofilm Biomass and Enzyme Activities The biofilm biomass metrics (AFDM and chlorophyll-a) and their relationships with water chemistry (NO3, NO2, SRP and TP) were summarized using a CCA biplot

(Figure 2.4). Samples from each time period (post-fertilizer, pre-harvest, and post-

22

harvest) assembled closely together regardless of watershed or stream of origin.

ANOVA-like permutation tests for the joint effect of each constraining variable for the

CCA of biofilm structural attributes (AFDM, chlorophyll-a, sediment content, P content, and percent organic matter) were significant across both axes (1000 permutations, p <

0.001). Axis 1 explained 60% of the spread across stream sites while axis 2 explained

20% of the variation. Samples separated along CCA1 axis by P (Spearman’s r = -0.43) and N (r = 0.33) concentrations in water. Higher algal biomass was associated with higher NO3 concentrations. Samples collected during post- harvest loaded closely to NO3, while pre-harvest samples loaded closely to SRP. In general, post-fertilizer samples assembled between pre-harvest and post-harvest samples.

The activities of extracellular enzymes and their relationships with the biofilm biomass metrics and water chemistry were summarized using a CCA biplot (Figure 2.5).

ANOVA-like permutation tests for the joint effect of each constraining variable for the

CCA of biofilm enzyme activities (PHOS, LAMP, BG, BX, PHOS:LAMP, and BG:BX) were significant across both axes (1000 permutations, p < 0.001). CCA axis 1 explained

70% of the variance and was correlated with biofilm biomass measured as chlorophyll-a

(Spearman’s r = 0.22). CCA axis 2 explained only 18% of the variance and was correlated with NO3 (Spearman’s r = -0.30) and total organic matter content (AFDM, r =

0.11). Samples from post-harvest were the most variable along CCA2.

Nutrient limitation I compared molar stoichiometry of N and P in water and in periphyton with biofilm ratios of LAMP:PHOS to assess how each metric of limitation compares with the other (Table 2.5). In general, metrics of nutrient limitation did not indicate consistent

23

limitation status across stream sites or time periods. Post-fertilizer application, the majority of stream sites were P-limited according to water column N:P. The exceptions to this trend were three sites within Baughman Creek and one site within Phelps Creek of the Grand River watershed. Eco-enzyme ratios of PHOS:LAMP also indicated that most streams were P limited post-fertilizer application although several sites had PHOS:LAMP ratios below 0.25 but still higher than the threshold of N-limitation, 0.10. Most sites in the Portage and Grand River watersheds exhibited P limitation pre-harvest while sites within the Maumee river watershed were N limited pre-harvest based on water N:P. In contrast, PHOS:LAMP ratios indicated P limitation across most sites pre-harvest.

Approximately half the stream sites exhibited P limitation while the other half indicated

N limitation for both water column N:P and PHOS:LAMP ratios across watersheds post- harvest. Periphyton N:P ratios overwhelming indicated that streams were P limited post- harvest with only 2 stream sites showing co-limitation.

Longitudinal changes in nutrient concentrations To evaluate how nutrient concentrations at our sites compared to main stem river concentrations, we compared downloaded data on the average SRP concentrations for the month prior to our sampling date. In general, water column phosphorus (SRP) concentrations were lower in the headwater streams that we sampled than in the larger main-stem river (Figure 2.6). In the mainstems, SRP was highest during post-fertilizer and lowest during pre-harvest time periods. Post-fertilizer application all streams had lower SRP concentrations in the upstream branches compared to the main stem.

24

Discussion Although there was not a strong relationship between algal biomass and water column nutrient concentrations, algal biomass (as chlorophyll) and total organic matter biomass (AFDM) were higher during time periods with higher nutrient concentrations.

Periphyton biomass was lowest in the watershed with the least agricultural land use.

However, the relationship between periphyton biomass and enzyme activity with nutrient concentrations within watershed and across time indicated complex relationships among

N and P availability, biomass accumulation, and enzyme activity. The extent of nutrient limitation, primarily by P, varied seasonally and across watersheds, and nutrient limitation status was not always consistent among the three metrics we used.

Nevertheless, we found that increasing nitrogen concentrations was associated with increased P limitation of stream biofilms.

Although water column nutrient concentration was similar to values observed in other Midwest streams (Miltner 2010), the Lake Erie tributaries have much higher inorganic N concentrations relative to P across seasons and streams. These consistently high water column N:P ratios suggested that periphyton in these streams are P limited

(Table 2.2). The seasonal results are consistent with previous studies that have observed either P or no limitation in agriculturally dominated watersheds in the Midwest (Johnson et al. 2009, Hill et al. 2012). Inorganic N fertilizer application has been steadily increasing (Pearl et al. 2016) resulting in increased bioavailability of N in streams and receiving bodies (Dodds et al. 2016). In contrast, there has been effort to reduce P fertilizer application which has resulted in improved water quality across many systems yet there is evidence suggesting that dual N and P reductions are necessary to stem

25

development of harmful algal blooms (Pearl et al. 2016). The rather high N:P ratios observed, especially post-fertilizer application, suggests that these low-order streams are serving as a potential conduit of N. Given the sensitivity of downstream ecosystems to N loading from watersheds, dual reductions of N and P are necessary in the Lake Erie watershed.

During both fall and spring, algal biomass and total organic matter content were considerably greater than in mid-summer (Figure 2.2). Land use influences both nutrient and light availability, which can have variable effects on accumulation of biomass in low- order stream biofilms (Johnson et al. 2009). Agricultural land use often increases light availability in streams by reducing riparian canopy cover (Fierro et al. 2017). With the exception of the Grand River watershed, our streams did not have riparian vegetation that extended over the stream channel. The canopy could have induced light limitation and contributed to the lower algal biomass observed in the Grand River watershed.

Periphyton biomass metrics were poorly predicted by water column nutrients. In general, water column N and P concentrations did not correlate with biomass across watersheds or time periods. Biomass accrual of attached biofilms is often not directly correlated with water column nutrient concentrations (Dodds 2007, Miltner 2010). When examined alone, dissolved N and P are correlated with attached algal biomass at low concentrations but those correlations become less relevant at high N or P concentration like the values observed in this study (Prairie et al.1989, Dodds and Smith 2016). Often,

N and P together describe benthic algal biomass better than when either nutrient is assessed alone (Lewis and McCutchan 2010). We observed drastic differences in N

26

relative to P supply shifting the balance toward P limitation which might have affected algal growth (Smith 1982).

Factors other than nutrients might control periphyton and attached algal biomass.

Flooding and associated scourings can limit biofilm biomass (Power and Stewart 1987,

Riseng et al. 2004). The previous scouring event occurred ~ 10 days prior to sampling post-fertilizer application which might explain why the observed biomass during this time period was lower than post-harvest even though water column nutrient concentrations were higher. Both other sampling events occurred > 30 days after the previous peak flow conditions so biomass accrual was not affected by scour.

One of the most robust relationships observed over space and time was the strong positive correlation between attached algal biomass and biofilm sediment content (Figure

2.3). Removal of riparian vegetation increases sediment delivery to streams in impacted systems (Larson et al. 2018). In streams, sediment transport is highest during storms and associated scouring events (Gao and Josefson, 2012). The deposition of sediments occurs in the receding phase of floods. The positive correlation we observed between epilithic algal biomass and sediment content might be a reflection of the time since the last scouring event. The effects of sediments on periphyton are complex and include reducing light availability through smothering and reducing hard substrata for colonization

(Brookes 1986, Wood and Armitage, 1997). Accumulation of sediments can initially reduce biofilm and biomass. However, this response is temporary and biomass can overcome sediment deposition over relatively short timescales (Izagirre et al.

2009). Alternatively, the attached algal community might contribute to sediment entrapment. High sediment loads can alter the periphyton community composition

27

through favoring filamentous algal growth (Izagirre et al. 2009). Long filamentous streamers, which were abundant in our streams, might trap sediments leading to the observed positive correlation between attached algal biomass and periphyton sediment content.

Agriculture is a highly variable practice across the landscape and changes in till practices, crop rotation, fertilizer application (both in terms of rates of application and broadcast versus incorporated application), and animal versus crop farming can impact the transfer of nutrients and sediments from farm fields to streams (Bosch et al. 2014,

Pittelkow et al. 2015, Garcia et al. 2016). In this study, crop cover was composed of both soybean and corn within and across watersheds making it difficult to parse out the effects of these crop types on stream function. However, one stream was directly impacted by runoff from a CAFO, which was reflected in elevated water column nutrient concentrations downstream. No-till agriculture is a management system that aims to reduce soil erosion, decrease fertilizer input costs, and sustain long-term crop productivity (Powlson et al. 2014). Tillage generally occurs in spring time just prior to planting (Ohio Farm Bureau, OFBF.org) which may affect phosphorus and sediment loading to streams during this relatively high precipitation period. Tillage and harvesting disurb the soil on the farm field which might contribute to the high periphyton sediment content during these time periods relative to pre-harvest.

Our estimates of eco-enzyme activity, and their relative activities among the various enzymes, are similar to values reported from other studies of aquatic ecosystems

(Burns and Ryder 2001, Sinsabaugh and Foreman 2001, Hill et al. 2010, Hill et al. 2012).

Eco-enzyme activity related to C acquisition (BG:BX, and BG activity) was not a simple

28

function of periphyton biomass (AFDM, Table 2.4). However, BG activity was orders of magnitude greater than BX activity indicating that labile C, likely of algal origin, is a major carbon resource for biofilm . Labile carbon uptake by biofilm bacteria increases in response to thick algal mats where labile dissolved organic carbon is plentiful (Sobczak 1996).

Biofilm phosphatase activity was not correlated with available P, a finding reported by other researchers (Clinton et al., 2010; Hill et al., 2010), although phosphatase activity is often negatively correlated with water and sediment P concentrations (Findlay et al. 2001, Sinsabaugh et al. 2009). In contrast to results of other studies, eco-enzyme activity involved in N-acquisition was not correlated with available

N (Clinton et al. 2010). This is surprising given the number of studies where dissolved inorganic nutrient concentrations correlate with nutrient-acquiring enzyme activities

(Foreman et al. 1998, Findlay et al. 2001, Sinsabaugh et al. 2009, Hill et al. 2010, Lang et al. 2012). It has been suggested that examining the ratios of enzymes associated with N, and P acquisition provides a more comprehensive view of response of attached communities to the nutrient availability in streams (Hill et al. 2012) because both N and P can stimulate decomposition and retention rates of biofilms (Rosemond et al. 2015).

Two main patterns emerged from comparison of metrics for estimating nutrient limitation. First, over time and across watersheds, P limitation was more widespread than

N limitation. This supports the widely held notion that freshwater ecosystems are P limited (Downing and McCauley 1992). Studies using NDS in agriculturally impacted streams have also shown P limitation and generally fail to show N limitation across seasons (Hill et al. 2012, Johnson et al. 2009, Reisinger et al. 2016). These studies used

29

biomass accrual as estimates of limitation and did not find seasonal variation in limitation status. In contrast, PHOS:LAMP ratios of attached biofilms and water column N:P demonstrates that biofilms are more strongly P limited post-fertilizer application. The evidence from dissolved nutrients, periphyton stoichiometry and microbial enzyme activity suggests that these low order streams are P limited due to high nitrogen concentrations particularly during early in the growing season (post-fertilizer and pre- harvest). Second, when water N:P indicated N limitation during low flow periods (pre- harvest), it was not reflected in ecoenzyme activities. As stated previously, dissolved nutrient ratios are indicators of the available pool of N relative to P, not necessarily limitation experienced by the biofilm community. In contrast, ecoenzyme activity for nutrient acquisition reflects a physiological demand and changes rapidly in response to inorganic nutrient availability (Burns et al. 2013). In essence, the stoichiometry of eco- enzyme activity provides a biological perspective on the influence of the imbalance of nutrients being transported from each watershed.

The relative consistency with which each metric showed P limitation suggest that algae in these low-order streams are responding to landscape level processes. Over time, the degree of P limitation varied and was most extreme post-fertilizer application when water column NO3 concentrations were highest. The high NO3 and SRP concentrations observed causes these streams to be considered nutrient replete, but ecoenzyme activities indicate that the magnitude of nutrient availability is not the sole factor. Rather, the imbalance of N relative to P in these streams is driving the patterns of limitation

(Reisinger et al. 2016).

30

The high PHOS:LAMP ratios observed in our study indicates that P is actively sequestered by biofilm communities which potentially slows the transport of P to downstream reaches of these streams. This is particularly interesting considering that high riverine SRP concentrations have been linked to blooms in Lake Erie

(Kane et al. 2014). Small streams are shallow, and in agricultural landscapes lack riparian trees. Thus, they experience high light availability relative to the murky and deep main stems of rivers (Johnson et al. 2009). Decreased light availability to the river bottom in large rivers could reduce the potential for biofilms to sequester nutrients from the water column. Alternatively, higher SRP concentrations could result from the larger watershed area contributing to higher P loading (Clement and Steinman 2017) or the remineralization of particulate P from upstream. Nevertheless, phosphorus is being sequestered by algae upstream where the greater surface area and periphyton biomass facilitates high rates of uptake.

Literature Cited Alexander, R. B., Smith, R. A., & Schwarz, G. E. (2000). Effect of stream channel size

on the delivery of nitrogen to the Gulf of Mexico. Nature, 403(6771), 758–761.

Battin, T. J., Besemer, K., Bengtsson, M. M., Romani, A. M., & Packmann, A. I. (2016).

The ecology and biogeochemistry of stream biofilms. Nature Reviews

Microbiology, 14(4), 251–263.

Battin, T. J., Kaplan, L. A., Newbold, J. D., Cheng, X., & Hansen, C. (2003). Effects of

current velocity on the nascent architecture of stream microbial biofilms. Applied

and Environmental Microbiology, 69(9), 5443–5452.

31

Biggs, B. J. F. (2000). Eutrophication of streams and rivers: dissolved nutrient-

chlorophyll relationships for benthic algae. Journal of the North American

Benthological Society, 19(1), 17–31.

Bosch, N. S., Evans, M. A., Scavia, D., & Allan, J. D. (2014). Interacting effects of

and agricultural BMPs on nutrient runoff entering Lake Erie.

Journal of Great Lakes Research, 40(3), 581-589.

Brookes, A. (1986). Response of aquatic vegetation to sedimentation downstream from

river channelisation works in England and Wales. Biological Conservation, 38(4),

351–367.

Burns, A., & Ryder, D. S. (2001). Response of bacterial extracellular enzymes to

inundation of floodplain sediments. Freshwater Biology, 46(10), 1299–1307.

Burns, R. G., DeForest, J. L., Marxsen, J., Sinsabaugh, R. L., Stromberger, M. E.,

Wallenstein, M. D., Weintraub, M. N. & Zoppini, A. (2013). Soil enzymes in a

changing environment: Current knowledge and future directions. Soil Biology and

Biochemistry, 58, 216–234.

Carpenter, S. R., Caraco, N. F., Correll, D. L., Howarth, R. W., Sharpley, A. N., & Smith,

V. H. (1998). Nonpoint pollution of surface waters with phosphorus and nitrogen.

Ecological Applications, 8(3), 559–568.

Chaffin, J. D., Bridgeman, T. B., Heckathorn, S. A., & Mishra, S. (2011). Assessment of

Microcystis growth rate potential and nutrient status across a trophic gradient in

western Lake Erie. Journal of Great Lakes Research, 37(1), 92–100.

Ciampitti, I. A., & Vyn, T. J. (2014). Understanding global and historical nutrient use

efficiencies for closing maize yield gaps. Agronomy Journal, 106(6), 2107–2117.

32

Clement, D. R., & Steinman, A. D. (2017). Phosphorus loading and ecological impacts

from agricultural tile drains in a west Michigan watershed. Journal of Great

Lakes Research, 43(1), 50–58.

Clinton, S. M., Edwards, R. T., & Findlay, S. E. G. (2010). Exoenzyme activities as

indicators of dissolved organic matter composition in the hyporheic zone of a

floodplain river. Freshwater Biology, 55(8), 1603–1615.

Cordell, D., Drangert, J.-O., & White, S. (2009). The story of phosphorus: Global food

security and food for thought. Global Environmental Change, 19(2), 292–305.

Davis, J. C., & Minshall, G. W. (1999). Nitrogen and phosphorus uptake in two Idaho

(USA) headwater wilderness streams. Oecologia, 119(2), 247–255.

Dodds, W. K. (2007). Trophic state, eutrophication and nutrient criteria in streams.

Trends in Ecology & Evolution, 22(12), 669–676.

Dodds, W., & Smith, V. H. (2016). Nitrogen, phosphorus, and eutrophication in streams.

Inland Waters, 6(2), 155–164.

Downing, J. A., & McCauley, E. (1992). The nitrogen : phosphorus relationship in lakes.

Limnology and Oceanography, 37(5), 936–945.

Elser, J. J., Bracken, M. E. S., Cleland, E. E., Gruner, D. S., Harpole, W. S., Hillebrand,

H., Ngai, J. T., Seabloom, E. W., Shurin, J. B., & Smith, J. E. (2007). Global

analysis of nitrogen and phosphorus limitation of primary producers in

freshwater, marine and terrestrial ecosystems. Ecology Letters, 10(12), 1135–

1142.

33

Espeland, E. M., Francoeur, S. N., & Wetzel, R. G. (2001). Influence of algal

photosynthesis on biofilm bacterial production and associated glucosidase and

xylosidase activities. Microbial Ecology, 42(4), 524–530.

Falkowski, P. G. (2000). Rationalizing elemental ratios in unicellular algae. Journal of

Phycology, 36(1), 3–6.

Fierro, P., Bertrán, C., Tapia, J., Hauenstein, E., Peña-Cortés, F., Vergara, C., Cerna, C.,

& Vargas-Chacoff, L. (2017). Effects of local land-use on riparian vegetation,

water quality, and the functional organization of macroinvertebrate assemblages.

Science of The Total Environment, 609, 724–734.

Findlay, S., Strayer, D., Goumbala, C., & Gould, K. (1993). Metabolism of streamwater

dissolved organic carbon in the shallow hyporheic zone. Limnology and

Oceanography, 38(7), 1493–1499.

Flecker, A. S. (1996). Ecosystem engineering by a dominant in a diverse

tropical stream. Ecology, 77(6), 1845–1854.

Francoeur, S. N., & Jensen, S. I. (1999). Substratum-associated . Water

Environment Research, 71(5), 1078-1092.

Francoeur, S. N., Smith, R. A., & Lowe, R. L. (1999). Nutrient limitation of algal

biomass accrual in streams: seasonal patterns and a comparison of methods.

Journal of the North American Benthological Society, 18(2), 242–260.

Galloway, J. N., Dentener, F. J., Capone, D. G., Boyer, E. W., Howarth, R. W.,

Seitzinger, S. P., Asner, G. P., Cleveland, C. C., Green, P.A., Holland, E. A.,

Karl, D. M., Micheals, A. F., Porter, J. H., Townsend, A. R., & Vörösmarty, C. J.

34

(2004). Nitrogen cycles: past, present, and future. Biogeochemistry, 70(2), 153–

226.

Gao, P., & Josefson, M. (2012). Temporal variations of suspended sediment transport in

Oneida Creek watershed, central New York. Journal of Hydrology, 426–427, 17–

27.

García, A. M., Alexander, R. B., Arnold, J. G., Norfleet, L., White, M. J., Robertson, D.

M., & Schwarz, G. (2016). Regional effects of agricultural conservation practices

on nutrient transport in the Upper Mississippi River Basin. Environmental Science

& Technology, 50(13), 6991-7000.

GLWQA (Great Lakes Water Quality Agreement). 2012. The 2012 Great Lakes Water

Quality Agreement. http://tinyurl.com/gt92hrh.

Hall, R. O., & Tank, J. L. (2003). Ecosystem metabolism controls nitrogen uptake in

streams in Grand Teton National Park, Wyoming. Limnology and Oceanography,

48(3), 1120–1128.

Hansson, L.-A. (1988). Chlorophyll a determination of periphyton on sediments:

identification of problems and recommendation of method. Freshwater Biology,

20(3), 347–352.

Hill, B. H., Elonen, C. M., Jicha, T. M., Cotter, A. M., Trebitz, A. S., & Danz, N. P.

(2006). Sediment microbial enzyme activity as an indicator of nutrient limitation

in Great Lakes coastal wetlands. Freshwater Biology, 51(9), 1670–1683.

Hill, B. H., Elonen, C. M., Seifert, L. R., May, A. A., & Tarquinio, E. (2012). Microbial

enzyme stoichiometry and nutrient limitation in US streams and rivers. Ecological

Indicators, 18, 540–551.

35

Holten, H., Kamp-Nielsen, L., & Stuanes, A. O. (1988). Phosphorus in soil, sediment and

water: An overview. Hydrobiologia, 170, 19-34.

Howarth, R. W., Billen, G., Swaney, D., Townsend, A., Jaworski, N., Lajtha, K., …

Zhao-Liang, Z. (1996). Regional nitrogen budgets and riverine N & P fluxes for

the drainages to the North Atlantic Ocean: Natural and human influences. In

Nitrogen Cycling in the North Atlantic Ocean and its Watersheds (pp. 75–139).

Springer, Dordrecht.

Izagirre, O., Serra, A., Guasch, H., & Elosegi, A. (2009). Effects of sediment deposition

on periphytic biomass, photosynthetic activity and algal community structure.

Science of The Total Environment, 407(21), 5694–5700.

Johnson, L. T., Tank, J. L., & Dodds, W. K. (2009). The influence of land use on stream

biofilm nutrient limitation across eight North American . Canadian

Journal of Fisheries and Aquatic Sciences, 66(7), 1081–1094.

Kane, D. D., Conroy, J. D., Peter, R., Baker, D. B., & Culver, D. A. (2014). Re-

eutrophication of Lake Erie: Correlations between tributary nutrient loads and

phytoplankton biomass. Journal of Great Lakes Research, 40(3), 496–501.

King, K.W., M.R. Williams, and N.R. Fausey. (2015a). Contributions of systematic tile

drainage to watershed scale phosphorus transport. Journal of Environmental

Quality, 44, 486-494.

Lang, D. A., King, R. S., & Scott, J. T. (2012). Divergent responses of biomass and

enzyme activities suggest differential nutrient limitation in stream periphyton.

Freshwater Science, 31(4), 1096–1104.

36

Larson, D. M., Dodds, W. K., & Veach, A. M. (2018). Removal of woody riparian

vegetation substantially altered a stream ecosystem in an otherwise undisturbed

grassland watershed. Ecosystems, 1–13.

Lewis, Jr, W. M., & McCutchan, Jr, J. H. (2010). Ecological responses to nutrients in

streams and rivers of the Colorado mountains and foothills. Freshwater Biology,

55(9), 1973-1983.

Liess, A., Drakare, S., & Kahlert, M. (2009). Atmospheric nitrogen-deposition may

intensify phosphorus limitation of shallow epilithic periphyton in unproductive

lakes. Freshwater Biology, 54(8), 1759–1773.

Lu, C., & Tian, H. (2017). Global nitrogen and phosphorus fertilizer use for agriculture

production in the past half century: shifted hot spots and nutrient imbalance.

Earth System Science Data, 9(1), 181–192.

Mcdowell, R. W., Larned, S. T., & Houlbrooke, D. J. (2009). Nitrogen and phosphorus in

New Zealand streams and rivers: Control and impact of eutrophication and the

influence of land management. New Zealand Journal of Marine and Freshwater

Research, 43(4), 985–995.

Meyer, J. L., Strayer, D. L., Wallace, J. B., Eggert, S. L., Helfman, G. S., & Leonard, N.

E. (2007). The contribution of headwater streams to in river

networks. Journal of the American Water Resources Association, 43(1), 86–103.

Miltner, R. J. (2010). A method and rationale for deriving nutrient criteria for small rivers

and streams in Ohio. Environmental Management, 45(4), 842–855.

Miltner, R. J., & Rankin, A. E. T. (1998). Primary nutrients and the biotic integrity of

rivers and streams. Freshwater Biology, 40(1), 145–158.

37

Paerl, H. W., Scott, J. T., McCarthy, M. J., Newell, S. E., Gardner, W. S., Havens, K. E.,

Hoffman, D. K., Wilhelm, S. W., & Wurtsbaugh, W. A. (2016). It Takes Two to

Tango: When and Where Dual Nutrient (N & P) Reductions are needed to protect

lakes and downstream ecosystems. Environmental Science & Technology, 50(20),

10805–10813.

Peterson, B. J., Wollheim, W. M., Mulholland, P. J., Webster, J. R., Meyer, J. L., Tank, J.

L., Marti, E., Bowden, W. B., Valett, H. M., Hershey, A. E., McDowell, W. H.,

Dodds, W. K., Hamilton, S. K., Gregory, S., & Morrall, D. D. (2001). Control of

nitrogen export from watersheds by headwater streams. Science, 292(5514), 86–

90.

Pittelkow, C. M., Linquist, B. A., Lundy, M. E., Liang, X., Van Groenigen, K. J., Lee, J.,

van Gestel, N., Six, J., Venterea, R.T., & van Kessel, C. (2015). When does no-till

yield more? A global meta-analysis. Field Crops Research, 183, 156-168.

Power, M. E., & Stewart, A. J. (1987). Disturbance and recovery of an algal assemblage

following flooding in an Oklahoma stream. The American Midland Naturalist,

117(2), 333–345.

Powlson, D. S., Stirling, C. M., Jat, M. L., Gerard, B. G., Palm, C. A., Sanchez, P. A., &

Cassman, K. G. (2014). Limited potential of no-till agriculture for climate change

mitigation. Nature Climate Change, 4(8), 678–683.

Prairie, Y. T., Duarte, C. M., & Kalff, J. (1989). Unifying nutrient–chlorophyll

relationships in lakes. Canadian Journal of Fisheries and Aquatic Sciences, 46(7),

1176–1182.

38

Price, K. J., & Carrick, H. J. (2016). Effects of experimental nutrient loading on

phosphorus uptake by biofilms: evidence for nutrient saturation in mid-Atlantic

streams. Freshwater Science, 35(2), 503–517.

R Development Core Team (2008). R: A language and environment for statistical

computing. R Foundation for Statistical Computing, Vienna, Austria. ISBN 3-

900051-07-0, URL http://www.R-project.org.

Redfield, A. C. (1958). The Biological control of chemical factors in the environment.

American Scientist, 46(3), 205 – 221.

Reisinger, A. J., Tank, J. L., & Dee, M. M. (2016). Regional and seasonal variation in

nutrient limitation of river biofilms. Freshwater Science, 35(2), 474–489.

Rier, S. T., & Stevenson, R. J. (2002). Effects of light, dissolved organic carbon, and

inorganic nutrients on the relationship between algae and heterotrophic bacteria in

stream periphyton. Hydrobiologia, 489(1-3), 179–184.

Riseng, C. M., Wiley, M. J., & Stevenson, R. J. (2004). Hydrologic disturbance and

nutrient effects on benthic community structure in midwestern US streams: a

covariance structure analysis. Journal of the North American Benthological

Society, 23(2), 309–326.

Romaní, A. M., Guasch, H., Muñoz, I., Ruana, J., Vilalta, E., Schwartz, T., Emtiazi, F., &

Sabater, S. (2004). Biofilm structure and function and possible implications for

riverine DOC dynamics. Microbial Ecology, 47(4), 316–328.

Romaní, A. M., & Sabater, S. (1999). Epilithic ectoenzyme activity in a nutrient-rich

Mediterranean river. Aquatic Sciences, 61(2), 122–132.

39

Rosemond, A. D., Benstead, J. P., Bumpers, P. M., Gulis, V., Kominoski, J. S., Manning,

D. W. P., Suberkropp, K. & Wallace, J. B. (2015). Experimental nutrient

additions accelerate terrestrial carbon loss from stream ecosystems. Science,

347(6226), 1142–1145.

Ryder, D. S., Watts, R. J., Nye, E., & Burns, A. (2006). Can flow velocity regulate

epixylic biofilm structure in a regulated floodplain river? Marine and Freshwater

Research, 57(1), 29–36.

Schindler, D. W., Hecky, R. E., Findlay, D. L., Stainton, M. P., Parker, B. R., Paterson,

M. J., Beaty, K. G., Lyng, M., & Kasian, S. E. M. (2008). Eutrophication of lakes

cannot be controlled by reducing nitrogen input: Results of a 37-year whole-

ecosystem experiment. Proceedings of the National Academy of Sciences,

105(32), 11254–11258.

Sinsabaugh, R. L., & Foreman, C. M. (2003). 18 - Integrating dissolved organic matter

metabolism and microbial diversity: An overview of conceptual models. In S. E.

G. Findlay & R. L. Sinsabaugh (Eds.), Aquatic Ecosystems (pp. 425–454).

Burlington: Academic Press.

Sinsabaugh, R. L., Hill, B. H., & Follstad Shah, J. J. (2009). Ecoenzymatic stoichiometry

of microbial organic nutrient acquisition in soil and sediment. Nature, 462(7274),

795–798.

Smil, V. (2001). Enriching the earth: Fritz Haber, Carl Bosch, and the transformation of

world food production. MIT Press, Cambridge, MA. Enriching the earth: Fritz

Haber, Carl Bosch, and the transformation of world food production. MIT Press,

Cambridge, MA.

40

Smith, V. H., Tilman, G. D., & Nekola, J. C. (1999). Eutrophication: impacts of excess

nutrient inputs on freshwater, marine, and terrestrial ecosystems. Environmental

Pollution, 100(1), 179–196.

Smucker, N. J., DeForest, J. L., & Vis, M. L. (2009). Different methods and storage

duration affect measurements of epilithic extracellular enzyme activities in lotic

biofilms. Hydrobiologia, 636(1), 153–162.

Sobczak, W. V. (1996). Epilithic bacterial responses to variations in algal biomass and

labile dissolved organic carbon during biofilm colonization. Journal of the North

American Benthological Society, 15(2), 143–154.

Stainton, M. P. (1980). Errors in molybdenum blue methods for determining

orthophosphate in freshwater. Canadian Journal of Fisheries and Aquatic

Sciences, 37(3), 472–478.

Sterner, R. W., & Elser, J. J. (2002). : the biology of elements

from molecules to the biosphere. Princeton University Press.

Tank, J. L., & Webster, J. R. (1998). Interaction of substrate and nutrient availability on

wood biofilm processes in streams. Ecology, 79(6), 2168–2179.

USDA NASS. 2016. 2012 Census Volume 1, Chapter 1: State Level. Washington, DC:

USDA National Agricultural Statistics Service. https://www.

agcensus.usda.gov/Publications/2012/Full_Report/Volume_1,_Chapter_1_State_

Level/

Vitousek, P. M., Aber, J. D., Howarth, R. W., Likens, G. E., Matson, P. A., Schindler, D.

W., Schlesinger, W. H. & Tilman, D. G. (1997). Technical Report: Human

41

alteration of the global : sources and consequences. Ecological

Applications, 7(3), 737–750.

Watson, S. B., Miller, C., Arhonditsis, G., Boyer, G. L., Carmichael, W., Charlton, M.

N., Confesor, R., Depew, D. C., Hook, T.O., Ludsin, S. A., Matisoff, G.,

McElmurry, S. P, Murray, M. W., Peter, R. R., Roa, Y. R., Steffen, M. M. &

Wilhelm, S. W. (2016). The re-eutrophication of Lake Erie: Harmful algal blooms

and . Harmful Algae, 56, 44–66.

Wood, P. J., & Armitage, P. D. (1997). Biological effects of fine sediment in the lotic

environment. Environmental Management, 21(2), 203–217.

Ylla, I., Borrego, C., Romaní, A. M., & Sabater, S. (2009). Availability of glucose and

light modulates the structure and function of a microbial biofilm. FEMS

Microbiology Ecology, 69(1), 27–42.

Zhang, X., Davidson, E. A., Mauzerall, D. L., Searchinger, T. D., Dumas, P., & Shen, Y.

(2015). Managing nitrogen for sustainable development. Nature, 528(7580), 51–

59.

42

Table 2.1. Land use and physical characteristics of the Grand, Maumee, and Portage river watersheds. Other = land use categories of Barren, Shrub, and Grassland.

Annual Lake Size Agriculture Forest Urban Other Precipitation Erie Physiographic Watershed (km2) (%) (%) (%) (%) (%) (mm) Soil Type Basin Region Low Lime Glaciated Glacial Alleghenny Grand 1831 33 6 43 11 6 1075 Drift Central Plateau High Lime Glacial Lake Huron-Erie Maumee 21540 78 2 7 11 1 870 Sediments Western Lake Plains High Lime Glacial Lake Huron-Erie Portage 1984 76 4 4 13 2 870 Sediments Western Lake Plains

43

Table 2.2. Mean (± SE) water column nutrient concentrations for each stream during the three sampling periods. NO2-N + NO3-N are nitrogen, SRP is soluble reactive phosphorus, and TP is total phosphorus in µg/L. N:P is the molar ratio of nitrogen to phosphorus in the water column.

Post-Fertilizer Pre-Harvest Post-Harvest NO -N + NO -N + NO -N + Watershed 2 SRP 2 SRP 2 SRP NO -N TP (µg/L) N:P NO -N TP (µg/L) N:P NO -N TP (µg/L) N:P Stream 3 (µg/L) 3 (µg/L) 3 (µg/L) (µg/L) (µg/L) (µg/L) Grand Baughman 355.6 (18) 53.5 (12) 719.2 (200) 13.2 (7.3) 408.0 (25) 30.1 (5.8) 63.7 (19) 64.0 (60) 1100 (18) 19.1 (7.5) 27.7 (3.5) 58.3 (30) Creek 401.1 (8.8) 35.5 (9.4) 355.1 (77) 31.9 (13) 204.7 (7.4) 46.3 (16) 84.3 (47) 13.6 (8.7) 273.6 (2.0) 15.5 (7.6) 42.1 (6.0) 66.8 (40) Center Creek 119.3 (3.0) 29.8 (8.8) 248.3 (85) 9.3 (1.4) 80.3 (2.1) 29.5 (8.6) 77.3 (12) 6.4 (1.7) 1060 (7.0) 12.3 (3.2) 43.7 (6.5) 120.7 (46) Phelps Creek 891.0 (26) 4.0 (0.6) 93.5 (27) 680.1 (397) 919.1 (43) 11.5 (2.3) 18.2 (3.5) 245.7 (160) 2660 (5.6) 3.3 (0.8) 5.1 (3.1) 515.3 (82) Swine Creek Maumee 7668 (9.5) 51.7 (13) 318.9 (108) 2373 (1400) 9.1 (0.2) 5.5 (1.7) 32.1 (3.2) 4.7 (1.5) 21100 (40) 19.5 (4.5) 108.9 (17) 3060 (1260) Blue Creek 6777 (18) 68.5 (18) 243.4 (34) 279.4 (81) 236.6 (9.6) 104.7 (19) 111 (66) 13.1 (6.6) 833.0 (3.0) 70.0 (35) 117.3 (25) 13.0 (4.5) Eagle Creek Flat Rock 6442 (20) 66.3 (9.9) 392.9 (81) 226.5 (40) 20.4 (1.0) 47.5 (16) 108 (7.3) 1.0 (0.5) 19920 (19) 3.3 (1.0) 51.6 (8.0) 2705 (570) Creek 9033 (43) 32.5 (5.0) 264.7 (41) 661.2 (103) 27.5 (0.5) 50.4 (9.5) 91.4 (15) 1.4 (0.4) 251.8 (1.4) 101.2 (92) 109.4 (45) 34.4 (14) Lye Creek 7574 (7.0) 49.7 (6.3) 285.8 (53) 358.6 (57) 25.6 (0.7) 27.7 (5.4) 53.9 (9.6) 3.0 (1.4) 80.1 (0.2) 30.9 (5.2) 84.9 (8.0) 2.0 (0.8) Ottawa Creek Portage Middle 8435 (20) 19.8 (3.7) 220.1 (51) 1078 (155) 3769 (36) 161 (15) 293 (33) 51.5 (4.7) 5064 (14) 34.5 (13) 61.4 (17) 440 (106) Branch Portage River North Branch 7403 (40) 46.8 (36) 327.9 (30) 1639 (850) 1725 (108) 81.6 (35) 154 (61) 24.8 (17) 19680 (59) 88.5 (58) 90.6 (42) 3980 (1260) Portage River

44

Table 2.3. Mean periphyton nutrient stoichiometry, δ15N, and eco-enzyme activities for each stream during the three sampling periods: post- fertilizer application (May 2017), pre-harvest (September 2017), and post-harvest (November 2016). Ecoenzyme activities are normalized to nmol/g C/h. Values in parentheses are standard error of the mean. Abbreviations: N:P = molar N:P of periphyton, LAMP = leucine amino peptidase, PHOS = phosphatase, BG = β-glucosidase, BX = β-xylosidase. Missing values are denoted with --.

Watershed Stream δ15N %N %P N:P LAMP PHOS BG BX BG:BX PHOS:LAMP Post-Fertilizer Grand Baughman Creek -- -- 1.1 -- 374500 41500 9000 970 9.28 0.11 Center Creek -- -- 1.0 -- 68400 36300 1800 270 6.67 0.53 Phelps Creek -- -- 1.0 -- 86800 21300 3300 250 13.2 0.25 Swine Creek -- -- 0.5 -- 76400 28500 3700 170 21.8 0.37 Maumee Blue Creek -- -- 0.8 -- 54600 34500 1800 420 4.29 0.63 Eagle Creek -- -- 1.0 -- 69600 16100 2500 450 5.56 0.23 Flat Rock Creek -- -- 0.9 -- 39100 13500 1900 280 6.79 0.34 Lye Creek -- -- 0.9 -- 102700 29400 4500 930 4.84 0.29 Ottawa Creek -- -- 1.0 -- 120200 34500 3800 980 3.88 0.29 Portage Middle Branch Portage River -- -- 0.9 -- 62100 18000 1100 150 7.33 0.29 North Branch Portage River -- -- 1.1 -- 18100 13600 700 120 5.83 0.75 Pre-Harvest Grand Baughman Creek -- -- 1.2 -- 1101559 267878 29950 2257 13.3 0.24 Center Creek -- -- 1.6 -- 1042558 484103 24922 2167 11.5 0.46 Phelps Creek -- -- 0.8 -- 1631872 272063 47786 3861 12.4 0.17 Swine Creek -- -- 0.8 -- 905859 287537 23722 2159 11.0 0.32 Maumee Blue Creek -- -- 1.5 -- 609935 113233 16080 1922 8.37 0.19 Eagle Creek -- -- 1.0 -- 1601924 132529 11928 2001 5.96 0.08 Flat Rock Creek -- -- 1.0 -- 2019537 124165 75751 17958 4.22 0.06 Lye Creek -- -- 0.8 -- 965445 217993 17288 3364 5.14 0.23 Ottawa Creek -- -- 0.7 -- 955697 260252 13988 2429 5.76 0.27 Portage Middle Branch Portage River -- -- 0.7 -- 674344 116451 12680 2552 4.97 0.17 North Branch Portage River -- -- 1.1 -- 863761 231321 19309 6960 2.77 0.27 Post-Harvest Grand Baughman Creek 7.1 13.5 0.9 42.5 (6.3) 75810 19858 6167 774 7.97 0.26 Center Creek 8.2 3.2 1.4 30.2 (4.3) 163462 23368 12068 1222 9.88 0.14 Phelps Creek 6.4 8.1 1.1 47.8 (6.4) 84727 20610 9328 674 13.8 0.24 Swine Creek 4.7 7.8 0.9 57.8 (9.7) 107153 35689 7380 815 9.06 0.33 Maumee Blue Creek 8.7 5.7 1.0 37.8 (8.0) 68750 20341 1731 136 12.7 0.30 Eagle Creek 8.3 3.9 1.1 37.4 (3.6) 652043 56979 6120 1224 5.00 0.09 Flat Rock Creek 8.6 2.8 1.4 26.1 (5.7) 31177 12087 1417 236 6.00 0.39 Lye Creek 7.9 4.4 0.7 32.5 (8.6) 71707 12040 5651 450 12.6 0.17 Ottawa Creek 7.7 5.3 0.8 47.5 (4.5) 261403 22959 2574 342 7.53 0.09 Portage Middle Branch Portage River 8.8 5.1 0.9 45.7 (8.7) 264111 27668 3002 339 8.86 0.10 North Branch Portage River 9.5 6.7 0.9 48.6 (4.5) 225643 30388 2208 584 3.78 0.13

45

Table 2.4. Spearman correlations (r) of stream water chemistry, and biofilm biomass, stoichiometry and eco-enzyme activities for post-fertilizer application (May 2017), pre- harvest (September 2017), and post-harvest (November 2016) sampling events.

Correlations in bold are significant at p < 0.05. Abbreviations: Chl = biofilm chlorophyll- a, AFDM = ash free dry mass, N:P = molar N:P ratio in periphyton, N:Pw = molar N:P in water column, LAMP = leucine amino peptidase activity, PHOS = phosphatase activity,

BG = β-glucosidase activity, BX = β-xylosidase activity.

Post-Fertilizer

AFDM N:P SRP NO3 N:Pw LAMP PHOS BG BX PHOS:LAMP BG:BX Chl 0.83 -- -0.01 0.44 0.34 0.18 0.43 0.21 0.41 0.18 -0.44 AFDM -- 0.01 0.57 0.43 0.37 0.6 0.44 0.64 0.05 -0.6 N:P ------SRP -0.02 -0.43 -0.28 -0.14 0 0.16 0.29 -0.33

NO3 0.83 0.36 0.5 0.2 0.3 -0.07 -0.3 N:Pw 0.3 0.41 0.07 0.18 -0.13 -0.23 LAMP 0.67 0.82 0.7 -0.67 -0.25 PHOS 0.66 0.67 0.03 -0.45 BG 0.84 -0.4 -0.34 BX -0.23 -0.76 PHOS:LAMP -0.12 BG:BX Pre-Harvest Chl 0.41 -- -0.09 -0.12 -0.07 -0.2 -0.19 0.05 0.25 0.06 -0.26 AFDM -- 0.24 -0.17 -0.27 0.24 -0.2 0.49 0.66 -0.31 -0.47 N:P ------SRP 0.42 -0.04 -0.04 -0.3 -0.12 0.38 -0.14 -0.5

NO3 0.86 0 0.05 0.04 -0.1 0.06 0.17 N:Pw 0.06 0.11 0.14 -0.26 0.03 0.4 LAMP 0.47 0.68 0.22 -0.82 0.3 PHOS 0.09 -0.16 0.04 0.25 BG 0.5 -0.67 0.12 BX -0.39 -0.76 PHOS:LAMP -0.12 BG:BX Post-Harvest - Chl 0.55 0.09 0.08 0.19 0.11 0.06 -0.02 -0.04 -0.29 -0.19 0.27 AFDM 0.15 -0.02 -0.04 0.04 -0.11 -0.35 0.27 -0.22 -0.2 0.6 N:P -0.17 -0.05 -0.02 -0.31 0.2 0.33 -0.13 0.08 0.13 SRP -0.18 -0.52 0.29 0.28 0.05 0 -0.19 0.11 NO3 0.88 -0.16 0.17 -0.35 -0.13 0.34 -0.23 N:Pw -0.27 0 -0.24 -0.09 0.36 -0.19 LAMP 0.56 0.34 0.4 -0.64 -0.03 PHOS 0.03 0.35 0.16 -0.36 BG 0.67 -0.38 0.46 BX -0.15 -0.28 PHOS:LAMP -0.27 BG:BX

46

Table 2.5. N-, P-, or co-limitation of stream sites based on molar stoichiometry of water and sediments, and on the relative activities of phosphorus- to nitrogen-acquiring eco- enzymes (PHOS:LAMP) for each watershed during post-fertilizer application (May

2017), pre-harvest (September 2017), and post-harvest (November 2016) sampling events.

Post-Fertilizer Pre-Harvest Post-Harvest Water Water Water Periphyton Watershed Stream Site PHOS:LAMP PHOS:LAMP PHOS:LAMP N:P N:P N:P N:P Baughman Grand 1 N P P P N N P Creek 2 N P N P N Co- P 3 N P N P N Co- P 4 P N P N P P P Center Creek 1 P P N P N N P 2 P P P P P P P 3 P P P P N Co- P Phelps Creek 1 P Co- N P P P P 2 P P N P P Co- P 3 P P P P P Co- P 4 N Co- P Co- N Co- P 5 P P P P N Co- P Swine Creek 1 P P P P P P P 2 P P P Co- P P P 3 P P P P P P P 4 P P P P P P P 5 P P P P P P P 6 P Co- P P P P P Maumee Blue Creek 1 P P N Co- P P P 2 P P N P P P P 3 P P N P P Co- P 4 -- P P P P P P 5 P P P P P P Co- 6 P P P P P Co- P Eagle Creek 1 P P -- P N N P 2 -- P N P P P P 3 P P P P N N P 4 -- P N Co- N P P 5 P Co- P N N N P Flat Rock 1 P Co- N N P P P Creek 4 P P N P P P Co- Lye Creek 1 P P N P N Co- P 2 P P N P N N P 3 P Co- N Co- P Co- P 4 P P N P P P N Ottawa Creek 1 -- P P Co- N N P 2 -- P N P N P P 3 P Co- N P N N P 4 P P N P N N P 5 P P N P N Co- P Middle Branch Portage 1 P P P P P N P Portage River 2 P P P Co- P Co- P 3 P P P P P Co- P 4 P P P P P Co- P 5 P P P P P Co- P 6 P P P P P P P North Branch 1 P P P P P P P Portage River 2 P P N Co- P P P 5 P P P P P N P

47

Table 2.6. Discharge and water column metrics for main stem rivers averaged from the month prior to sampling upstream sites. Nitrate (NO3), nitrite (NO2), soluble reactive phosphorus (SRP), and total phosphorus (TP) are reported in µg/L. Total suspended solids is in mg/L and Discharge is in L/s. Data were downloaded from the National

Center for Water Quality Research.

Sample Discharge Suspended TP SRP NO2 + NO3 Month Year River Period (L/s) solids (mg/L) (µg/L) (µg/L) (µg/L) Nov 2016 Post-harvest Maumee 53700 30.7 150 50 3470 May 2017 Post-fertilizer 775500 172.7 450 110 5750 Sept 2017 Pre-harvest 24300 28.4 170 40 220 Nov 2016 Post-harvest Grand 21100 74.6 160 60 3420 May 2017 Post-fertilizer 45300 314.3 270 20 1280 Sept 2017 Pre-harvest 10500 28.8 150 80 4090 Nov 2016 Post-harvest Portage 730 8.0 150 100 2890 May 2017 Post-fertilizer 37500 76.8 250 80 8850 Sept 2017 Pre-harvest 1040 6.2 130 80 1600

48

(! Blue Creek North Branch Portage River Eagle Creek ! Baughman Creek (! Flat Rock Creek (! Center Creek 012.5 25 50 75 (! Lye Creek Phelps Creek Kilometers Ottawa Creek Swine Creek Middle Branch Portage River Lake Erie

!!! ((!

(! (!(!(! (! (!(! (!(! (! (!(! (!

Figure 2.1. Map of sampling locations for three watersheds within the Lake Erie

Watershed drainage basin and land use from the National Land Cover Database (2011).

Maumee River Watershed: Blue Creek, Eagle Creek, Flat Rock Creek, Lye Creek, and

Ottawa Creek. Portage River Watershed: Middle Branch Portage River and North Branch

Portage River. Grand River Watershed: Baughman Creek, Center Creek, Phelps Creek, and Swine Creek. Land use categories: crop agriculture (brown), pasture (yellow), forest

(green), developed (red), water (blue).

49

200 150

Grand Grand Maumee Maumee Portage 150 Portage  2

m 100  2 mg m   g

100   AFDM

50 Chlorophyll-a

50

0 0

Post Fertilizer Pre Harvest Post Harvest Post Fertilizer Pre Harvest Post Harvest

Figure 2.2. Mean algal biomass as chlorophyll-a (mg/m2) and periphyton organic matter content (AFDM in g/m2) for each watershed post-fertilizer application, pre-harvest, and post-harvest. Horizontal lines represent median values and boxes represent inter-quartile ranges (25-75% percentiles). Whiskers show the range and points represent outlier values.

50

100.6 Post Fertilizer Pre Harvest Post Harvest

0.4  10 2 m mg   0.2 10

Chlorophyll-a 100

10-0.2

10-0.2 100 100.2 100.4 100.6 Sediment Contentg m2

Figure 2.3. Mean site periphyton algal biomass (chlorophyll-a in mg/m2) and sediment content (in g/m2) for post-fertilization, pre-harvest, and post-harvest in log-log scale.

51

pNO3 pNP.1

pSRP pNO2

pTP CCA2 (20%) CCA2 1 505 0 -5 -10

-10 -5 0 5

CCA1 (66%)

Figure 2.4. Canonical correspondence analysis of biofilms for the Maumee (squares),

Portage (triangles), and Grand (circles) River watersheds post-fertilizer (black), pre- harvest (dark grey), and post-harvest (light grey). Response variables are biofilm chlorophyll-a (mg/m2), ash free dry mass (g/m2), percent organic matter, sediment content (g/m2), and P content (g/m2) and the predictor variables soluble reactive phosphorus (SRP in µg/L), nitrate (NO3, in µg/L), nitrite (NO2, in µg/L), water column

N:P (NP), and total phosphorus (µg/L).

52

pAsh pAFDM CCA2 (18%) CCA2 pSRP pChl

pNO3 1 020 10 0 -10 pPOM

-30 -20 -10 0 10

CCA1 (70%)

Figure 2.5. Canonical correspondence analysis of biofilms for the Maumee (squares),

Portage (triangles), and Grand (circles) River watersheds post-fertilizer (black), pre- harvest (dark grey), and post-harvest (light grey). Response variables are biofilm extracellular enzyme activities (phosphatase, leucine aminopeptidase, β-glucosidase, and

β-xylosidase in nmol/g C/h) and enzyme ratios (PHOS:LAMP and BG:BX) and the predictor variables are water column SRP and NO3, and biofilm chlorophyll-a, AFDM, sediment content (Ash), and percent organic matter (POM).

53

Post Fertilizer Pre Harvest Post Harvest

50 g/L) 

0 Change in SRP( in Change

-50

Blue Eagle Flat Lye Ottawa Middle North

Figure 2.6. The difference in water column soluble reactive phosphorus concentrations

(SRP, µg/L) between upstream sites in this study and the main stem rivers for the Portage and Maumee River. Stream abbreviations: Maumee River Watershed: Blue Creek = Blue,

Eagle Creek = Eagle, Flat Rock Creek = Flat, Lye Creek = Lye, Ottawa Creek = Ottawa,

Portage River Watershed: Middle Branch Portage River = Middle, North Branch Portage

River = North.

54

Chapter Three

SEDIMENT MICROBIAL ASSEMBLAGE COMPOSITION IN A HETEROGENEOUS RESOURCE LANDSCAPE: LONGITUDINAL TRANSITIONS IN BIOFILM COMMUNITIES

Introduction Headwater streams are intimately linked to the adjacent landscapes through terrestrial particulate carbon inputs from riparian vegetation (Hynes 1975, Vannote et al.

1980). As you move downstream, the canopy opens and streams widen. Before the river becomes too deep for light to penetrate to the bottom, attached algae are also an important carbon source. Consequently, light availability and terrestrial particulate carbon (C) covary longitudinally leading to variation in allochthonous (terrestrial origin) and autochthonous (attached algae) resources. The effect of this variation in carbon availability on stream structure and function has largely focused on consumers (Wallace and Webster 1996, Wallace et al. 1999, Manning et al. 2016). Variation in C quality and quantity is important for bacteria, which carry out the mineralization of particulate organic C (Rosemond et al. 2015, Collins et al. 2016, Fabian et al. 2017). Respiration from headwaters contributes substantially to global carbon fluxes, emitting a large amount of carbon dioxide (1.55 Pg C y–1) into the atmosphere (Battin et al. 2008). Thus, biofilm communities represent a metabolically active component of carbon cycling in freshwater ecosystems and understanding how microbial communities change in response

55

to carbon availability may provide insight to the responses of carbon cycles to land use change.

Midwestern streams historically had woody riparian vegetation that stabilized banks and contributed terrestrial organic C while simultaneously intercepting light.

Agricultural practices have removed the forested vegetation in low-order streams

(Sweeney et al. 2004). This vegetation removal resulted in increased sediment and nutrient loadings and reduced terrestrial organic C input (Findlay and Sinsabaugh, 2006).

The heterogeneity in water chemistry and organic matter caused by land use alterations might shape the structure of biofilm communities including those growing on sediments

(Jones et al. 2009, Shade et al. 2012).

Unconsolidated sediments in streams are a composite of inorganic, detrital, and living components. Flow dynamics, riparian land use, upstream habitat, and geological features determine the composition and distribution of stream sediments (Wetzel 2001).

Stream bottoms are structurally heterogeneous owing to particle sorting associated with variation in flow. In small streams, the stream bottom in areas of low flow is dominated by unconsolidated sediments. Increased discharge associated with rain or snowmelt events regularly suspends and resorts stream sediments. Sediment particle size determines both the total surface area available for microbial colonization as well as oxygen concentrations in the pore water (Dale et al. 1974, Battin et al. 2003b). Thus, unconsolidated sediments are a dynamic and ephemeral substratum for bacterial communities that are rarely studied relative to hard substrates such as rocks.

The amount of organic matter in stream sediments reflects a balance between organic matter loading to the sediments, resuspension and export of organic matter, and

56

decomposition rates. The organic component is derived from terrestrial detritus, primarily from leaf litter, dissolved organic carbon, and living and dead algae, plants, fungi, bacteria, and metazoans. Terrestrial organic matter can either enter streams directly from the canopy through leaf fall or laterally from stream banks (Lisboa et al. 2015). Leaves can represent 41–98% of total organic matter moving from terrestrial to stream ecosystems (Franca et al. 2009). The high standing stock of terrestrial particulate organic matter is a vital carbon resource for stream consumers and (Brett et al.

2017). Within-stream retention time and export of organic matter depends on flow velocity and decomposition rates (Lisboa et al. 2015). Terrestrial detritus has a high amount of structural material, is relatively difficult to break down and is broken down over time into smaller size fractions (Sabater et al. 2006). Thus, this highly abundant and available carbon resource is of low quality to consumer and organsims.

Within-stream sources of organic carbon are primarily phototrophic eukaryotic algae and cyanobacteria growing on bottom surfaces. High light availability and high nutrient concentrations increase the growth and biomass of (Warren et al.

2017). Grazing by metazoans and high flow velocity reduces the accumulation of attached algal and cyanobacterial biomass (Power et al. 1988, Ryder et al. 2006). Algal detritus is broken down easily and relatively fast by bacteria which preferentially decompose labile matter over more refractory compounds that are often found in terrestrial matter (Shivers et al. 2016). High light from the removal of riparian forested vegetation in agricultural streams shifts the balance of organic carbon availability from recalcitrant and relatively low quality terrestrial detritus to labile and high quality algae.

57

The heterotrophic microbial community members, which include fungi, bacteria, and archaea, are primarily responsible for the breakdown and recycling of detritus. The breakdown of sediment detritus by decomposers is regulated by ambient temperature, nutrient availability and the lability of the detritus itself (Sabater et al. 2006). Light also affects bacterial community composition through fluctuations in phototrophic biomass and primary productivity (Wagner et al. 2015). Freshwater sediments contain some of the most diverse heterotrophic communities (Battin et al. 2016), and ultimately, environmental variables determine the members of these diverse biofilm community

(Goodman et al. 2015). Nevertheless, two phyla, Bacteroidetes and Proteobacteria, dominate community relative abundances in stream sediments.

From a watershed perspective, the interaction between organic C availability and microbial community composition has important implications for organic C movement and processing along hydrological flow paths. Agricultural practices cause variation in the relative availability of the resources of light, nutrients, sediments, and carbon. This heterogeneous resource landscape should generate spatial variation in microbial community composition. My main objective was to assess the compositional variability in sediment microbial communities within a stream environment that has different carbon resource availability due to variation in riparian canopy cover. I investigated how the quality and source of organic carbon is related to sediment microbial community structure. Further, I qualitatively examined whether differences in microbial community structure were related to environmental characteristics at each site.

58

Methods Study site The Little Miami River is a national scenic river flowing through the southwestern corner of Ohio (Ohio EPA 1995). The Little Miami is one of the most biologically diverse rivers in the region, home to 113 fish species, some of which are rare and endangered (Harrington 1999). This region experiences warm summers (15-30°C) and moderately cold winters (-10 – 0°C) with an average annual rainfall between 90-110 cm. The Little Miami is embedded in a predominately agricultural watershed that has protected forest fragments.

I selected sampling locations along a stretch of the Little Miami that varied in forested riparian area in an attempt to capture a natural variation in terrestrial organic carbon subsidies and attached algal biomass (Figure 3.2). The headwater site (LAt) has high light (L), abundant attached algae (A) and low terrestrial organic C input (t). Site

LAt has row crop agriculture (usually corn) on both banks and is close to a dairy farm.

The riparian vegetation consists of grass a few meters wide along the banks. The river at

LAt is highly channelized and there is no tree canopy. A buffer strip of riparian trees begins immediately downstream from site LAt. Although the Little Miami River continues through agricultural land, this buffer extends, virtually uninterrupted, for 30 kilometers, well below the most downstream site that we sampled. The next site downstream (~ 4 km downstream, laT) has low light (l), low attached algal abundance

(a), and high terrestrial organic C input (T). At site laT the stream is still relatively small, and this site has the most robust canopy with a forested riparian buffer strip of approximately 200 m wide. The third site (lat) has relatively low light (l), low attached algal abundance (a) and low terrestrial organic C input (t). The site lat has a relatively 59

small riparian buffer strip composed of primarily mixed brush and immature trees that do not shade the entire stream (buffer width ~ 70 m). The most downstream site (LAT) has relatively high light (L), abundant attached algal biomass (A), and high terrestrial organic

C input (T). LAT is wider (16.6 m) and the stream has travelled through several kilometers of forested area, including John Bryan State Park and the Clifton Gorge State

Nature Preserve. LAT has a riparian area composed of mature trees (buffer width ~ 870 m) but the stream has widened so light penetrates to the sediment surface in the middle portions of the stream. Thus, LAT likely has greater algal organic C per unit area than lat and laT.

Physicochemical Characteristics Field measurements of pH, conductivity, and dissolved oxygen were taken in the stream thalweg near midday. I collected water samples in triplicate for soluble reactive phosphorus (SRP) and chlorophyll-a analyses. Samples were stored on ice and transferred to the laboratory for processing the same day. I filtered chlorophyll-a samples on PALL

AE glass fiber filters and frozen at -20°C for further analysis. SRP samples were filtered on preashed PALL AE glass fiber filters (1 µm nominal pore size) and frozen at -20°C.

Filters were used for determination of suspended solids as loss on ignition (LOI) at

500°C for 1 hour (APHA 2005). I measured SRP by the molybdate method (Stainton et al. 1980). Chl-a was extracted at 4°C for 24 hours in 90% buffered ethanol. I measured chlorophyll-a and phaeophyton fluorometrically with an acidification step (Arar and

Collins 1997).

In addition to measuring light availability using a Licor sensor, I downloaded data from the Ohio Agriculture and Research Development Center (OARDC) meteorological

60

station to assess differences in light availability to the . I converted light irradiance data from meteorological data as in Francoeur 2017 and Winslow et al. 2014.

Specifically, I used solar radiation measurements from the OARDC weather station located in Clark County Ohio. Because the meteorological station reported irradiance in kJ/m2/h, I converted units to PAR using several steps: 1) to convert shortwave energy to quanta I multiplied shortwave radiation (kJ/m2) by the constant 5.03 (Wetzel, 2001), 2) to account for the proportion of PAR in shortwave radiation, I multiplied quanta by the constant 0.46 (Kirk 1994, Hiriart-Baer et al. 2008), and 3) to account for surface reflection, I reduced values of PAR by 5% (e.g. Middelboe and Markager, 2003). I used this data and measured Licor data to calculate percent light reaching the stream at each site.

Sediment Particle Size Distribution I collected 10 samples of sediment at each site along a transect using a sediment corer or trowel. This sample was refrigerated and then dried at 60 °C. I measured grain- size distribution by dry-sieving each sample through a sieve series (16-, 4-, 1.68-, 1-, 0.5-

, 0.297-, 0.25-, 0.125-, 0.105-, 0.074-, 0.062-, 0.053-, and 0.044-mm mesh). For each sieve size used, I calculated the percentage by mass of the sediment sample that is finer than the ith sieve size.

푀 푁 = = 1− × 100% 푀 where 푀 is the initial dry mass of the sample, 푀 is the mass of sediment retained on the jth sieve, and Nj, the percentage by mass of the sediment sample that is finer than the ith sieve size. From these values, I calculated the uniformity coefficient (퐶 ), and the coefficient of gradation: 61

퐷 퐶 = 퐷

(퐷) 퐶 = (퐷 × 퐷) where values of 퐷 are found by plotting particle-size distribution curves. The particle diameters are plotted in log scale, and the corresponding percent finer in arithmetic scale.

In this manner, the diameter in the particle-size distribution curve corresponding to 10% finer is defined as 퐷 and so on for each value of 퐷. 퐶 and 퐶 are metrics describing the graded nature of sediments where a value > 4 or ~ 1, respectively, indicate well- graded sediments.

Organic Component of Sediment I collected 8 replicates of undisturbed sediments by inserting an acrylic tube (5 cm diameter) into the sediments along a transect within a run habitat at each site. The top

6.6 mm of sediment was collected by pushing the core up above the top of the acrylic corer until 6.6 mm was exposed. I used a spatula to collect this 6.6 mm slice from each core and a small subsample was reserved for analysis of water content. I sieved sediment samples to remove large stones. Half of each replicate was used for energy content analysis and the other half was split into subsamples for analyses of algal biomass, organic matter content, and elemental stoichiometry. All subsamples were lyophilized, homogenized, and frozen at -20°C.

Sediment Energy Content I used a combination of wet sieving and decanting of sediment sample to collect the organic matter fraction. Large organic debris was visually identified and collected from larger sieve fractions. All organic material from each replicate was combined into

62

one final sample for analysis. I measured organic matter content of the remaining sediment as loss on ignition at 500°C. To measure the calorie content, I combusted the organic matter fraction using a Parr 6200 Bomb Calorimeter.

Attached algal biomass and total organic matter content I subsampled 25 - 40 mg of homogenized sediments for chlorophyll-a extraction and 1-2 g for organic matter content measured as ash-free dry mass (AFDM).

Chlorophyll-a was analyzed as described above. AFDM was measured as LOI at 500 °C for 4 hours.

Sediment stoichiometry To remove excess carbonate, I acidified subsamples of sediment with 1N-HCl until the sample stopped emitting CO2 gas. I measured P content on a subsample of sediment by the molybdate method (Stainton et al. 1980) after combustion for 1 hour at

500°C. Percent C and percent N were analyzed using an isotope-ratio mass spectrometer

(Finnigan MAT Delta Plus) coupled to an elemental analyzer (Carlo Erba NC2500) at the

Cornell University Stable Isotope Laboratory. Replicates were run on 8% of the samples to determine analytical error.

Microbial community composition I collected 8 sediment samples from each site using sterile 50-ml centrifuge tubes.

Samples were collected across two transects perpindicular to streamflow and immediately placed on ice for transport to the lab. Once at the lab, each tube was centrifuged at 4500 rpm. Overlying water was decanted and discarded. I homogenized the sediment sample with a sterile spatula and subsampled this bulk sample into three microcentrifuge tubes. I immediately froze samples at -81°C for further analyses. One tube was used for bulk community composition analysis (DNA) and one sample for total lipids analysis. 63

DNA was extracted using the PowerSoil® DNA Kit according to the manufacturer’s protocol. Prokaryotic microbial community interrogation was performed by amplifying the V4 variable region of the 16S ribosomal RNA gene. Amplification was conducted using the Ion Torrent (Life technologies) compatible fusion primer set which consists of the sequencing adapters (A and trP1), 6 nucleotide barcodes (for multiplexing) and the universal primers 515F (5-GTGCCAGCMGCCGCGGTAA) and 806R (5-

GTGCCAGCMGCCGCGGTAA). Amplified products were purified using QiaQuick

PCR purification columns (Qiagen) and were size selected using the Pippin Prep electorphoresis gel system (Sage Science). Purified and size-selected products were equimolarly pooled and used to construct the sequencing library with the Ion PGM

Template OT2 400 kit (Life technologies). The high-quality library was sequenced on an

Ion Torrent PGM 316 chip v2 following the manufacturer's guidelines. The sequencing run resulted in an average of 25,000 reads per sample.

Raw sequencing output was first quality-filtered to remove low-quality reads

(Average Q < 25) and short reads (< 100bp). High-quality reads were then demultiplexed using the 6 nt error-correcting barcodes. Demultiplexed sequence data were analyzed in

Qiime using the default pipeline (Caporaso et al. 2010). Operational taxonomic units were determined by clustering at 97% sequencing similarity using the UCLUST algorithm (Edgar 2010) and annotated using the GreenGenes database v13.8 (DeSantis et al. 2006).

Microbial community biomass I analyzed total lipids (TL) on ~ 100 mg of sediment from each site. In order to obtain enough mass to extract lipids, I combined replicates into 2 samples per site. Total

64

lipids were extracted using 4 ml of phosphate buffer and 2:1 methanol:chloroform (White et al. 1979, DeForest et al. 2012). A 19:0 phospholipid standard (1,2-dinonadecanoyl-sn- glycerol-3-phosphocholine, Avanti Polar Lipids, Alabaster, Alabama, USE) was added to determine analytical recovery. Lipids were extracted with chloroform: methanol: water

(4:2:1) from freeze-dried sediments. Sonication (10 min) was used to enhance lipid extraction, and samples were centrifuged to facilitate phase separation. The chloroform phase was transferred to a new tube. Chloroform was evaporated under an N2 gas stream, and the remaining lipids were dissolved in toluene. Methanolic H2SO4 (1% v/v) was added to produce FA methyl esters (FAMEs), and samples were transmethylated overnight at 50°C in a water bath.

FAMEs samples were analyzed using a gas chromatograph (Shimadzu Ultra) equipped with mass detector (GC-MS) at the University of Jyväskylä (Finland).

Biomarker abundance was calculated by dividing individual biomarkers by total biomass

(i.e., % mol fraction). I removed rare biomarkers if they had low biomass (<1% mol fraction) from subsequent analyses. I combined known biomarkers into the groups for bacteria, fungi, algae to obtain relative abundance and biomass (µmol Fatty Acid/ mg dry mass) of these groups. Algal and cyanobacterial markers included: 16:2ω6, 16:2ω4,

16:3ω3, 16:4ω3, 18:4ω3, 20:2ω6, 20:3ω3. Bacterial markers included: 16:1ω9, 16:1ω8,

16:1ω7, 16:1ω6, i-17:0, a-17:0, 15:1ω7, i-16:0, i-15:0, a-15:0, 18:1ω9, 18:1ω7, 18:3ω6,

20:1ω9, 22:1ω9, 22:1ω7.

Statistical Analyses I used Pearson product–moment correlations to examine the extent of multicollinearity among independent environmental variables. I used one-way ANOVA

65

and Tukey’s Honestly Significant Difference (HSD) post-hoc test to assess whether organic matter quantity or quality, light intensity, SRP, energy content, and microbial biofilm biomass differed significantly among sites. Community diversity and evenness were assessed by calculating the Shannon’s H’ (diversity) and Simpson’s E (evenness) on

OTU data. The Bray-Curtis index is one of the best indices for detecting gradients in species composition (Michin 1987), and I used it to assess community dissimilarity across sites. I used the relative abundance data to produce rank-abundance plot as log(10) relative abundance vs. rank (from most abundant to least), for each site. I used principal components analysis (PCA) to visualize patterns among site microbial communities and

Canonical Correspondence Analysis (CCA) to assess the environmental drivers of community composition. Significance of the constraints used in CCA was assessed using the permutest function in R Version 3.2.3 (R Core Team, 2015) with 1000 permutations.

Significant models required a p < 0.05.

Results Sediment heterogeneity and environmental conditions

Sediment particle size distribution assessed using 퐶 and 퐶 did not vary significantly between sites (ANOVA, F3,12 = 0.836 and 1.126, respectively, all p > 0.05,

Table 3.2). Mean SRP ranged from 500 ± 19.0 to 1120 ± 177 µg/L (± SD) and the most upstream site (LAt) had significantly higher SRP concentrations than three downstream sites as determined by post-hoc Tukey’s HSD (ANOVA, F3,14 = 37.45, p < 0.05, Figure

3.3a). Mean solar radiation (measured at noon) ranged between 160 μmol/m2/s and 1500

μmol/m2/s across the study sites. The percent of total solar radiation reaching the stream bottom varied significantly among sites (ANOVA, F3,14 = 71.92, p < 0.001).

66

Variability of organic matter and biofilm biomass Sediment organic matter content varied between 110 and 517 mg/m2, and was highly variable within sites, especially LAt (average SD = 142 mg/m2, Figure 3.3b). LAt had significantly higher organic matter content than lat and LAT (ANOVA, F3,14 = 4.849, p < 0.05, Figure 3.3b). Sediment elemental composition did not vary significantly among study sites. Algal biomass, measured as chlorophyll-a was highly variable within sites and ranged from 13.0 (± 7.0, SD) to 36.3 (± 20, SD) mg/m2. Attached algal chlorophyll-a did not vary significantly across study sites (ANOVA, F3,14 = 2.014, p > 0.05, Figure

3.3c). However, algal biomass (which includes cyanobacteria) measured as fatty acid biomarkers had significantly higher values at LAt (0.25 ± 0.1 µg fatty acid/mg dry mass,

F3,12 = 34.47, p < 0.001, Figure 3.4a). Even though fatty acid content varied, sediment energy content averaged 15060 j/g organic matter and did not vary significantly across study sites (F3,13 = 0.7275, p > 0.05, Table 3.1). Bacterial biomass ranged between 0.11 and 0.80 (± 0.01 to 0.05, SD) µg fatty acid/mg dry mass and was significantly higher at

LAt than at the other three sites (F3,12 = 31.61, p < 0.001, Figure 3.4b). Algal biomass expressed as per mole fraction was lower at LAT compared to the other three sites, although this difference was not significant (Figure 3.4c). Bacterial biomass expressed as per mole fraction was significantly lower at LAT than at the other three sites (Figure

3.4d).

Microbial diversity The average number of reads per sample was 55,157, with a minimum number of

34, 014 at LAT site and the maximum number of 83,559 at the LAt site. The number of

OTUs decreased from upstream (20138) to downstream (8533). Alpha diversity,

67

expressed as Shannon’s H and Simpson’s E, was similar in stream sediments across the upstream to downstream gradient (Table 3.3). Habitat complexity, measured as sediment particle size distribution metrics, did not correlate with diversity. Bacterial OTU richness estimates were twice as high in the headwater site (LAt) compared to the site farthest downstream (LAT). The rank-abundance distributions showed a strong of a few OTUs and a long tail of rare OTUs across all sites (Figure 3.5a). The number of rare

OTUs was higher in the most downstream site, LAT.

Microbial community composition Biofilm community composition from all four sites were similar, but showed distinct patterns (Figure 3.5b). Bray-Curtis Dissimilarity values were all below 0.3 indicating that all sites were similar in composition. Nevertheless, LAt was most dissimilar to laT (0.20) and LAT (0.21). In all samples, the phylum Proteobacteria represented the most abundant group, on average 53.4%. Within this phylum, the classes

Beta- (average 21.5%) and Deltaproteobacteria (average 10.4%) were the two most frequently observed. Alphaproteobacteria accounted for 8.5%, while

Gammaproteobacteria accounted for 7.1% of the microbial community abundance. The other Proteobacteria (including Epsilon-, Zeta-, and TA18) accounted for 5.7%. The relative abundance of the phylum Proteobacteria varied across study sites.

Betaproteobacteria was highest at laT (33.8%) and lowest at LAT (11.1%).

Deltaproteobacteria and TA18 was highest in the LAT site at 17.1 and 10.6%, respectively. TA18 was much lower at the 3 other sites which averaged 0.18%.

Aside from Proteobacteria, the phyla Bacteroidetes, Planctomycetes, and

Cyanobacteria were among the most abundant phyla at each site. The second most

68

abundant phylum was Bacteroidetes (average 9.2%). The highest relative abundance of

Bacteroidetes occurred at LAt (12.9%) and the lowest was at laT (5.7%). Planctomycetes accounted for 5.2%, followed in importance by Cyanobacteria (average 5.4%). The highest abundance of Cyanobacteria was found at LAt, which had 11.9% while the other

3 sites averaged 3.2%. Other abundant phyla included Acidobacteria and Chloroflexi

(averaged 3.8 and 3.9%, respectively), Verrucomicrobia (average 3.4%), Actinobacteria

(average 3.2%), Gemmatimonadetes (3.2%), and Nitrospirae (average 1.7%). The average percentage of “Unassigned and Other sequences” was 3.1%.

PCA analyses of the relative abundances revealed clear differences in sediment biofilm communities among sites (Figure 3.6). PC1 explained 53% of the variation across sites and separated the most upstream site, LAt from the most downstream site, LAT.

PC1 suggested increased Deltaproteobacteria,Chloroflexi and decreased

Alphaproteobacteria relative abundance going downstream. PC2 explained 33% of the variation across sites and separated laT and lat from LAt and LAT. The phyla

Bacteroidetes and Cyanobacteria loaded closely to LAt. Betaproteobacteria loaded closely to lat and laT suggesting lower relative abundance along PC2, while

Planctomycetes, and the Proteobacteria classes TA18, Epsilon- and Deltaproteobacteria increased at LAT along PC2.

CCA revealed significant environmental drivers of microbial community composition (Figure 3.7). CCA models were run on pairs of environmental variables that significantly varied between sites to avoid overfitting the model. Local physical and chemical variables explained 78% (CCA1) of the variation in microbial community composition with light intensity and SRP as significant variables (p < 0.05). Contrary to

69

our expectation, sediment particle size distribution, algal biomass (as chlorophyll-a and fatty acid biomarker), and total organic matter (as AFDM) did not have a significant effect on microbial community composition.

Discussion Small streams are intricately linked to the surrounding landscape causing them to be especially sensitive to changes in land use, however, there is little information on the effects of agricultural practices on benthic microbial community composition in lotic ecosystems (Findlay, 2010). Biofilm microbial assemblages drive biogeochemical cycling in stream ecosystems. By analyzing the sediments along a longitudinal stretch of the Little Miami River, we demonstrated that microbial community composition was correlated with local environmental conditions (e.g. light, SRP), which were directly influenced by surrounding land use, suggesting that both land use practices and local stream properties influence sediment microbial communities in streams.

The impacts of agricultural practices on physicochemical variables were evident but did not vary along an upstream-downstream gradient. Bacterial decreased from upstream to downstream (OTUs, Table 3.3). Despite variation in physicochemical variables, bacterial community diversity and evenness did not differ from upstream to downstream. The most impacted site in terms of riparian forested vegetation removal and nutrient concentration had the highest species richness. This result is inconsistent with studies of metazoans which often observe a decline in species richness under nutrient enrichment (Suding et al. 2005). Other studies have shown an increase in heterotrophic microbial species richness with inorganic and organic nutrients associated with anthropogenic activities (Wakelin et al. 2008). The similarity in evenness

70

and the relatively similar community composition across sites (Bray-Curtis) implies that either the bacteria observed are generalists, capable of using a variety of carbon sources

(Wittebolle et al. 2009), or there is a high level of dormancy among these sediment- associated communities. Low-order streams are dynamic environments that experience temporal changes in physical conditions. A highly dynamic environment selects for dormancy, which can affect species richness because dormant individuals are included in microbial assessments of 16SrDNA (Lennon and Jones 2011). The ability to enter a dormant state is a common response to unfavorable conditions in microbial communities, such that 20-80% of bacteria recovered from environmental samples appear to be metabolically inactive (Cole 1999).

When autotrophic production increases in response to increased nutrients, bacteria benefit from an increased availability of labile C (Carr et al. 2005). The energy content of sediments did not vary significantly among sites. However, total organic matter and algal biomass were highest at the most impacted site where both heterotrophic biomass and richness was highest (Figure 3.3). When autotrophic carbon is plentiful, heterotrophic organisms depend less on terrestrial and detrital organic matter (Findlay et al. 1993,

Romani 2004a) largely due to the labile nature of algal exudates (Ylla et al. 2009).

Higher availability of algal exudates at the headwater site might explain why we saw greater richness and biomass of heterotrophs.

Headwaters are critical reservoirs for microbial richness in lotic ecosystems

(Besemer et al. 2013), which contradicts predictions by the river continuum concept and patterns observed with macro-organisms, which increase in richness from headwaters to downstream (Vannote et al. 1980, Finn et al. 2011, Kiffney et al. 2006). Many

71

assumptions to the longitudinal expectation of diversity in streams relies on the dispersal of upstream community members to downstream as well as the input from increased watershed drainage area (Wilson 1992, Economo and Keitt, 2010). Although some bacterial taxa may be dispersal limited, the majority are widely distributed and are strongly structured by local environmental conditions (Van der Gucht et al. 2007).

Furthermore, freshwater and terrestrial taxonomic profiles often overlap supporting the idea that terrestrial systems act as a source of microbial diversity (Tamames et al. 2010).

Given that overall community compositions were similar across sites, it is possible that dispersal played a role in our study even though overall richness decreased downstream.

Microbial community structure and diversity correlate significantly with pH in soils and sediments across acidic and alkaline conditions (Rouske et al. 2010, Xiong et al.

2012). Water column pH in this study spanned less than 1 pH unit and was not correlated with richness or individual phyla. Richness and diversity of microbial communities is highest at or near-neutral pH (Lauber et al. 2009), which might contribute to the high

OTUs observed here. Although pH has been shown to be a main contolling factor of microbial community structure, individual species can grow in environments spanning 4 pH units (Rosso et al. 1995), thus pH is likely not a contributing factor explaining the variation in microbial communtities in this study.

The majority of the abundant community members associated with sediments in our study represented the same phyla (and classes) that dominate freshwater bacterial biofilms growing on hard substrates (Besemer et al. 2012, Pohlon et al. 2013, Battin et al.

2016). Many of these abundant phyla have been characterized in terms of their ecological role in soils and sediments (Table 3.1). This allows us to draw inferences about

72

environmental conditions and community composition across our sites, even at the coarseness of phyla and class. Sediment microbial communities from each site were dominated by Proteobacteria, which is a ubiquitous and metabolically diverse phylum often found in freshwater sediments (Methe et al. 1998, Fazi et al. 2005). Within

Proteobacteria, the most highly abundant class Betaproteobacteria is comprised of mostly aerobic bacteria and is often found in organic-rich soils and eutrophic sediments.

Betaproteobacteria had the greatest relative abundance at the site with the greatest terrestrial input (laT) and was lowest at the least impacted site (LAT). The predominately aerobic Deltaproteobacteria had the greatest relative abundance at the two sites with potentially high terrestrial organic carbon input which is comprised of complex C molecules. This class is known for being capable of oxidizing a range of complex polymeric carbon compounds (Couradeau et al. 2011, Robinson et al. 2016). We did not measure interstitial oxygen but the recently described TA18 was drastically higher at

LAT and previous research has only found this subclass of Proteobacteria in oxic sediments (Robinson et al. 2016).

Bacteriodetes was the second-most abundant phyla and was most abundant at the headwater site. Members of Bacteriodetes are adept at breaking down large complex molecules (Robinson et al. 2016). This group is often a dominant member of sediment microbial communities in marine and freshwater sites and is characteristic of well- developed stream biofilms (Fang et al. 2017, Besemer et al. 2009). Additionally, this phylum is typically present in the gastrointestinal tract of mammals and in high-nutrient environments (Mao et al. 2015, Zeglin 2015). The high relative abundance of

Bacteroidetes at the two sites with the smallest riparian buffer zone might reflect

73

increased influence of animal manure through runoff from the adjacent farm fields.

Furthermore, the class Sphingobacteria was the primary driver of the high relative abundance of Bacteroidetes phylum at LAt. Sphingobacteria is known for its affinity for nutrient-rich environments. The high relative abundance of Cyanobacteria at the headwater site (LAt) makes sense because they are that do well under high light and high nutrient concentrations (Wagenhoff et al. 2013). Cyanobacteria are one of the most important autotrophs in streams (Power et al. 1985, Lamberti and Steinman,

1997). Furthermore, both Bacteroidetes and Cyanobacteria loaded closely to the most impacted site (Figure 3.6). Planctomycetes are known for their ability to break down a wide variety of C substrates and are typically found in eutrophic systems. Therefore, it is not surprising that Planctomycetes loaded between site LAT and LAt in the CCA. In summary, differences in the relative abundance of the dominant phyla and classes across longitudinal gradients may be linked to changes from anthropogenic land uses.

The overall similarities of microbial community composition among the upstream and downstream sites likely reflects OTU-specific tolerance to local conditions

(Jyrkankallio-Mikkola et al. 2017) and the linkage between landscape and freshwater microbial communities (Veach et al. 2016, Battin et al. 2016). Our results underline the role of water chemistry (SRP) in structuring aquatic microbial communities (Soininen et al. 2004) but also give support to the inclusion of physical variables (light) into shaping communities (Jyrkankallio-Mikkola et al. 2017). However, the longitudinal connectivity within a stream makes it difficult to determine if differences in relative abundance of community members are due to site-specific characteristics that differ longitudinally within the study streams or due to dispersal limitation, or species sorting (Besemer et al.

74

2013).The assembly of local communities can be explained by metacommunity theory, whereby local processes, including the abiotic environment and biotic interactions, and regional processes, such as dispersal, regulate the formation of entire communities

(Besemer et al. 2012). On a local scale, the interplay of niche availability and has been suggested to drive the patterns of bacterial biodiversity in biofilms through species sorting. Differences in physical and chemical characteristics of our stream suggests that land use practices act as a selective force on sediment associated microbial communities in streams (Roberto et al. 2018).

In general, unexplained variation in microbial community composition remains high regardless of the number of explanatory variables (Meier et al. 2015, Jyrkankallio-

Mikkola et al. 2017). Nevertheless, we did observe significant relationships between our measured physical and chemical variables and sediment microbial communities. Light and SRP explained a large proportion of the variation in community composition and this was driven by the community observed at the headwater site LAt (Figure 3.7). Microbial community composition was not predicted by any of the metrics describing carbon characteristics that we measured. Others have shown that spatial and temporal variation in community composition and activity of sediment-associated microbial communities is linked to the size distribution and composition of both particulate and dissolved organic carbon inputs (Sinsabaugh and Findlay 1995, Battin et al. 2008, Sobczak and Findlay

2002). Stream sediment microbial communities form and exist in a complex and dynamic environment. Without accounting for all sources of available carbon we have an unclear picture of the realized organic carbon pool in this stream. Supply of dissolved organic

75

carbon can alter community composition and determine how hyporheic sediment microbial communities are structured (Findlay et al. 2003).

Indirect effects of watershed land use affect physical-chemical characteristics of streams, which in turn, influence sediment microbial community composition. These microbial communities are key drivers of carbon and nutrient cycles and serve as a carbon source to consumers in streams. Changing environmental factors at the watershed scale directly impact metazoan communities (Hynes 1975, Mulholland et al. 2008), and we have provided evidence that environmental variation in nutrients and light can explain variation in microbial community assemblages. Nutrient impacts on stream microbiota from land use change are often accompanied by differences in riparian vegetation and cover and thus carbon availability (Findlay and Sinsabaugh, 2006). Given the ubiquity of agricultural land use in the Midwest, these results raise questions about the effects of human activity on stream ecosystems.

Literature Cited APHA 2005. Standard Methods for the Examination of Water and Wastewater, 21st edn.

Total organic carbon (TOC): high temperature combustion method (5310 A and

5310 B).

Arar, E. J., & Collins, G. B. (1997). Method 445.0: In vitro determination of chlorophyll

a and pheophytin a in marine and freshwater algae by fluorescence. Washington,

DC, USA: United States Environmental Protection Agency, Office of Research

and Development, National Exposure Research Laboratory.

76

Battin, T. J., Besemer, K., Bengtsson, M. M., Romani, A. M., & Packmann, A. I. (2016).

The ecology and biogeochemistry of stream biofilms. Nature Reviews

Microbiology, 14(4), 251–263.

Battin, T. J., Kaplan, L. A., Findlay, S., Hopkinson, C. S., Marti, E., Packman, A. I.,

Newbold, J. D. & Sabater, F. (2008). Biophysical controls on organic carbon

fluxes in fluvial networks. Nature Geoscience, 1(2), 95–100.

Battin, T. J., Kaplan, L. A., Newbold, J. D., Cheng, X., & Hansen, C. (2003). Effects of

current velocity on the nascent architecture of stream microbial biofilms. Applied

and Environmental Microbiology, 69(9), 5443–5452.

Besemer, K., Hödl, I., Singer, G., & Battin, T. J. (2009). Architectural differentiation

reflects bacterial community structure in stream biofilms. The ISME Journal,

3(11), 1318–1324.

Besemer, K., Peter, H., Logue, J. B., Langenheder, S., Lindström, E. S., Tranvik, L. J., &

Battin, T. J. (2012). Unraveling assembly of stream biofilm communities. The

ISME Journal, 6(8), 1459–1468.

Besemer, K., Singer, G., Quince, C., Bertuzzo, E., Sloan, W., & Battin, T. J. (2013).

Headwaters are critical reservoirs of microbial diversity for fluvial networks.

Proceedings of the Royal Society of London B: Biological Sciences, 280(1771),

20131760.

Brett, M. T., Bunn, S. E., Chandra, S., Galloway, A. W. E., Guo, F., Kainz, M. J., …

Wehr, J. D. (2017). How important are terrestrial organic carbon inputs for

secondary production in freshwater ecosystems? Freshwater Biology, 62(5), 833–

853.

77

Carr, G. M., Morin, A., & Chambers, P. A. (2005). Bacteria and algae in stream

periphyton along a nutrient gradient. Freshwater Biology, 50(8), 1337–1350.

Cole, J. J. (1999). Aquatic microbiology for ecosystem scientists: new and recycled

paradigms in ecological microbiology. Ecosystems 2, 215–225.

Collins, S. M., Kohler, T. J., Thomas, S. A., Fetzer, W. W., & Flecker, A. S. (2016). The

importance of terrestrial subsidies in stream food webs varies along a stream size

gradient. Oikos, 125(5), 674–685.

Couradeau, E., Benzerara, K., Moreira, D., Gérard, E., Kaźmierczak, J., Tavera, R., &

López-García, P. (2011). Prokaryotic and eukaryotic community structure in field

and cultured microbialites from the alkaline Lake Alchichica (Mexico). PLOS

ONE, 6(12), e28767.

Dale, N. G. (1974). Bacteria in intertidal sediments: Factors related to their distribution.

Limnology and Oceanography, 19(3), 509–518.

DeForest, J. L., Smemo, K. A., Burke, D. J., Elliott, H. L., & Becker, J. C. (2012). Soil

microbial responses to elevated phosphorus and pH in acidic temperate deciduous

forests. Biogeochemistry, 109(1-3), 189–202.

Economo, E. P., & Keitt, T. H. (2010). Network isolation and local diversity in neutral

metacommunities. Oikos, 119(8), 1355–1363.

Fabian, J., Zlatanovic, S., Mutz, M., & Premke, K. (2017). Fungal–bacterial dynamics

and their contribution to terrigenous carbon turnover in relation to organic matter

quality. The ISME Journal, 11(2), 415–425.

78

Fang, H., Chen, Y., Huang, L., & He, G. (2017). Analysis of biofilm bacterial

communities under different shear stresses using size-fractionated sediment.

Scientific Reports, 7(1), 1299.

Fazi, S., Amalfitano, S., Pernthaler, J., & Puddu, A. (2005). Bacterial communities

associated with benthic organic matter in headwater stream microhabitats.

Environmental Microbiology, 7(10), 1633–1640.

Findlay, S. (2010). Stream microbial ecology. Journal of the North American

Benthololgical Society. 29, 170–181.

Findlay, S. E. G., Sinsabaugh, R. L., Sobczak, W. V., & Hoostal, M. (2003). Metabolic

and structural response of hyporheic microbial communities to variations in

supply of dissolved organic matter. Limnology and Oceanography, 48(4), 1608–

1617.

Findlay, S., Howe, K., & Fontvielle, D. (1993). Bacterial-algal relationships in streams of

the Hubbard Brook Experimental Forest. Ecology, 74(8), 2326–2336.

Findlay, S., & Sinsabaugh, R. L. (2006). Large-scale variation in subsurface stream

biofilms: A cross-regional comparison of metabolic function and community

similarity. Microbial Ecology, 52(3), 491–500.

França, J. S., Gregório, R. S., Paula, J. D. de, Júnior, J. F. G., Ferreira, F. A., & Callisto,

M. (2009). Composition and dynamics of allochthonous organic matter inputs and

benthic stock in a Brazilian stream. Marine and Freshwater Research, 60(10),

990–998.

79

Francoeur, S. N., Winslow, K.A.P., Miller, D., & Peacor, S. D. (2017). Mussel-derived

stimulation of benthic filamentous algae: The importance of nutrients and spatial

scale. Journal of Great Lakes Research, 43(1), 69–79.

Goodman, K. J., Parker, S. M., Edmonds, J. W., & Zeglin, L. H. (2015). Expanding the

scale of aquatic sciences: the role of the National Ecological Observatory

Network (NEON). Freshwater Science, 34(1), 377–385.

Gucht, K. V. der, Cottenie, K., Muylaert, K., Vloemans, N., Cousin, S., Declerck, S.,

Jeppesen, E., Conde-Porcuna, J., Schwenk, K., Zwart, G., Degans, H., Vyverman,

W., & Meester, L. D. (2007). The power of species sorting: Local factors drive

bacterial community composition over a wide range of spatial scales. Proceedings

of the National Academy of Sciences, 104(51), 20404–20409.

Harrington, S. (1999). Occurrence and distribution of fish species in the Great and Little

Miami River Basins, Ohio and Indiana, pre-1900 to 1998. US Department of the

Interior, US Geological Survey National Water-Quality Assessment Program.

Hiriart‐Baer, V. P., Arciszewski, T. J., Malkin, S. Y., Guildford, S. J., & Hecky, R. E.

(2008). Use of pulse-amplitude-modulated fluorescence to assess the

physiological status of Cladophora sp. along a water quality gradient. Journal of

Phycology, 44(6), 1604–1613.

Hynes, H. B. N. (1975). The stream and its valley. Internationale Vereinigung für

theoretische und angewandte Limnologie: Verhandlungen, 19(1), 1-15.

Jones, S. E., Newton, R. J., & McMahon, K. D. (2009). Evidence for structuring of

bacterial community composition by organic carbon source in temperate lakes.

Environmental Microbiology, 11(9), 2463–2472.

80

Jyrkänkallio-Mikkola, J., Meier, S., Heino, J., Laamanen, T., Pajunen, V., Tolonen, K. T.,

Tolkkinen, M., & Soininen, J. (2017). Disentangling multi-scale environmental

effects on stream microbial communities. Journal of , 44(7), 1512–

1523.

Kiffney, P. M., Greene, C. M., Hall, J. E., & Davies, J. R. (2006). Tributary streams

create spatial discontinuities in habitat, biological productivity, and diversity in

mainstem rivers. Canadian Journal of Fisheries and Aquatic Sciences, 63(11),

2518–2530.

Kirk, J. T. O. (1994). Light and Photosynthesis in Aquatic Ecosystems. Cambridge

University Press.

Lauber, C. L., Hamady, M., Knight, R., & Fierer, N. (2009). Pyrosequencing-based

assessment of soil pH as a predictor of soil bacterial community structure at the

continental scale. Applied and Environmental Microbiology, 75(15), 5111–5120.

Lennon, J. T., & Jones, S. E. (2011). Microbial seed banks: the ecological and

evolutionary implications of dormancy. Nature Reviews Microbiology, 9(2), 119–

130.

Lisboa, L. K., Silva, A. L. L. da, Siegloch, A. E., Júnior, J. F. G., & Petrucio, M. M.

(2015). Temporal dynamics of allochthonous coarse particulate organic matter in

a subtropical Atlantic rainforest Brazilian stream. Marine and Freshwater

Research, 66(8), 674–680.

Manning, P., Gossner, M. M., Bossdorf, O., Allan, E., Zhang, Y.-Y., Prati, D., Bluthgen,

N., Bock, S., Bohm, S., Borschig, C., Holzel, N., Jung, K., Klaus, V. H., Klein, A.

M., Kleinebecker, T., Krauss, J., Lange, M., Muller, J., Pasalic, E., Socher, S. A.,

81

Tschapka, M., Turke, M., Weiner, C., Werner, M., Gockel, S., Hemp, A., Renner,

S.C., Wells, K., Buscot, F., Kalka, E. K. V., Linsenmair, K. E., Weisser, W. W.,

& Fischer, M. (2016). Grassland management intensification weakens the

associations among the diversities of multiple plant and animal taxa. Ecology,

96(6), 1492–1501.

Meier, S., Luoto, M., & Soininen, J. (2015). The effects of local, buffer zone and

geographical variables on lake metacommunities. Hydrobiologia,

743(1), 175–188.

Methé, B. A., Hiorns, W. D., & Zehr, J. P. (1998). Contrasts between marine and

freshwater bacterial community composition: Analyses of communities in Lake

George and six other Adirondack lakes. Limnology and Oceanography, 43(2),

368–374.

Middelboe, A. L., & Markager, S. (2003). Depth limits and minimum light requirements

of freshwater macrophytes. Freshwater Biology, 37(3), 553–568.

Mulholland, P. J., Helton, A. M., Poole, G. C., Hall, R. O., Hamilton, S. K., Peterson, B.

J., Tank, J. L., Askenas, L. R., Cooper, L. W., Dahm, C. N., Dodds, W. K.,

Findlay, S. E. G., Gregory, S. V., Grimm, N. B., Johnson, S. L., McDowell, W.

H., Meyer, J. L., Valett, H. M., Webster, J. R., Arango, C. P., Beaulieu, J. J.,

Bernot, M. J., Burgin, A. J., Crenshaw, C. L., Johnson, L. T., Niederlehner, B. R.,

O’Brien, J. M., Potter, J. D., Sheibley, R. W., Sobota, D. J. & Thomas, S. M.

(2008). Stream denitrification across and its response to anthropogenic

nitrate loading. Nature, 452(7184), 202–205.

82

OEPA 1995. Biological and Water Quality Study of The Little Miami River and Selected

Tributaries. OEPA Technical Report MAS/1994-12-11. Ohio EPA Division of

Surface Water. Columbus, OH.

Osborne, L. L., & Kovacic, D. A. (1993). Riparian vegetated buffer strips in water-

quality restoration and stream management. Freshwater Biology, 29(2), 243–258.

Pohlon, E., Fandino, A. O., & Marxsen, J. (2013). Bacterial community composition and

extracellular enzyme activity in temperate streambed sediment during drying and

rewetting. PLOS ONE, 8(12), e83365.

Power, M. E., Stewart, A. J., & Matthews, W. J. (1988). Grazer control of algae in an

Ozark Mountain stream: Effects of short-term exclusion. Ecology, 69(6), 1894–

1898.

R Core Team. (2015). R: A Language and Environment for Statistical Computing, R

Foundation for Statistical Computing, Vienna, Austria.

Roberto, A. A., Van Gray, J. B., & Leff, L. G. (2018). Sediment bacteria in an urban

stream: Spatiotemporal patterns in community composition. Water Research, 134,

353–369.

Robinson, G., Caldwell, G. S., Wade, M. J., Free, A., Jones, C. L. W., & Stead, S. M.

(2016). Profiling bacterial communities associated with sediment-based

aquaculture systems under contrasting redox regimes. Scientific

Reports, 6, 38850.

Romaní, A. M., Guasch, H., Muñoz, I., Ruana, J., Vilalta, E., Schwartz, T., Emtiazi, F. &

Sabater, S. (2004). Biofilm structure and function and possible implications for

riverine DOC dynamics. Microbial Ecology, 47(4), 316–328.

83

Rosemond, A. D., Benstead, J. P., Bumpers, P. M., Gulis, V., Kominoski, J. S., Manning,

D. W. P., Suberkropp, K. & Wallace, J. B. (2015). Experimental nutrient

additions accelerate terrestrial carbon loss from stream ecosystems. Science,

347(6226), 1142–1145.

Rosso, L., Lobry, J.R., Bajard, S., & Flandrois, J.P. (1995). Convenient model to describe

the combined effects of temperature and pH on microbial growth. Applied

Environmental Microbiology, 61: 610–616.

Rousk, J., Baath, E., Brookes, P.C., Lauber, C.L., Lozupone, C., Caporaso, J.G., Knight,

R., & Fierer, N. (2010). Soil bacterial and fungal communities across a pH

gradient in an arable soil. ISME Journal, 4: 1340–1351.

Sabater, S., Guasch i Padró, H., Muñoz Gràcia, I., & Romaní i Cornet, A. M. (2006).

Hydrology, light and the use of organic and inorganic materials as structuring

factors of biological communities in Mediterranean streams. Limnetica, 25(1-2),

335-348.

Shade, A., Peter, H., Allison, S. D., Baho, D., Berga, M., Buergmann, H., Huber, D. H.,

Langenheder, S., Lennon, J. T., Martiny, J. B.H., Matulich, K. L., Schmidt, T. M.

& Handelsman, J. (2012). Fundamentals of microbial community resistance and

resilience. Frontiers in Microbiology, 3, 417.

Shivers, S. D., Opsahl, S. P., & Covich, A. P. (2016). Microbial bioavailability of

dissolved organic carbon from leachates of freshwater autotrophs. Aquatic

Microbial Ecology, 76(3), 233–241.

84

Sinsabaugh, R. L., & Findlay, S. (1995). Microbial production, enzyme activity, and

carbon turnover in surface sediments of the Hudson River . Microbial

Ecology, 30(2), 127–141.

Sobczak, W. V., & Findlay, S. (2002). Variation in bioavailability of dissolved organic

carbon among stream hyporheic flowpaths. Ecology, 83(11), 3194–3209.

Soininen, J., & Könönen, K. (2004). Comparative study of monitoring South-Finnish

rivers and streams using macroinvertebrate and benthic diatom community

structure. Aquatic Ecology, 38(1), 63–75.

Stainton, M. P. (1980). Errors in molybdenum blue methods for determining

orthophosphate in freshwater. Canadian Journal of Fisheries and Aquatic

Sciences, 37(3), 472–478.

Suding, K. N., Collins, S. L., Gough, L., Clark, C., Cleland, E. E., Gross, K. L., ... &

Pennings, S. (2005). Functional-and abundance-based mechanisms explain

diversity loss due to N fertilization. Proceedings of the National Academy of

Sciences, 102(12), 4387-4392.

Tamames, J., Abellán, J. J., Pignatelli, M., Camacho, A., & Moya, A. (2010).

Environmental distribution of prokaryotic taxa. BMC Microbiology, 10, 85.

Vannote, R. L., Minshall, G. W., Cummins, K. W., Sedell, J. R., & Cushing, C. E.

(1980). The river continuum concept. Canadian Journal of Fisheries and Aquatic

Sciences, 37(1), 130–137.

Veach, A. M., Stegen, J. C., Brown, S. P., Dodds, W. K., & Jumpponen, A. (2016).

Spatial and successional dynamics of microbial biofilm communities in a

grassland stream ecosystem. Molecular Ecology, 25(18), 4674–4688.

85

Wagenhoff, A., Lange, K., Townsend, C. R., & Matthaei, C. D. (2013). Patterns of

benthic algae and cyanobacteria along twin-stressor gradients of nutrients and fine

sediment: a stream mesocosm experiment. Freshwater Biology, 58(9), 1849–

1863.

Wakelin, S. A., Colloff, M. J., & Kookana, R. S. (2008). Effect of wastewater treatment

plant effluent on microbial function and community structure in the sediment of a

freshwater stream with variable seasonal flow. Applied and Environmental

Microbiology, 74(9), 2659–2668.

Wallace, J. B., Eggert, S. L., Meyer, J. L., & Webster, J. R. (1999). Effects of resource

limitation on a detrital-based ecosystem. Ecological Monographs, 69(4), 409–

442.

Wallace, J. B., & Webster, J. R. (1996). The role of macroinvertebrates in stream

ecosystem function. Annual Review of Entomology, 41(1), 115–139.

Wang, Y., Sheng, H.-F., He, Y., Wu, J.-Y., Jiang, Y.-X., Tam, N. F.-Y., & Zhou, H.-W.

(2012). Comparison of the levels of bacterial diversity in freshwater, intertidal

wetland, and marine sediments by using millions of Illumina Tags. Applied and

Environmental Microbiology, 78(23), 8264–8271.

Warren, D. R., Collins, S. M., Purvis, E. M., Kaylor, M. J., & Bechtold, H. A. (2017).

Spatial variability in light yields colimitation of primary production by both light

and nutrients in a forested stream ecosystem. Ecosystems, 20(1), 198–210.

Wetzel, R. G. (2001). Limnology: Lake and River Ecosystems, 3rd. Acadimic pres, AQn

Elsevier imprint, Sanfrancisco, New York, London.

86

White, D. C., Davis, W. M., Nickels, J. S., King, J. D., & Bobbie, R. J. (1979).

Determination of the sedimentary microbial biomass by extractible lipid

phosphate. Oecologia, 40(1), 51–62.

Winslow, K. P., Francoeur, S. N., & Peacor, S. D. (2014). The influence of light and

nutrients on benthic filamentous algal growth: A case study of Saginaw Bay, Lake

Huron. Journal of Great Lakes Research, 40, 64–74.

Wittebolle, L., Marzorati, M., Clement, L., Balloi, A., Daffonchio, D., Heylen, K., Vos,

P.D., Verstraete, W. & Boon, N. (2009). Initial community evenness favours

functionality under selective stress. Nature, 458(7238), 623–626.

Xiong, J., Liu, Y., Lin, X., Zhang, H., Zeng, J., Hou, J., Yang, Y., Yao, T., Knight, R., &

Chu, H. (2012). Geographic distance and pH drive bacterial distribution in

alkaline lake sediments across Tibetan Plateau. Environmental Microbiology,

14(9), 2457–2466.

Ylla, I., Borrego, C., Romaní, A. M., & Sabater, S. (2009). Availability of glucose and

light modulates the structure and function of a microbial biofilm. FEMS

Microbiology Ecology, 69(1), 27–42.

Zeglin, L. H. (2015). Stream microbial diversity in response to environmental changes:

review and synthesis of existing research. Frontiers in Microbiology, 6, 454.

87

Table 3.1. Dominant microbial community members and their relevant ecological functions and habitat.

Relevant Ecological Phylum Class Metabolism Habitat Function Anaerobic and Complex carbon break Soils, sediments, guts Bacteroidetes aerobic down of animals Complex polysaccharide Abundant in animal Bacteroidia Obligate anaerobic break down guts Commensal and Marine and Flavobacteria pathogens of fish, do not freshwater degrade cellulose Nutrient-rich Sphingobacteria Degrade wide range of C environments, sediments, soils Heterotrophic, Key players in C, S, and Dominate many Proteobacteria phototrophic, N cycles environments chemotrophic Aerobic and N fixation associated anaerobic Alpha- with plants, autotrophic phototrophs and methane oxidation heterotrophs Breakdown phenol and Organic-rich soils, Heterotrophic, lignin, nitrification, Beta- sediments, eutrophic chemolithotrophic ammonia, S, and H aquatic systems oxidizers Obligate and facultative aerobic and anaerobic, Gamma- heterotrophic, Ubiquitous chemoautotrophic, pathogenic, symbiotic Estuary and marine Delta- Anaerobic Sulfate reducers sediments Intestines of Anaerobic, Sulfate reducers, do not mammals and birds, Epsilon- heterotrophic, process C dominate chemoautotrophic hydrothermal vents High oxygen TA18 Aerobic Newly described environments Photosynthesis, nitrogen Common in all Cyanobacteria Autotrophic fixation aquatic environments Aquatic Stramenopiles Phototrophic Photosynthesis environments Thermophilic Extreme Chloroflexi phototrophs, Photosynthesis environments, soils, heterotrophic sediments, freshwater Aerobic, Actinobacteria Wide substrate use Ubiquitous in soils heterotrophic Aerobic, Ammonia oxidation, use Aquatic and mesophilic, Planctomycetes wide range of C eutrophic oligotrophic, substrates environments heterotrophs Soils, sediments, Low C substrate Acidobacteria oligotrophic availability environments

88

Table 3.2. Physical characteristics, sediment structure, and energy content of sediments at the stream sites sampled. Sediment particle size metrics include the coefficient of gradation (Cc) and the uniformity coefficient (Cu). Energy content is reported in joules/g organic matter. Values in parentheses are SE. Site abbreviations: LAt = Dolly Varden, laT = Selma Pike, lat = Garlough, LAT = Grinnell.

Stream Stream Riparian Cond. Energy Site km width (m) width (m) pH (µS) CC CU Content 2.07 1.44 LAt 8.10 7 0 7.80 753 (0.24) (1.64) 14810 (464) 2.06 1.03 laT 11.56 13.6 427.9 6.93 742 (0.17) (0.85) 15070 (93.7) 1.88 1.04 lat 13.15 5.3 71.1 7.47 729 (0.15) (1.60) 14980 (245) 1.85 1.24 LAT 21.20 16.6 1734 7.71 672 (0.17) (1.37) 15340 (172)

89

Table 3.3. Diversity and evenness metrics of microbial communities analyzed from stream sediments.

Site Stream km Shannon’s H Simpson’s E OTUs Reads LAt 8.10 4.71 0.98 20138 83559 laT 11.56 4.98 0.95 13499 47153 lat 13.15 5.07 0.94 12657 55903 LAT 21.20 5.05 0.92 8533 34014

90

01 2 4 6

Kilometers

Flow Direction

Figure 3.1. Map of sampling locations along the Little Miami River and land use from the

National Land Cover Database (2011). Land use categories: crop agriculture (brown), pasture (yellow), forest (green), developed (red), water (blue).

91

Figure 3.2. Photographs of each sampling location. Upstream site (LAt), next site downstream (laT), site 3 (lat) and the furthest site downstream (LAT).

92

a b c b  2 m  L mg

a   g

 ab   b ab SRP a a OrganicMatter 0 001500 1000 500 0 1000 800 600 400 200 0

12 24 54 88 12 24 54 88 Percent Light Percent Light

c a  2 m a mg  

a a Chlorophyll-a 02 04 06 70 60 50 40 30 20 10 0

12 24 54 88 Percent Light

Figure 3.3. Soluble reactive phosphorus concentration in µg/L (a, SRP), sediment organic matter content in mg/m2 (b), and chlorophyll-a in mg/m2 (c) across percent light.

Reported p-values in the text are for the overall ANOVA. Within a single graph, boxes with the same letter above them are not significantly different (p < 0.05, Tukey’s HSD test). Horizontal lines represent median values and boxes represent inter-quartile ranges

(25-75% percentiles) and whiskers show the range.

93

a b b   c mgDM mgDM FA  g FA   g      ab a a

b a a Algal Biomass Algal Bacterial Biomass Bacterial . . . . 0.4 0.3 0.2 0.1 0.0 1.2 1.0 0.8 0.6 0.4 0.2 0.0

12 24 54 88 12 24 54 88 Percent Light Percent Light

c a d a

a ab b a a c Algal Biomass (per fraction) mole Biomass Algal Bacterial Biomass (per fraction) mole Biomass Bacterial 01 025 20 15 10 5 0 50 40 30 20 10 0

12 24 54 88 12 24 54 88 Percent Light Percent Light

Figure 3.4. Sediment fatty acid content among stream sites. Quantitative fatty acids are expressed as µg fatty acid per mg dry mass of sediment, not organic matter. Algal biomass includes cyanobacteria. Reported p-values in the text are for the overall

ANOVA. Within a single graph, boxes with the same letter above them are not significantly different (p < 0.05, Tukey’s HSD test). Horizontal lines represent median values and boxes represent inter-quartile ranges (25-75% percentiles) and whiskers show the range.

94

101 a

0 10 Relative Abundance (%)Abundance Relative

-2 10

LAt laT lat 10-3 LAT

0 200 400 600 800 1000

Rank

b 100

Acidobacteria Actinobacteria 75 Bacteroidetes Chloroflexi Cyanobacteria Gemmatimonadetes Nitrospirae Other 50 P-Alphaproteobacteria P-Betaproteobacteria P-Deltaproteobacteria P-Epsilonproteobacteria

Relative Abundance Relative P-Gammaproteobacteria P-TA18 25 Planctomycetes Unassigned Verrucomicrobia

0

laT lat LAT LAt

Figure 3.5. Rank-abundance curve of the microbial communities across the four sites (a) using relative abundance data. Community composition of the most abundant phyla or classes (for Proteobacteria) of the stream sediments expressed as relative abundances (b).

95

Cyano UnkBacter Plancto

EpsilonTA18

Delta

PC2 (31%)PC2 Alpha

Chloro

LAt Beta laT lat LAT 3- 10123 2 1 0 -1 -2 -3

-4 -2 0 2 4

PC1 (55%)

Figure 3.6. The first two axes of the principal component analysis (PC1, PC2) synthesizing sediment microbial community composition. Abbreviations: Cyano =

Cyanobacteria, Bacter = Bacteroidetes, Plancto = Planctomycetes, Unk = Unassigned,

Chloro = Chloroflexi, and Alpha, Beta, Delta, Epsilon, and Gamma = Proteobacteria classes.

96

pLight pSRP

Cyanobacteria

Bacteroidetes Unassigned Alphaproteobacteria Planctomycetes Verrucomicrobia ActinobacteriaAcidobacteria EpsilonproteobacteriaTA18 GammaproteobacteriaGemmatimonadetesDeltaproteobacteria CCA2 (18%) CCA2 BetaproteobacteriaNitrospiraeChloroflexiOther 10-. . . . 1.5 1.0 0.5 0.0 -0.5 -1.0

-1.0 -0.5 0.0 0.5 1.0 1.5 2.0

CCA1 (77%)

Figure 3.7. CCA of microbial relative abundances with environmental predictors soluble reactive phosphorus (SRP) and light availability (Light) for each site, LAt (closed circle),

LAT (closed triangle), lat (closed square), and laT (open triangle).

97

Chapter Four

LIGHT AVAILABILITY DETERMINES ATTACHED ALGAL PRODUCTIVY AND BENTHIC CONSUMER GROWTH AND PRODUCTION IN EXPERIMENTAL MESOCOSMS

Introduction Lake littoral zones have higher zoobenthic biomass and production than do profundal zones (Dermott 1978, Kajak 1978, Lindegaard 1992, Bergtold and

Traunspurger 2005) owing to changes in habitat and resources with depth (Brinkhurst

1974, Johansen 1974, Vadeboncoeur et al. 2011). Lower temperatures in the profundal zone lead to slower growth of profundal zoobenthos relative to their littoral counterparts

(Hamburger et al. 1994), while high habitat complexity is associated with higher in the littoral zone than the profundal zone (Strayer and Findlay 2010,

Vadeboncoeur et al. 2011). Additionally, the relative availability of low-quality detrital resources and high quality algal resources varies with depth due to declines in disturbance (Rowan and Rasmussen 1995) and light (Vadeboncoeur et al. 2003, 2014).

The relative importance of resource availability versus temperature in driving zoobenthic production patterns with depth is poorly resolved. Stable isotope analysis points to a strong reliance of zoobenthos on attached algae, a chiefly littoral resource (Vadeboncoeur et al. 2003, Vander Zanden et al. 2006, Devlin et al. 2013), but a direct, energetic dependence of benthic secondary production on attached algal production has rarely been quantified. I experimentally evaluated the bottom-up effect of light availability on

98

attached algal production, and the consequences of this variation for the production of a dominant constituent of the zoobenthic community, chironomid midges.

Temperature determines individual growth rates and total production of zoobenthic invertebrates in lakes (Brinkhurst 1974, Johansen 1974, Brey 2012). The hypolimnia of deep stratified temperate lakes can be 10-15°C colder than the epilimnia

(Wetzel 2001). These strong temperature gradients contribute to low profundal secondary production relative to the littoral zone (Brinkhurst 1974, Johansen 1974, Brey

2012). However, the mixed layers of stratified lakes are often nearly isothermal (Burns et al. 2005) making temperature alone insufficient to explain the decline in littoral zoobenthic production with depth (Lindegaard 1992, Babler et al. 2008, Butkas et al.

2011). Within lakes, resource availability can have a more obvious effect on zoobenthic biomass than temperature does. Terrestrial and phytoplankton detritus accumulate in areas of low disturbance. Thus, depositional areas support greater zoobenthic biomass than non-depositional areas, regardless of whether these depositional areas occur in the epilimnion or the hypolimnion (Rasmussen and Rowan 1997). Similarly, chironomid production corresponds to seasonal variation in phytoplankton biomass rather than strictly temperature (Rasmussen 1984). Thus, the quality and quantity of food resources are also a strong determinant of secondary production.

Benthic primary consumers can subsist on terrestrial detritus, phytoplankton, macrophytes, or attached algae. Nevertheless, the few studies that measure benthic primary and secondary production concurrently conclude that littoral zoobenthos rely substantially on attached algae (Strayer and Likens 1986, Vander Zanden et al. 2006,

Karlsson et al. 2009, Northington et al. 2010), and this is supported by stable isotope

99

analysis (Devlin et al. 2011). Attached algal productivity is strongly light limited in lakes, and beyond the zone of wave disturbance, production declines with depth (Vadeboncoeur et al. 2008, 2011, 2014, Malkin et al. 2010). Spatial variation in the productivity of this high quality resource may drive variation in littoral zoobenthic production.

I conducted a laboratory experiment to test the direct linkages between attached algal production and zoobenthic production. I used a light gradient to generate a gradient in primary production, and monitored the survival and growth of the generalist consumer

Chironomus dilutus across a gradient in algal productivity. I expected that attached algal primary production and biomass would increase in response to increased light availability. My primary objective was to assess the relationship between benthic primary production and benthic secondary production. If zoobenthos are primarily consuming benthic algae, then zoobenthic production will be coupled with attached algal production.

Methods Overview I conducted a laboratory experiment in which I manipulated light availability to generate a gradient of attached algal production on unconsolidated sediments. I cultured attached algae and eggs of Chironomus dilutus across a gradient of light intensities at 23°

C. Larval densities of C. dilutus were measured after 31 days of development and I collected all of the adults that emerged from each light treatment for 38 days after sampling the larvae. I tested whether the biomass and production of larval and adult chironomids was related to attached algal productivity, algal P content, and algal biomass

(as chlorophyll-a).

100

Experimental Setup I placed 4.6 cm of sandy sediment in the bottom of each of nine 35 L aquaria that had a total bottom area of 1290 cm2. Sediments were collected from a depth of 2 m in the littoral zone of Sparkling Lake, WI, USA, homogenized by hand, and air-dried. I filled each aquarium with artificial water made from deionized water, half-strength Chu #10 algal medium, and SeaChem lake salts (Chu 1942). Water was recirculated between the aquaria and a common 1,650 L reservoir. The reservoir was fitted with a chiller (Model

D1-100 chiller; Frigid Units, Inc) to maintain a uniform temperature of 23 °C, and a

Smart UV Water Sterilizer (Emperor Aquatics, Inc) to minimize phytoplankton growth. I placed a Thermocron ® iButton (DS1922T, Maxim-Integrated) at the sediment-water interface in each aquarium to record temperature at 10-minute intervals.

Immediately after set up, I exposed all 9 aquaria to similar light intensity (200 -

220 μmol m-2 s-1). Two days after starting the recirculating system, I inoculated all 9 aquaria with a mixed-algal slurry collected from rocks in the Little Miami River, OH.

After 14 days, robust biofilms had developed on the sediments, and I manipulated the lamps (Metal Halide, 1000W or 400W) to create 7 different light treatments (0 – 220

mol m-2s-1, Table 4.1). Two aquaria were set up at duplicate light intensities, one at high light and one at low light. I set the photoperiod to 12-h light:12-h dark.

I acquired C. dilutus egg sacs from a laboratory culture from the Environmental

Protection Agency (EPA Mid-continent Ecology Division, Duluth, MN). I used the ring- count method to quantify the number of eggs in each egg sac (Sadler 1935, Benoit et al.

2009). Briefly, for each egg mass I counted the mean number of eggs contained in five rings at approximately equal distances along the length of the tubular egg mass. I

101

multiplied the mean by the total number of rings in the egg mass to estimate the number of eggs. I inoculated 7 of the 9 aquaria at an average initial density of 2.0 ± 0.10 eggs cm-

2 (± SD), dividing egg sacs when necessary to approximate target densities. I estimated the initial mass of individual eggs based on egg volume of Chironomus islandicus

(diameter of 125 µm, length 340 µm; Lindegaard and Jonasson 1979) and a specific gravity of 1.05. I assumed that dry weight was 13% of wet weight (Jonasson 1974). Both the Sparkling Lake sediments and the Little Miami River algal slurry may have contained chironomid eggs. Therefore, two aquaria were not inoculated with C. dilutus, one at high light intensity (174.1 mol m-2 s-1) and one at low light intensity (23.2 mol m-2 s-1), so that I could assess the effect of these “wild” chironomid eggs.

Organic matter characterization On day 31 of the experiment, I collected four replicate sediment plugs using a 20 cc syringe corer and froze (-20 °C) the top 15 mm for chlorophyll-a, organic matter, and phosphorus analyses. I lyophilized, weighed, and homogenized each sample. A subsample (~ 20 mg) of each replicate was extracted for chlorophyll at 4°C for 24 hours in 90% buffered ethanol. I measured chlorophyll and phaeophyton fluorometrically with an acidification step (Arar and Collins 1997). I measured organic matter content as loss on ignition (LOI) at 500 °C for 1 h (APHA 2005). I measured P content by the molybdate method (Stainton et al. 1977).

I visually assessed biofilm thickness throughout the experiment. I took photographs through the sides of each aquaria weekly and measured the thickness on day

31 using a ruler.

102

Primary Productivity I used oxygen exchange methods to analyze benthic gross primary productivity

(GPP) and benthic net ecosystem productivity (NEP) (Vadeboncoeur et al. 2014). Clear and opaque acrylic cores were used to collect undisturbed sediments and we measured the evolution of oxygen over time (YSI Model Pro ODO). GPP was estimated using the net O2-exchange in clear acrylic cores plus the O2 consumption in paired dark cores.

Cores were incubated for 35 minutes but clear chambers in high light intensity treatments were terminated after 20 minutes to avoid air bubble accumulation. I used a photosynthetic quotient of 1.2 to convert primary productivity estimates in mg O2 to mg

C (Williams et al. 1979).

I used a diving pulse-amplitude modulated fluorometer (PAM) to assess primary production parameters in situ. The diving PAM measures relative electron transport rate

(ETR) of photosystem II as a function of light intensity using chlorophyll fluorescence. I used rapid light curves to derive ETRmax once per week over the course of the experiment

(Devlin et al. 2016). I measured ETRmax in 3 locations in each mesocosm between 11:00 and 14:00. Each rapid light curve consisted of exposing the attached algae to nine progressively increasing light intensities for 30 s each (1-900 µmol/m2/s). The fiberoptic probe was placed in a holder at a distance of 3 mm above the sediment-water interface.

Secondary Production I collected larvae after 31 days of growth using 5 cm diameter acrylic corers.

Each sediment core was sieved (250 and 1000 µm) and the retained larvae were collected and measured for head capsule width and body length using a dissecting microscope. I collected cores until I obtained at least 30 individuals per treatment for each of the 7

103

aquaria that received egg innocula (maximum of 6 cores). I limited larval collection because sampling was destructive and I wanted to leave enough area undisturbed to allow the remaining larvae to complete their life cycle. For the two treatments without an egg inoculum, I collected 4 cores and sieved the sediments for larvae. I did not find larvae in these two treatments. I used published length to dry weight regressions to convert head capsule width to biomass for chironomid larvae (Benke et al. 1999).

푑푟푦 푚푎푠푠 = 2.7842 × 퐿. Eq. (1) where, 퐿 is the head capsule width in mm and dry mass is in mg.

I collected adults as they emerged from each aquarium. I counted, pooled, dried

(24 h at 60°C), and weighed daily collections of adults emerging from each treatment to estimate average dry mass. I used the total number of flies emerged for adult density and average mass of all emerged adults for adult mass in each treatment. After 70 days, the two lowest light treatments had not produced adults. Therefore, I collected sediments and searched for larvae. I did not find any, and I ended the experiment.

I used the increment summation method to calculate production in each aquarium

(Rigler and Downing 1984). I first calculated specific growth rate (µ) for the time intervals for which I had individual mass (egg to larva at day 31) in each treatment using the equation:

( ) 휇 = Eq. 2 where, 푀 is the final mass, 푀 is the initial mass, and 푡 is the time elapsed in days.

Because I only had three estimates of mass (egg, larva at day 31, and emerged adults), I used 휇 from each aquarium to calculate mass in mg at a daily time-step using the following equation: 104

∗ 푀 = 푀 ∗ 푒 Eq. 3 where, 푀 is the estimated mass at time 푡, 푀is the mass at the beginning of the time interval (initial 푊= mass of egg), and 휇 is the estimated specific growth rate for a given aquarium.

I calculated mortality rates for each treatment in two ways. First, I calculated mortality rates between day 0 and day 31 using the equation:

( ) 푚 = Eq. 4

2 2 where, 휌 is the final density (individuals/m ), 휌 is the initial density (individuals/m ), and 푡 is the time in days for each time interval. I applied this mortality rate at a daily time step to get daily estimates of density for each aquaria between day 0 and 31.

To assess the validity of this method on overall production estimates, I varied initial (day 0-1) mortality rates to equal -0.99 and varied our estimated mortality rate by ±

10%. Additionally, I used survivorship curves and life tables from Lake Myvatn to back- calculate daily mortality rates from treatment-specific average mortality rates (calculated between day 0 and 31). I estimated daily mortality rates from the published survivorship curves to calculate daily mortality rate for each aquaria from the average mortality rate calculated in Eq. 4.

Using data from the survivorship growth curves, production tables for each treatment were constructed and daily estimates of production were calculated and summed using the following equation:

(휌 ∗ 퐼 ) 푃 = 푡

105

2 where, 푃 is the production (mg/m ) for an individual treatment, 휌 is the average

2 density (individuals/m ) between each increment, 퐼 is the dry weight gained (mg) in an increment, and 푡 is 1.

Larval production was converted to carbon units (mg C/m2) for comparison to attached algal productivity. I converted dry mass to ash-free dry mass (mg AFDM) and mean C content (mg C) using the conversion factor of 0.93 and 0.478, respectively

(Benke et al. 1999). I calculated trophic efficiency (퐸퐸) for each treatment as:

푃 퐸퐸 = ∗ 100 푃

2 where 푃 is larval production (mg C/m ), and 푃 is the gross ecosystem productivity of attached algae (mg C/m2/d) for each treatment.

Statistical Analysis I used general linear models to assess the effects of light availability on attached algal productivity, biomass, P content, and total biofilm organic matter content. To assess chironomid responses to attached algal primary productivity, I regressed larval production and biomass against primary productivity. I used general linear mixed effects models to assess whether daily temperatures varied across treatments. All statistical analyses were conducted using lme4 and nlme packages in R Version 3.2.3 (R Core

Team, 2015).

Results Biofilms on sediments The recirculating system maintained mean average daily temperatures in individual aquaria ranging from 22.8 – 23.0 °C (± 0.3, SD), and daily temperature did not vary significantly between treatments (t7, 19134 = 1.5697, p = 0.16). Light strongly affected

106

the thickness of the algal biofilms on the surface of the sediment. In the 2 lowest light treatments, no algal biofilms were visible. At light intensities between 20 and 50

mol/m2/s, algal biofilms were 1-2 mm thick while at the highest light intensities biofilms were up to 15 mm thick at day 31 of the experiment. Algal biomass measured as chlorophyll-a ranged from 0.12 to 26.4 (± 0.2 – 11.5 SE) mg/m2, and the light gradient generated a weak positive response in algal biomass (Figure 4.1a). The highest algal biomass occurred in the aquarium exposed to 45 mol/m2/s which developed a monoculture of the cyanobacterium Oscillatoria. Biofilm thickness and chlorophyll-a

2 were not significantly correlated (F1,7 = 2.899, p = 0.13, r = 0.19, Table 4.2). Surficial sediment organic matter content was not correlated with light intensity across

2 2 experimental aquaria (232 ± 23 g/m , F1,7 = 0.3211, p = 0.59, r = -0.09, Figure 4.1b).

Biofilm P content ranged from 5.4 to 13.6 (± 0.6 to 2.9) g/m2, and %P ranged from 0.12 to 0.24 % (± 0.03 to 0.14, Table 4.1). Percent P and P content of the biofilm did not vary

2 significantly across the light gradient (F1,7 = 0.2627, p = 0.62, r = -0.1, and F1,7 = 0.6377, p = 0.45, r2 = -0.04, Table 4.2).

Light intensity and primary production Gross primary productivity varied between 90 to 1260 mg C/m2/d. Respiration exceeded primary productivity for the two lowest light treatments (NEP = -193 and -169 mg C/m2/d; Figure 4.2). There were strong positive relationships between light

2 availability and primary productivity (GPP: F1,7 = 31.12, p < 0.01, r = 0.79, and NEP:

2 F1,7 = 27.52, p < 0.01, r = 0.79, respectively; Table 4.2). The two treatments without C. dilutus larva averaged 40 % lower GPP and 35% lower NEP than treatments containing

107

C. dilutus larvae at similar light intensities (Figure 4.2a). In situ estimates of ETRmax increased over time (Figure 4.2b).

Chironomid density and larval production Average individual mass of C. dilutus larvae on day 31 ranged from 0.5 – 6.9 mg across the light and primary productivity gradient (Figure 4.2). Benthic NEP was the best

2 predictor of average individual larval mass (F1,5 = 167.2, p < 0.001, r = 0.97, Figure

4.2a). Specific growth rate between egg and larva on day 31 was positively correlated with benthic NEP (Figure 4.2b) and was 1.6 times higher in the highest light treatment than the lowest light treatment. Furthermore, larvae reached adulthood 16 days earlier in high light treatments (Figure 4.2d) than low light treatments. The two lowest light treatments never produced adults after 70 days. Average mass of adults was 1.9 times greater in high light treatments than low light treatments (Figure 4.2f).

Larval production ranged from 5.8 to 110 mg C/m2/d (Figure 4.3a). Light intensity and benthic NEP were the best predictors of larval production (p < 0.001, r2 =

0.94 and p < 0.01, r2 = 0.84, Figure 4.3a). GPP explained 80% of the variation in larval production. The slope between NEP and larval production was 0.11. EE varied from -

0.03 to 0.28 in the treatments with C. dilutus but there was no relationship with light or primary productivity (Figure 4.3b).

To assess the validity of the crude estimates of larval production, I ran several scenarios varying mortality rates at the beginning of the growth curve. The greatest difference from the measured estimates of secondary production occurred when I used an initial mortality rate (between day 0 - 1) of 99% for all treatments. Larval production estimates were reduced by half in this scenario. Varying estimated mortality rates by ±

108

10% caused only slight variation in secondary production. Estimates of larval production using back-calculated daily mortality rates from Lake Myvatn data (Jónasson et al. 1990) were approximately 90% of the estimates obtained from my calculated average larval mortality rates.

Discussion Light availability explained most of the variation in biofilm GPP and NEP and zoobenthic production. Attached algal productivity was coupled with zoobenthic growth and production across all light treatments. The gradient in secondary production in the mesocosms mirrors the pattern of zoobenthic production often seen in lakes with depth.

Temperature did not significantly vary across the treatments and was not a driver of secondary production.

Light availability drove attached algal primary productivity in the mesocosms, similar to the relationships observed along depth gradients in lake littoral zones (Figure

4.1e). In littoral areas, primary production of attached algae on sediments is regulated by the light availability at the sediment surface (Karlsson et al. 2009, Vadeboncoeur et al.

2014). Benthic primary productivity across a gradient of light availability similar to our range in light intensities (between 10 – 60% of surface light) in Sparkling Lake varied between 100 – 750 mg m-2 d-1 (Vadeboncoeur et al. 2014). It is possible that primary productivity in the mesocosms changed over time throughout the experiment as biofilm accumulated. I measured primary production using O2-exchange methods only on day 31.

However, we used PAM fluorometry as an index of changes in primary productivity during the first 30 days of the experiment. The in situ estimates of production parameters indicated that primary productivity, measured as ETRmax, steadily increased over time

109

(Figure 4.1f). Thus, the measured primary productivity on day 31 overestimated the primary productivity that occurred over the course of the experiment and the estimates of are conservative.

Algal biomass, measured as chlorophyll-a, did not have a positive response to light availability, and was not significantly correlated with primary productivity (Figure

4.1b). The visual assessment of biofilms revealed dramatic differences in thickness across the light gradient but these changes were not evident in the assessment of algal biomass using chlorophyll-a. Although chlorophyll-a content is often used as a proxy for algal biomass, it is a poor index of algal biomass in attached biofilms when light intensity varies significantly (Baulch et al. 2009). In lakes, attached algal productivity and biomass are often decoupled across gradients in light availability (Vadeboncoeur et al. 2014). Nor was the total amount of organic carbon was a function of light. Organic carbon content of the sediments did not change significantly during the 31 days of the experiment despite the drastic change in productivity and visual accumulation of algal biomass. The percent organic matter averaged less than 5% across sites so even a doubling of organic matter in mesocosms would be difficult to discriminate using AFDM. The inability to detect changes in organic matter likely stems from the high inorganic content of the sediments.

Grazers can effectively remove algal biomass (Power 1988, Hillebrand 2009,

Rosemond et al. 1999) and it is possible that grazers suppressed the accumulation of algal biomass at high light intensities. I might not have observed a drastic increase in algal biomass in more productive aquaria because it was efficiently transferred into consumer biomass. Field studies verify that algal mats under grazed conditions do not accrue biomass because it is efficiently transferred into consumer tissue (Hill et al. 2010).

110

Furthermore, herbivores can efficiently track primary production over patch-level scales irrespective of algal biomass (Liess and Hillebrand 2004). Thus, attached algal production can exert control over consumer growth rates even when consumers have a strong effect on algal biomass (Steinman 1996).

At fine scales, tube-dwelling chironomids can generate a net positive effect on their resources (Herren et al. 2017). When light availability was comparable in the mesocosms, aquaria with chironomid larvae exhibited higher primary productivity than aquaria without larvae. Chironomid larvae can pump large volumes of water through the sediment that increases the flux of oxygen and nutrients (Holker et al. 2015, Roskosch et al. 2012). Consistent with the increase observed in this study, others have observed a

71% increase in primary productivity of attached algae in mesocosms with C. islandicus compared to treatments without larvae (Herren et al. 2017). The proposed mechanism behind this was the ability of these consumers to cycle the equivalent of the yearly external loading of N and P to the . Although I did not measure nutrient cycling rates, the mesocosms had a relatively low P concentration and it is possible that primary productivity increased due to the ability of these invertebrates to increase nutrient availability to the algae.

I directly measured the parameters necessary to calculate secondary production of the larval stage only at the beginning (eggs) and near the end of the larval development period (day 31). I intentionally reduced my sampling effort during the larval growth period because of the relatively small size of each aquaria (1290 cm2) and the destructive nature of sampling for invertebrates. For these reasons, it was necessary to estimate production between the beginning of the experiment and day 31 by making a series of

111

assumptions about the shape of the mortality and growth curves. The different mortality scenarios had only a minimal effect on the calculated values of larval production and did not vary the relationship observed with NEP (Figure 4.3). Thus, I am confident that the relationships observed for larval production across the primary productivity gradient are realistic. Nevertheless, absolute values of estimated larval production should be considered with caution when trying to make direct comparisons to field studies.

The strong positive relationship between benthic NEP, light intensity, and larval production indicates that the often observed decline in zoobenthic production with depth could be driven by the subsequent decline in attached algal productivity. Larval production was lowest in the two lowest light treatments which also had a negative NEP

(Figure 4.1e). Light intensity in these two treatments was below 10 μmol m-2 s-1 and it is possible that at this light intensity, attached algal productivity is so low that the zoobenthos would benefit from terrestrial or phytoplankton detritus. I did not amend the diets in this experiment and the two treatments with negative NEP never produced adults.

There is likely a threshold of attached algal productivity beyond which the zoobenthos relies on and uses C from the other available energy sources in lakes in addition to attached algae (i.e. terrestrial and phytoplankton detritus). For example, the use of sedimented detritus leads to increased zoobenthic biomass in depositional areas of lakes

(Rasmussen and Rowan 1997) and the light-limited profundal zones can sustain zoobenthic .

In the few studies that simultaneously measure attached algal primary productivity and benthic secondary production, algal productivity often contributed to explaining zoobenthic production. A synthesis of zoobenthic production along depth

112

intervals by Babler et al. (2008) revealed either a negative relationship or no relationship with depth in lakes. In these comparisons, the available descriptive data (lake size, trophic status, depth, and percentage of area in the littoral zone) did not describe the observed patterns of secondary production. However, Northington et al. (2008) observed a weak but significant relationship between benthic autochthonous production and secondary production in oligotrophic arctic lakes. Mean zoobenthic production in high dissolved organic carbon (DOC) lakes had reduced availability of high O2-habitat (Craig et al. 2015), however, the model with benthic primary productivity was itself significant in explaining the spatial variation in zoobenthic production. Reduced O2 availability could have resulted from low levels of attached algal production. DOC can lead to light attenuation in lakes which can stunt attached algal primary production (Jones 1992, Ask et al. 2012). Thus, resource availability, in the form of attached algae, can often explain patterns in zoobenthic productivity along depth gradients in lakes.

Attached algal productivity was efficiently transferred into chironomid production. It is valuable to explore my data within the context of real-lake estimates of trophic efficiency because of the potentially important ecological implications although I cannot make direct comparisons because many authors measure entire zoobenthic assemblages and use different metrics to estimate efficiencies (i.e. NPP or GPP, and production rather than productivity). The trophic efficiency was 8% and 11% in Hjarboek

Fjord and Lake Myvatn S-basin, respectively, between net zoobenthic primary consumer and net attached algal production (Lindegaard 1992). In Lake Thingvallavatn total zoobenthic production averaged 6% of the estimated available food but ranged between

10-11% in the littoral zone. Castle Lake had low ecological efficiencies in the lake (<

113

3%) but the benthic pathway had the highest efficiency of 0.5-1% using primary production estimates (Vander Zanden et al. 2006). These general results support the hypothesis that relative to phytoplankton pathways, benthic production is efficiently passed up the to higher consumers (Hecky and Hesslein 1995) although energy transfer across the phytoplankton-consumer link can be equally or more efficient than benthic pathways (Jonasson 1992, Vander Zanden et al. 2006).

Benthic consumers are spatially concentrated and the macrozoobenthos like C. dilutus are larger than their counterparts. Additionally, benthic primary producers are highly nutritious (based on C:N, C:P, and polyunsaturated fatty acid content) and spatially concentrated relative to other energy sources in lakes

(Vadeboncoeur and Power 2017). The high nutritional quality and dense aggregation of attached algal carbon might explain why benthic primary production pathways with high trophic efficiencies in lake ecosystems is often observed. This might also contribute to the consistently observed contribution of attached algal C to biomass of top consumers from stable isotope studies of lake food webs (Vadeboncoeur et al. 2003, Vander Zanden et al. 2006, Devlin et al. 2013).

Although the idea that autochthonous primary production contributes energy to a substantial portion of secondary production has existed, and been supported, for decades

(Lindeman 1942, Strayer and Likens 1986, Lindegaard 1994, Blumenshine et al. 1997), the efficiency with which autochthonous production is converted to consumer biomass can have significant ecological consequences. We experimentally show that variation in light can explain much of the variation in zoobenthic production that is often observed in lakes. Attached algal C is concentrated in space and is a highly nutritious resource for

114

consumers. Therefore, it is not surprising that, on average, zoobenthic production can comprise up to 42% of whole-lake secondary production, which is considered a conservative estimate because the productivity of the is often not measured

(Vadeboncoeur et al. 2002). If attached algal production is efficiently transferred into zoobenthic consumer biomass and production as our experiment suggests then the ecological consequences for lake productivity might be large. Thus, C fixed by attached algae is potentially a vital commodity fueling whole-ecosystem production even though the littoral zone can comprise a relatively small area within the total

(Vadeboncoeur et al. 2011). It is notoriously difficult to estimate zoobenthic production accurately, but investigating benthic community primary productivity and secondary production is necessary to understand the flow of energy not only through benthic communities but through entire lake food webs.

Literature Cited Arar, E. J., & Collins, G. B. (1997). Method 445.0: In vitro determination of chlorophyll

a and pheophytin a in marine and freshwater algae by fluorescence. Washington,

DC, USA: United States Environmental Protection Agency, Office of Research

and Development, National Exposure Research Laboratory.

Ask, J., Karlsson, J., & Jansson, M. (2012). Net ecosystem production in clear-water and

brown-water lakes. Global Biogeochemical Cycles, 26(1) GB1017.

Babler, A. L., Solomon, C. T., & Schilke, P. R. (2008). Depth-specific patterns of benthic

secondary production in an oligotrophic lake. Journal of the North American

Benthological Society, 27(1), 108–119.

115

Baulch, H. M., Turner, M. A., Findlay, D. L., Vinebrooke, R. D., & Donahue, W. F.

(2009). Benthic algal biomass — measurement and errors. Canadian Journal of

Fisheries and Aquatic Sciences, 66(11), 1989–2001.

Benke, A. C., Huryn, A. D., Smock, L. A., & Wallace, J. B. (1999). Length-Mass

Relationships for freshwater macroinvertebrates in North America with particular

reference to the Southeastern United States. Journal of the North American

Benthological Society, 18(3), 308–343.

Benoit, D.A., Sibley, P.K., Juenemann, J.L., & Ankley, G.T. (2009). Chironomus tentans

life‐cycle test: Design and evaluation for use in assessing toxicity of contaminated

sediments. Environmental Toxicology and Chemistry, 16(6), 1165–1176.

Bergtold, M., & Traunspurger, W. (2005). Benthic production by micro-, meio-, and

in the profundal zone of an oligotrophic lake. Journal of the North

American Benthological Society, 24(2), 321–329.

Blumenshine, S. C., Vadeboncoeur, Y., Lodge, D. M., Cottingham, K. L., & Knight, S. E.

(1997). Benthic-pelagic links: Responses of benthos to water-column nutrient

enrichment. Journal of the North American Benthological Society, 16(3), 466–

479.

Brey, T. (2012). A multi-parameter artificial neural network model to estimate

macrobenthic invertebrate productivity and production. Limnology and

Oceanography: Methods, 10(8), 581–589.

Brinkhurst, R. O. (1974). Themes in Benthic Research. In The Benthos of Lakes (pp. 3–

6). Palgrave, London.

116

Butkas, K. J., Vadeboncoeur, Y., & Vander Zanden, M. J. (2010). Estimating benthic

invertebrate production in lakes: a comparison of methods and scaling from

individual taxa to the whole-lake level. Aquatic Sciences, 73(1), 153–169.

Cebrian, J. (1999). Patterns in the fate of production in plant communities. The American

Naturalist, 154(4), 449–468.

Chu, S. P. (1942). The influence of the mineral composition of the medium on the growth

of planktonic algae: Part I. Methods and Culture Media. ,

30(2), 284–325.

Craig, N., Jones, S. E., Weidel, B. C., & Solomon, C. T. (2015). Habitat, not resource

availability, limits consumer production in lake ecosystems. Limnology and

Oceanography, 60(6), 2079–2089.

Dermott, R. (1978). Benthic diversity and substrate-fauna associations in Lake Superior.

Journal of Great Lakes Research, 4(3), 505–512.

Devlin, S. P., Vander Zanden, M. J., & Vadeboncoeur, Y. (2013). Depth-specific

variation in carbon isotopes demonstrates resource partitioning among the littoral

zoobenthos. Freshwater Biology, 58(11), 2389–2400.

Devlin, S. P., Vander Zanden, M. J., & Vadeboncoeur, Y. (2016). Littoral-benthic

primary production estimates: Sensitivity to simplifications with respect to

periphyton productivity and basin morphometry. Limnology and Oceanography:

Methods, 14(2), 138–149.

Hamburger, K., Dall, P. C., & Lindegaard, C. (1994). Energy metabolism of Chironomus

anthracinus (Diptera: Chironomidae) from the profundal zone of Lake Esrom,

117

Denmark, as a function of body size, temperature and oxygen concentration.

Hydrobiologia, 294(1), 43–50.

Herren, C. M., Webert, K. C., Drake, M. D.,Vander Zanden, M.J., Einarsson, Á., Ives, A.

R., & Gratton, C. (2017). Positive feedback between chironomids and algae

creates net mutualism between benthic primary consumers and producers.

Ecology, 98(2), 447–455.

Hillebrand, H. (2005). Light regime and consumer control of autotrophic biomass.

Journal of Ecology, 93(4), 758–769.

Hill, W. R., Smith, J. G., & Stewart, A. J. (2010). Light, nutrients, and herbivore growth

in oligotrophic streams. Ecology, 91(2), 518–527.

Hölker, F., Vanni, M. J., Kuiper, J. J., Meile, C., Grossart, H.-P., Stief, P., Adrian, R.,

Lorke, A., Dellwig, O., Brand, A., Hupfer, M., Mooij, W. M., Nutzmann, G., &

Lewandowski, J. (2015). Tube-dwelling invertebrates: tiny ecosystem engineers

have large effects in lake ecosystems. Ecological Monographs, 85(3), 333–351.

Jónasson, P.M. (1972). Ecology and production of the profundal benthos in relation to

phytoplankton in Lake Esrom. Oikos 14 (Supplement): 1–148.

Jónasson, P.M. (1992). The ecosystem of Thingvallavatn: a synthesis. Oikos, 64: 405–

434.

Jones, R. I., Grey, J., Sleep, D., & Arvola, L. (1999). Stable -isotope analysis of

zooplankton carbonnNutrition in humic lakes. Oikos, 86(1), 97–104.

Kajak, Z. (1978). The characteristics of a temperate eutrophic, dimictic lake (Lake

Mikołajskie, Northern Poland). Internationale Revue Der Gesamten

Hydrobiologie und Hydrographie, 63(4), 451–480.

118

Karlsson, J., Byström, P., Ask, J., Ask, P., Persson, L., & Jansson, M. (2009). Light

limitation of nutrient-poor lake ecosystems. Nature, 460(7254), 506–509.

Liess, A., & Hillebrand, H. (2004). Invited review: direct and indirect effects in

herbivore–periphyton interactions. Archiv für Hydrobiologie, 159(4), 433-453.

Lindegaard, C. (1992). The role of zoobenthos in in deep, oligotrophic Lake

Thingvallavatn, Iceland. Hydrobiologia, 243-244(1), 185–195.

Lindegaard, C. (1992). Zoobenthos ecology of Thingvallavatn: Vertical distribution,

abundance, and production. Oikos, 64(1/2), 257–304.

Lindegaard, C., & Jónasson, P. M. (1979). Abundance, population dynamics and

production of zoobenthos in Lake Mývatn, Iceland. Oikos, 32(1/2), 202–227.

Lindeman, R. L. (1942). The trophic-dynamic aspect of ecology. Ecology, 23(4), 399–

417.

Malkin, S. Y., Bocaniov, S. A., Smith, R. E., Guildford, S. J., & Hecky, R. E. (2010). In

situ measurements confirm the seasonal dominance of benthic algae over

phytoplankton in nearshore primary production of a large lake. Freshwater

Biology, 55(12), 2468–2483.

Northington, R. M., Keyse, M. D., Beaty, S. R., Whalen, S. C., Sokol, E. R., & Hershey,

A. E. (2010). Benthic secondary production in eight oligotrophic arctic Alaskan

lakes. Journal of the North American Benthological Society, 29(2), 465–479.

Power, M. E., Stewart, A. J., & Matthews, W. J. (1988). Grazer control of algae in an

Ozark Mountain stream: Effects of short-term exclusion. Ecology, 69(6), 1894–

1898.

119

Rasmussen, J. B. (1984a). Comparison of gut contents and assimilation efficiency of

fourth instar larvae of two coexisting chironomids, Chironomus riparius Meigen

and Glyptotendipes paripes (Edwards). Canadian Journal of Zoology, 62(6),

1022–1026.

Rasmussen, J. B. (1984b). The life-history, distribution, and production of Chironomus

riparius and Glyptotendipes paripes in a prairie . Hydrobiologia, 119(1), 65–

72.

Rasmussen, J. B., & Rowan, D. J. (1997). Wave velocity thresholds for fine sediment

accumulation in lakes, and their effect on zoobenthic biomass and composition.

Journal of the North American Benthological Society, 16(3), 449–465.

Rigler, F. H., & Downing, J. A. (1984). The calculation of secondary productivity. A

manual on methods for the assessment of secondary productivity in fresh waters.

Second Edition, IBP Handbook, 17, 19-58.

Rosemond, A. D., Mulholland, P. J., & Brawley, S. H. (2000). Seasonally shifting

limitation of stream periphyton: response of algal populations and assemblage

biomass and productivity to variation in light, nutrients, and herbivores. Canadian

Journal of Fisheries and Aquatic Sciences, 57(1), 66–75.

Roskosch, A., Hette, N., Hupfer, M., & Lewandowski, J. (2012). Alteration of

Chironomus plumosus ventilation activity and bioirrigation-mediated benthic

fluxes by changes in temperature, oxygen concentration, and seasonal variations.

Freshwater Science, 31(2), 269–281.

Sadler, W.O. 1935. Biology of the midge Chironomus tentans Fabricus, and methods for

its propagation. Mem. Cornell Univ. Agric. Exp. Stat. 173.

120

Sierszen, M. E., Peterson, G. S., & Scharold, J. V. (2006). Depth-specific patterns in

benthic–planktonic food web relationships in Lake Superior. Canadian Journal of

Fisheries and Aquatic Sciences, 63(7), 1496–1503.

Stainton, M. P. (1980). Errors in molybdenum blue methods for determining

orthophosphate in freshwater. Canadian Journal of Fisheries and Aquatic

Sciences, 37(3), 472–478.

Steinman, A. D. (1996). Effects of grazers on freshwater benthic algae. Algal ecology:

freshwater benthic ecosystems, ed. R.J. Stevenson, M.L. Bothwell, R.L. Lowe, pp.

341–73. San Diego: Academic

Strayer, D. L., & Findlay, S. E. G. (2010). Ecology of freshwater shore zones. Aquatic

Sciences, 72(2), 127–163.

Strayer, D., & Likens, G. E. (1986). An energy budget for the zoobenthos of Mirror Lake,

New Hampshire. Ecology, 67(2), 303–313.

Vadeboncoeur, Y., Devlin, S. P., McIntyre, P. B., & Vander Zanden, M. J. (2014). Is

there light after depth? Distribution of periphyton chlorophyll and productivity in

lake littoral zones. Freshwater Science, 33(2), 524–536.

Vadeboncoeur, Y., Jeppesen, E., Vander Zanden, M. J., Schierup, H.-H., Christoffersen,

K., & Lodge, D. M. (2003). From Greenland to green lakes: Cultural

eutrophication and the loss of benthic pathways in lakes. Limnology and

Oceanography, 48(4), 1408–1418.

Vadeboncoeur, Y., Lodge, D. M., & Carpenter, S. R. (2001). Whole-lake fertilization

effects on distribution of primary production between benthic and pelagic

habitats. Ecology, 82(4), 1065–1077.

121

Vadeboncoeur, Y., McIntyre, P. B., & Vander Zanden, M. J. (2011). Borders of

biodiversity: Life at the edge of the World’s large lakes. BioScience, 61(7), 526–

537.

Vadeboncoeur, Y., Peterson, G., Vander Zanden, M. J., & Kalff, J. (2008). Benthic algal

production across lake size gradients: interactions among morphometry, nutrients,

and light. Ecology, 89(9), 2542–2552.

Vadeboncoeur, Y., Vander Zanden, M. J., & Lodge, D. M. (2002). Putting the lake back

together: Reintegrating benthic pathways into lake food web models. BioScience,

52(1), 44–54.

Vander Zanden, M. J., Chandra, S., Park, S.-K., Vadeboncoeur, Y., & Goldman, C. R.

(2006). Efficiencies of benthic and pelagic trophic pathways in a subalpine lake.

Canadian Journal of Fisheries and Aquatic Sciences, 63(12), 2608–2620.

Wetzel, R. G. (2001). Limnology: Lake and River Ecosystems, 3rd. Acadimic pres, AQn

Elsevier imprint, Sanfrancisco, New York, London.

Williams, P. J. B., & Raine, R. C. T. (1979). Agreement between the c-14 and oxygen

methods of measuring phytoplankton production-reassessment of the

photosynthetic quotient. Oceanologica Acta, 2(4), 411-416.

122

Table 4.1. Mean and standard errors (in parentheses) of measured variables: light intensity (μmol/m2/s), water temperature (°C), , sediment P (g/m2), sediment percent P

(%), percent organic matter, and the density of eggs added (no./m2) of each treatment.

Light Intensity Temp (°C) P (g/m2) %P Organic Matter Egg Density (μmol/m2/s) (%) 1 22.9 (0.01) 9.3 (1.5) 0.2 (0.1) 3.0 20500 8 23.0 (0.01) 13.6 (2.9) 0.2 (0.0) 6.0 20000 22 23.0 (0.01) 9.6 (2.5) 0.2 (0.1) 4.0 18200 23 22.8 (0.01) 9.1 (0.3) 0.2 (0.0) 5.0 0 43 22.8 (0.01) 9.1 (2.0) 0.1 (0.0) 3.0 21400 91 22.9 (0.01) 5.6 (2.6) 0.1 (0.0) 4.0 20700 175 22.5 (0.01) 6.8 (1.0) 0.1 (0.0) 3.0 0 177 22.9 (0.01) 5.4 (0.6) 0.1 (0.0) 5.0 19700 211 22.8 (0.01) 12.5 (6.4) 0.2 (0.1) 3.0 19300

123

Table 4.2. Simple linear regression statistics for light, production, and biomass parameters: light intensity (μmol/m2/s), chlorophyll-a (Chl-a, mg/m2), sediment organic matter (AFDM, g/m2), sediment P (g/m2), sediment percent P (%), gross primary production (GPP, mg C/m2/d), net primary production (NEP, mg C/m2/d), larval biomass density (mg/m2), and larval production (mg C/m2/d).

Dependent Independent r2 (adj) Model F DF Variable Variable p-value NEP Light 0.79 0.001 27.52 7 GPP Light 0.79 0.001 31.12 7 Chl-a Light 0.14 0.169 2.342 7 GPP 0.08 0.232 1.172 7 Thickness 0.19 0.132 2.899 7 AFDM Light -0.09 0.588 0.3211 7 P Light -0.04 0.451 0.6377 7 % P Light -0.10 0.6241 0.2627 7 Larval Biomass Light 0.8722 0.001307 41.96 5 NEP 0.9652 4.929e-05 167.2 5 GPP 0.7114 0.01059 15.79 5 Chl-a 0.01768 0.3407 1.108 5 Larval Production Light 0.9445 0.0001588 103.1 5 NEP 0.8351 0.002504 31.38 5 GPP 0.7971 0.004259 24.57 5 Chl-a -0.1033 0.5374 0.438 5

124

a d  2  m  mg   Chl:C OrganicContentMatter 0 0 0 0 500 400 300 200 100 0 . . . . 0.4 0.3 0.2 0.1 0.0

0 50 100 150 200 250 0 50 100 150 200 250

b e  2   1  m  d  2  mg  m   C  mg   0 1000 500 0 PrimaryProduction AttachedAlgal Chlorophyll 02 040 30 20 10 0

0 50 100 150 200 250 0 50 100 150 200 250 2 1 Light Intensitymolm s 

c f max ETR BiofilmThickness (mm) 015 10 5 0 02 04 50 40 30 20 10 0

0 50 100 150 200 250 0 5 10 15 20 25 30 35 2 1 Light Intensitymolm s  Day

125

Figure 4.1. Mean organic matter content (a), algal biomass (b), biofilm thickness (c), biofilms Chl:C ratio (d), and primary productivity (e) as a function of light intensity.

Mean ETRmax over time (f). Circle and square symbols represent aquaria with and without Chironomus dilutus larvae, respectively. In graph (e), closed symbols represent gross primary production and open symbols represent net ecosystem production (NEP).

Bars represent standard error of the mean. Regression statistics in Table 4.2.

126

a d  2  m  mg   DaystoPeak Emergence LarvalBiomass 8 6 4 2 0 03 04 05 60 55 50 45 40 35 30

-200 0 200 400 600 800 -200 0 200 400 600 800

b e  1  d   TotalAdultsNo. SpecificGrowth Rate 010102020300 250 200 150 100 50 0 .001 .00.25 0.20 0.15 0.10

-200 0 200 400 600 800 -200 0 200 400 600 800

c f  2   2 m   m  mg   No.   LarvalDensity TotalAdultBiomass 0 0 0 400 300 200 100 0 0030 0050 0070 8000 7000 6000 5000 4000 3000 2000

-200 0 200 400 600 800 -200 0 200 400 600 800 2 1 2 1 NEPmgCm d  NEPmgCm d 

127

Figure 4.2. Larval biomass (a), specific growth rate (b), larval density (c), days to peak emergence (d), number of adults emerged (e), and total adult biomass (f) as a function of benthic net ecosystem production. Error bars represent standard error of the mean.

Regression statistics are in Table 4.2.

128

a c  1  d   2 1   m d   2  mg m    C  mg  

LarvalProduction Estimate Estimate -10% 04 08 0 120 100 80 60 40 20 0

EstimatedLarval Production Estimate -varied Estimate +10%

04 08 0 120 100 80 60 40 20 0 Initial -0.99

-200 0 200 400 600 800 0 50 100 150 200 250 2 1 NEPmgCm d  Light Intensitymolm2s1

b EcologicalEfficiency 01000102030.4 0.3 0.2 0.1 0.0 -0.1

0 50 100 150 200 2 1 Light Intensitymolm s 

Figure 4.3. Secondary production of Chironomus dilutus as a function of net ecosystem production (a). Production efficiency across the light gradient (b). Estimates of larval production for various mortality rate scenarios: Estimate = production estimate using calculated mortality rates in this paper; Estimate + and – 10% = estimated mortality rate

± 10%; Estimate varied = production estimated using mortality rates that varied over time based on survivorship curves from Jónasson (1979); Initial -0.99 = my calculated mortality rate with the mortality rate as -0.99 for day 0-1. Error bars represent standard error of the mean. Regression statistics in Table 4.2.

129

Chapter Five

PRIMARY CONSUMERS IN AGRICULTURAL STREAMS RELY ON CARBON FIXED BY ATTACHED ALGAE, INDEPENDENT OF THE DENSITY OF RIPARIAN VEGETATION

Introduction Stream food webs are fueled by two major basal resources: algal biofilms adhered to a substrate and detritus transported from the surrounding terrestrial landscape. Spatial variation in the total amount of food available, and the availability of high quality algal carbon relative to low quality terrestrial detritus influences stream food web structure and the flow of energy and material through an ecosystem (Vannote et al. 1980, Nakano and

Murakami 2001). In agricultural landscapes, fertilizer runoff and riparian deforestation increase resources (light, nutrients) for algae and simultaneously decrease food resources for organisms that consume terrestrial detritus (Dodds 2007). The variation in available food resources is likely to affect the abundance of different types of primary consumers and alter food web structure.

Forested headwater streams are fueled by terrestrial detritus, and the energetic importance of autochthonous algal carbon increases downstream as the canopy opens

(Collins et al. 2016, Vannote et al. 1980). The standing stock of terrestrial carbon in forested streams often greatly exceeds the amount of algal carbon. Many macroinvertebrate consumers rely on terrestrial detritus, and the removal of riparian vegetation in agricultural landscapes reduces the supply of detrital material (Reid et al.

130

2008a). At the same time, riparian removal increases light availability to the stream bed, promoting the growth of attached algae. When terrestrial detritus comprises a smaller proportion of the organic carbon pool, consumers may shift their diet to algal-derived carbon or drop out of the community.

As food sources for macroinvertebrates, terrestrial vascular plants and differ greatly in their quality. Biochemical properties of a resource, including carbon-to- nutrient ratios and fatty acid content, determine its food quality for consumers. Although abundant in forested streams, terrestrial detritus is more recalcitrant, it has a higher carbon-to-nutrient ratios, and has a lower relative abundance of essential fatty acids

(EFAs) than algae (Brett et al. 2017). Consequently, in streams that lack riparian trees, macroinvertebrates have access to abundant and high quality algae as an energy source, and a reduced supply of particulate terrestrial organic carbon.

Nutrient enrichment from agriculture (through fertilizer runoff) stimulates periphytic algal growth, often causing proliferation of algal biomass (Dodds 2007).

Increased algal production is often passed on directly to consumers. Nutrient addition increased abundance and production of both macroinvertebrate primary and secondary consumers in a low-order stream (Cross et al. 2006). Higher epilithic chlorophyll-a levels and increased macroinvertebrate abundance was observed in streams along a natural nutrient gradient (Niyogi et al. 2007). However, excessive enrichment also causes large daily fluctuations in dissolved oxygen concentrations, due to high respiration rates in the algal mat during the night. These fluctuations can lead to the loss of species with high oxygen requirements from the invertebrate assemblage (Miltner 2010).

131

Macroinvertebrate species richness declines with nutrient-induced increases in algal biomass (Wang et al. 2007).

Nutrient fertilization also causes the algal assemblage to change, which has consequences for primary consumers. Algal divisions have distinct biochemical composition and ecological functions. Each of the phyla that typically dominate benthic assemblages (diatoms, chlorophytes, and cyanobacteria) has a characteristic and distinct essential fatty acid profile (Frederickson et al. 1986, Dunstan et al. 1994, Taipale et al.

2013). Diatoms produce eicosapentaenoic acid, or EPA, which is an essential fatty acid

(EFA) for most invertebrates (Parrish et al. 1995, Napolitano et al. 1990). EFAs are required for macroinvertebrate growth and development and must be obtained from the diet because many invertebrates lack the enzymes necessary to produce the fatty acids de novo (Brett and Muller-Navarra, 1997). Under high nutrient concentrations, periphyton algal communities become dominated by green filamentous algae (Biggs et al. 1998b).

Green algae have low concentrations of certain EFAs (linoleic acid and α-linoleic acid) and lack the ability to produce the longer-chained polyunsaturated fatty acids (PUFAs) such as EPA and arachidonic acid (Brett and Meuller-Navarra 1997, Taipale et al. 2013).

Thus, when abundant nutrients shift benthic algal assemblages from dominance by diatoms to dominance by green algae, the essential fatty acid pool available to consumers is diluted by the shorter-chain fatty acids produced by the green algae. Consequently, although land use alterations theoretically shift the basal resource pool from one dominated by a relatively poor terrestrial carbon source to the better quality algae, the food quality of the algal assemblage may itself decline.

132

I used carbon and nitrogen stable isotopes and fatty acid analyses to compare food sources for a suite of common benthic invertebrates across a gradient of agricultural land use and riparian forest vegetation. The Midwestern region of the USA is dominated by agriculture, especially row crops. My goal was to compare food web structure across a gradient in agricultural impact and the presence or absence of a riparian vegetation corridor in a landscape where the effect of agriculture is inescapable. First, I assessed the variation in periphyton quality and quantity metrics across the land use gradient. Second,

I assessed the reliance of consumer functional feeding groups on autochthonous versus allochthonous energy sources across the land use gradient and riparian condition. In addition to stable isotopes, fatty acid profiles can be used to assess the reliance of consumers on basal resources from different origins (Pollero et al. 1981, Desvilettes et al.

1994, Guo et al. 2016). I hypothesized that in streams where riparian vegetation corridors were intact, terrestrial carbon would be the dominant carbon food source for benthic primary consumers, and that the contribution of autochthonous carbon would be greater in degraded streams than in stream reaches with intact riparian vegetation. I also expected that as the contribution of agricultural land to total watershed area increases, periphyton would have a lower relative abundance of EFAs (such as eicosapentaenoic acid, EPA) due to a decreasing abundance of diatoms relative to chlorophytes. Assessing these relationships will allow us to identify how food webs in highly impacted stream ecosystems are simultaneously affected by elevated nutrient concentrations and altered riparian land use.

133

Methods I sampled riffle habitats of 12, wadeable low-order streams throughout Ohio,

USA in August – September of 2013. I sampled during low flow conditions when scouring events are infrequent and trees are completely leafed-out. Previous flood events occurred > 30 days prior to sampling. The region is classified as eastern broadleaf forest, but greater than 50% of total land area in Ohio has been converted to agriculture (row crop, pasture, and high density animal farming). I used published land use values to choose sample locations across a gradient in the proportion of the drainage basin devoted to agriculture. I also chose sites that had varied riparian vegetation. I collected water, terrestrial vegetation, periphyton and invertebrates from 2-3 adjacent riffles in each stream to characterize water chemistry, attached algal quality and quantity, terrestrial resource quality, and food web structure.

Sample Collection and Preparation To assess water chemistry and quality, I measured pH, conductivity (Oakton pH/Con 10 Series, Eutech Instruments, Singapore), dissolved oxygen and temperature

(YSI ProODO, YSI, Yellow Springs, Ohio) in the thalweg at each riffle sampled a midday. I collected water samples in triplicate for soluble reactive phosphorus (SRP), total phosphorus (TP), total suspended solids (TSS) and chlorophyll-a analyses. I collected cobbles for periphyton samples, terrestrial detritus (collected in Fall 2014 after leaf fall), and macroinvertebrate samples. Cobbles, terrestrial detritus (naturally occuring submerged leaf packs), and macroinvertebrates were collected from within the riffle. I used kick nets and Surber samplers to collect macroinvertebrates for stable-isotope (SIA), stoichiometric, and fatty acid (FA) analyses. I collected fish for FA analyses using seine

134

nets and fishing poles in October of 2013. All samples were stored on ice and transferred to the laboratory for processing the same or next day.

Metazoans I euthanized the fish, identified them to species, and collected an anterior dorsal muscle sample from each individual. I sorted and identified living macroinvertebrates to the lowest practical taxonomic level and composited the samples to yield one pooled sample for each taxon for each river. Large-bodied taxa such as crayfish, fish, and

Megaloptera were not pooled. I placed the invertebrates in the refrigerator in shallow water for 24 h in order to minimize the material in their guts. This was especially important for insects because I used entire individuals for stable-isotope and fatty acid analyses. After gut clearing, I removed the shells from clams and snails. I also removed the carapace from crayfish, retaining only the tail muscle. Fish and macroinvertebrate samples were stored in glass vials at -20°C and subsequently freeze-dried (Benchtop 6K

Freeze Dryer, VirTis, New York).

I classified each taxon into one of the following functional feeding groups (FFG): grazer, filter feeder, collector, shredder, , and predator (Cummins and Klug,

1979, Merritt and Cummins, 1996; Table 2). Grazers feed selectively on attached algae.

Filter feeders and collectors mix carbon resources. Filter feeders consume organic matter filtered from the water column and collectors consume organic matter that is located on the benthic surface of the stream. Thus, these two FFGs mix C resources from terrestrial and algal origin. Shredders feed selectively on terrestrial leaf litter. consume detritus, organisms, and primary producers while predators selectively feed on other organisms. Typically, corbiculid clams are classified as filter feeders and psephenid

135

beetles are classified as grazers. I placed these two organism groups into their own FFG because of where I found them within each stream. Corbiculids were usually embedded within the sediments, directly above or below the riffle habitat, while psephenids were consistently found on the bottom surface of rocks.

Water Quality I filtered chlorophyll-a, TSS, and SRP samples using pre-ashed PALL AE glass- fiber filters (1 µm nominal pore size) on the same day that we collected them. TP and

SRP samples were stored frozen at -20°C. I dried TSS filters at 105°C for 24 hours.

Chlorophyll-a was extracted at 4°C for 24 hours in 90% buffered ethanol. I quantified chlorophyll-a fluorometrically (Aquaflour, Turner Designs, Sunnyvale CA) and corrected for pheophytin with an acidification step (Arar and Collins 1997). I measured TP and

SRP spectrophotometrically using the acid-molybdate method (UV-1700

Spectrophotometer, Shimadzu, Japan; Stainton et al. 1977).

Periphyton I collected cobbles (n=2-3) at 2 or 3 riffle habitats within each stream along a transect perpendicular to stream flow. In the laboratory, a periphyton sample was isolated by placing a plastic cap on a flat section of each rock (diameter of 6.3 cm) and scraping all uncovered algae off of the rock with a wire brush (Flecker 1996). I then removed the cap, scraped and collected the periphyton from the defined area. Periphyton samples were then frozen, freeze-dried, weighed, and ground into a fine powder (Hansson 1988). I subsampled periphyton for chlorophyll-a, ash-free dry mass (AFDM), phosphorus content, SIA, stoichiometry, and FA analyses. Chlorophyll-a and phosphorus content

136

were analyzed as described above. I quantified AFDM as loss on ignition (LOI) at 500°C for 1 hour (APHA 2005).

Stable-isotopes, stoichiometry, and fatty acids I measured the isotopic and stoichiometric composition of macroinvertebrates, periphyton, and terrestrial detritus. Subsamples of dried periphyton were acidified by adding drops of 1N-HCl until the sample stopped emitting CO2 gas. I composited replicate periphyton samples to get a single sample for each riffle. I rinsed leaf packs to remove biofilm and invertebrate colonizers and then composited replicates for each riffle.

Small-bodied taxa were pooled at the level of stream while large-bodied taxa were not pooled. Stable carbon (δ13C) and nitrogen (δ15N) isotopes, %C and %N were analyzed using an isotope-ratio mass spectrometer (Finnigan MAT Delta Plus) coupled to an elemental analyzer (Carlo Erba NC2500) at the Cornell University Stable Isotope

Laboratory. Replicates were run on 8% of the samples to determine analytical error.

I analyzed total lipids (TL) on 0.2 g of periphyton and 0.05 g of consumer tissue from functional feeding groups (Table 1). Due to their small mass, I could not get TL information from psephenids and the collector FFG of chironomids. Consumers were combined as described above. Total lipids were extracted using 4 mLof phosphate buffer and 2:1 methanol:chloroform (White et al. 1979, DeForest et al. 2012). A 19:0 phospholipid standard (1,2-dinonadecanoyl-sn-glycerol-3-phosphocholine, Avanti Polar

Lipids, Alabaster, Alabama, USE) was added to determine analytical recovery. Silicic acid chromatography using solid-phase extraction columns (500 mg 6 mL-1, Thermo scientific) was used to separate neutral lipid, glycolipid, and polar lipid fractions by eluting with chloroform, acetone, and methanol, respectively (Zelles 1999). Separated

137

polar lipids were then subjected to an alkaline methanolysis to form fatty acid methyl esters (FAMEs). FAMEs were separated and quantified using a HP GC-FID (HP6890 series, Agilent Technologies, Inc. Santa Clara, CA, USA) gas chromatograph. Biomarker abundance was calculated by dividing individual biomarkers by total biomass (i.e., % mol fraction). We removed rare biomarkers (<1% mol fraction) from subsequent analyses.

Site selection and land use characterization The published values for watershed agricultural land use that we used initially for site selection often included the entire drainage basin (Ohio EPA Division of Surface

Water). These values may not accurately reflect agricultural impacts on the portion of the watershed upstream of the sample site. Therefore, I used ArcGIS 10.3.1 to determine watershed land use for each sampling location. I used the USGS National Elevation

Dataset (NED) and the Pour Point tool in ArcGIS to delineate watersheds. Once each watershed was defined, I calculated land use summaries from the USGS National Land

Cover Database (2011). I categorized land used for cultivated crops, hay, or pasture as

Agriculture. The Forest land use category included deciduous forest, broadleaf forest, mixed forest and herbaceous woody vegetation. The Developed category included lands categorized as developed low intensity, developed medium intensity, developed high intensity, and developed open space, determine by the intensity of land covered with structures. Wetlands dominated by woody plants and emergent herbaceous plants were combined in the Wetland category. I used these derived values in this paper.

To determine the riparian buffer land use, I combined site observations of vegetation type and density with Google Earth (2014) estimates of land cover. I used the

138

ruler function in Google Earth to measure the width of vegetated buffer at a sample location (ranged from 0 to 850 m) and the distance the vegetation extended upstream of the sampling location (ranged from 0 to 36 m). I drew a 50 m buffer on either side of the stream that extended 500 m upstream of the sample site and estimated the proportion of land within this buffer strip composed of woody vegetation (ranged from 0 to 0.9). I also scored the type of vegetation (brush, mature trees, grass, etc.) and how sparse or dense the riparian vegetation was from field observations at each site sampled (scores from 0 to

3). A riparian area with dense mature trees scored highest (3) while an open riparian area composed of grass species scored lowest (0).

I averaged the scores of the four riparian vegetation metrics (width, length upstream, proportion of buffer area, and vegetation score) for each site and averaged site scores by stream. Thus, the riparian index was heavily weighted by buffer width at the stream site. This stream average score ranged from 1.5 to 166 across streams from a possible range of 0 to 250. In this manner, locations with a low riparian score had little to no woody riparian vegetation at the sampling site or upstream of the location. Streams with a high riparian score had robust and dense woody riparian vegetation at and upstream of the sampling location.

Statistical Analysis All statistical analyses were conducted in the statistical software R version 3.3.2

(R Core Team 2015), using the extension package vegan for ordination techniques

(Oksanen et al. 2013) and the package lme4 for linear and ANCOVA models (Bates et al.

2012). Watershed land use percentage data was arcsine square root transformed. Water quality variables and periphyton metrics were log10(x)-transformed when necessary as

139

determined by examination of Shapiro-Wilk test of normality and quantile–quantile plots.

Ratios and FA % mole fraction data were not transformed. I averaged water quality (TSS,

SRP, TP, chlorophyll-a) and periphyton (chlorophyll-a, AFDM, Ash, %P, %N, δ13C,

δ15N) metrics at each site within a stream to assess within-stream variation in water quality and periphyton metrics across the agricultural gradient to validate averaging the sites at the stream level. I used one-way ANOVA and Tukey HSD post-hoc tests for each of these metrics.

I used general linear models (GLM) to analyze differences in water chemistry and periphyton metrics across the agricultural land use gradient. For this analysis, geometric means were calculated for each stream. Candidate models were defined to represent the following classes of variables, reflective of potential drivers of periphyton metrics: 1) water quality variables (TP, SRP, TSS), 2) land use and riparian score. I ran GLM of periphyton and terrestrial detritus δ13C and δ15N signatures across the land use gradient and water quality metrics. I used primary producer δ15N and δ13C in ANCOVA analyses of consumer stable-isotope signatures. I used consumer FFG as a factor. I used Akaike’s information criterion (AIC; Akaike, 1981) forward and backward stepwise predictor selection to identify models that maximize explanatory power while minimizing the number of predictors (AIC selection criteria: AIC > 4).

I used principle components analysis (PCA) to visually assess differences in the quality and quantity of periphyton. The first PCA explored relationships between periphtyon molar N:P, percent N, percent P, AFDM, and chlorophyll-a biomass. The second PCA explored relationships between periphyton fatty acid profiles across sites. I used the major FA groups and quality biomarkers: sum of saturated FA (SAFA), sum of 140

monounsaturated FA (MUFA), sum of polyunsaturated FA (PUFA), sum of highly unsaturated FA (HUFA), sum of ω3 FA, sum of ω6 FA, sum of bacterial FA (BAFA), sum of eicosapentaenoic acid (EPA) and sum of Fungi biomarkers. I used Pearson’s correlation coefficient to explore relationships between periphyton categories in ordination space. In addition, I fit environmental factors (water quality metrics – TP,

SRP, TSS; and land use characteristics -- % Agriculture, Forest, and Developed, and the riparian score) to the PCA to identify what environmental factors explained the variability in periphyton quantity and quality. I used standard options and 1000 permutations in the R package vegan (Version 3.3.2, R Core Team 2013, Oksanen et al.

2013).

I used nonmetric multidimensional scaling (NMDS) to characterize variation in periphyton and consumer FA composition over the agriculture gradient. I compared consumer and periphyton FA composition using the sum of seven FA groups: SAFA,

MUFA, PUFA, ω3 FA, ω6 FA, BAFA, and the ratio ω3:ω6. These seven FA groups characterize the major FA present in periphyton and consumers. I ran NMDS using a

Jaccard distance, Wisconsin double standardization, 2 dimensions, and √(x) transformations. I required a run stress of < 0.2 to accept the model. I used Pearson correlation coefficient to explore relationships between individual FA groups.

To assess how the agricultural impact gradient related to the variation in consumer FA profiles, I extracted scores for each organism and corrected for the site- specific periphyton score. I replotted corrected scores in NMDS space and used

ANCOVA models to assess drivers of the variation in FA composition by FFG in NMDS

141

space. I considered FFG a factor, and periphyton δ15N and terrestrial detritus δ15N as the predictor variables.

Results Land use and water chemistry Scioto Brush Creek in southwestern Ohio had the least agricultural land in the watershed (24%) and the Middle Branch Portage River in northwestern Ohio had the most (89%; Table 5.1). Mean (± SE) water column TP ranged from 8.2 ± 1.1 to 184.6 ±

31.6 and SRP ranged from 2.0 ± 0.2 to 31.8 ± 0.9 µg/L. Water column TP was

2 significantly positively correlated (F1,10 = 8.145, r = 0.39, p < 0.05) with watershed

2 agriculture but not developed land (F1,10 = 0.403, r = -0.05, p > 0.05; Table 5.4). In fact,

TP was the only water quality metric whose variation was correlated with any land use category. Including the riparian vegetation score did not improve upon linear regression models. Additionally, we found no effect of site on any water chemistry metric

(ANOVA, all p > 0.05).

Periphyton biomass and stoichiometry AFDM and chlorophyll-a did not vary predictably across the agricultural gradient or with the point estimates of water chemistry metrics (p > 0.05). Mean C:N and C:P of terrestrial detritus were 29.5 and 157.2, which was approximately 2x greater than the C:N and C:P of periphyton (14.5 and 72.7, respectively; Table 5.3). C:N and C:P did not vary predictably with agriculture or water chemistry metrics.

The first two principal components (PCs) accounted for 72% of the variation in periphyton community metrics across the study streams (Figure 5.2). PC1 explained 44% of the variation and was negatively associated with chlorophyll-a (Pearson’s r, -0.70) and positively correlated with % P (Pearson’s r, 0.80). PC2 explained 28% of the variation 142

and was positively associated with % N (Pearson’s r, 0.84). AFDM did not have a strong effect on sample ordination. We fit vector averages of environmental variables for water chemistry (TP, SRP, and TSS) and land use data (agriculture, developed, forest, wetland, and riparian vegetation score). Neither water chemistry metrics nor land use data were significant predictors of periphyton quantity (AFDM, chlorophyll-a) or quality (%P, %N,

NP; all p > 0.05).

Stable isotopes of primary resources and consumers Across streams, δ13C values of periphyton ranged from -16.8 to -33.2 ‰ (Table

5.3). δ13C values of terrestrial detritus varied from -28.3 to -32.3 ‰. Terrestrial detritus was typically more depleted in δ13C than periphyton within a given stream. Neither basal resource sampled had a δ13C signature that varied predictably with agricultural land use or point estimates of water chemistry (all p > 0.05). In contrast, terrestrial detritus and periphyton δ15N were positively correlated with agricultural land use and with each other

(Pearson’s r = 0.84, p < 0.001, Figure 5.3a, 5.3b). Periphyton δ15N was positively

2 correlated with water column TP (F1,10 = 24.97, r = 0.69, p < 0.001, Table 5.4).

Periphyton δ15N signature was on average 3x higher than the δ15N signature of terrestrial detritus and both basal resources were more enriched in δ15N in streams with higher percentages of land devoted to agriculture.

Although basal resource δ13C signatures did not overlap, consumer δ13C signatures did not provide reliable separation with which to infer feeding relationships

(Table 5.5, Table 5.3). FFG δ13C values often overlapped even though we would expect them to separate based on feeding ecology. For example, the δ13C signatures of grazers (-

25.4 to -28.8‰) and shredders (-25.5 to -35.2‰) did not have separation across the

143

agricultural gradient or within an individual stream (Table 5.5). Furthermore, other FFG also had similar δ13C signatures: omnivores (-25.7 to -28.5‰), filter feeders (-26.4 to -

31.3), and collectors (-24.4 to -28.1). Thus, we could not discern the carbon source supporting consumers or the feeding relationships using δ13C isotopic signatures.

Consumer δ15N varied between FFG and across streams (Table 5.6). The δ15N of periphyton was the best predictor of consumer δ15N signatures and the intercept varied depending on FFG (Figure 5.4 a-d). For example, expected specialist FFGs, shredders and grazers, had δ15N signatures that ranged from 1.9 to 3.5 and 5.6 to 14.8‰ , respectively, among streams. There were relatively few representatives of predatory organisms, and we did not did not collect predators at many (6) sites. Therefore, I omitted this FFG from ANCOVA analyses. For all streams, periphyton δ15N explained more

15 15 variation in consumer tissue δ N signatures than leaf δ N (ANCOVA, Periphyton: F8,69

2 2 = 8.296, R = 0.43, p < 0.0001; Leaf: F8,69 = 4.569, R = 0.27, p < 0.001, Figure 5.4a-d).

ANCOVA of the δ15N signature of periphyton and consumer FFG δ15N was significant for Filter Feeder, Grazer, Omnivore, and Shredder FFGs while the ANCOVA of the δ15N signature of leaf packs was significant for Filter Feeder, Collector, and Shredder FFGs

(Table 5.7).

Total lipids of periphyton and consumers Mean periphyton fatty acid content of major groups (PUFA, HUFA, SAFA,

MUFA, ω3 and ω6) did not vary predictably across the agricultural gradient when assessed alone. However, variation in periphyton fatty acid profiles were explained by the

PC ordination (Figure 5.5). PC1 explained 47% of the variation and was positively associated with biofilm chlorophyll-a and AFDM (Pearson’s r = 0.34 and 0.25,

144

respectively). PC2 explained 25% of the variation and was positively associated with agriculture (Pearson’s r = 0.63). The physiologically important EPA ranged from 1.7 to

5.8 % mole fraction across study streams but did not vary predictably across the agricultural gradient. The ratio of diatom indicator fatty acids (16:1ω7 and 20:5ω3) to green and cyanobacteria indicator fatty acids (18:2ω6 and 18:3ω3) declined across the watershed agricultural gradient (Figure 5.6).

Variation in FA profiles among the 7 groups that represented the stream food web organisms of this study was illustrated by the NMDS (Figure 5.7). The composition of stream consumers in NMDS clustered more strongly by taxonomic group than by stream.

Organisms whose sample consisted of only muscle or dissected tissue (e.g. crayfish and snails) rather than whole-body (e.g. tipulids and hydropsychids) tended to cluster together. Grazers, predators, and omnivores loaded most closely to periphyton. Shredders and filter feeders loaded further away from periphyton. The FA groups with the greatest effect on sample ordination were MUFA, ω3:ω6, and PUFA. MUFA and PUFA were responsible for the distribution of samples along the NMDS1 which separated algae from invertebrates except for shredders (Pearson’s r = MUFA = 0.95, PUFA = -0.95). NMDS2 appeared to separate samples more by stream and was most strongly correlated with

ω3:ω6 (Pearson's r = -0.93). Periphyton δ15N signature was the best predictor of FFG

NMDS ordination based on ANCOVA models of periphyton corrected NMDS scores

2 (F6,69 = 72, R = 0.77, p < 0.0001; Figure 5.8).

Discussion I used FA and carbon:nutrient ratios to assess the quality of stream resources (leaf litter, algal biofilms) and combined bulk stable isotope (C, N) and total lipids analyses to

145

trace the movement of these resources in the stream food web. In general, attached algal biomass and quality were not correlated with water quality metrics. I found that the common functional feeding groups in these Midwestern streams consistently relied on C fixed by attached algae irrespective of riparian vegetation across an agricultural land use gradient. Consumer lipid profiles reflected changes in periphyton fatty acid profiles along the agricultural impact gradient (measured here using periphyton δ15N), a pattern that has also been observed in grazers in subtropical streams (Guo et al. 2016). Additionally, consumer FFGs had the same relationship with respect to each other based on δ15N regardless of agricultural impact and potential resource availability.

We did not see a strong relationship between nutrient concentrations and the amount of periphyton in the stream, irrespective of whether periphyton was quantified as

AFDM or chlorophyll-a. The point estimate of water column nutrient concentrations is likely not reflective of long-term conditions experienced by biofilm communities. In general, strong relationships between nutrients and biofilm chlorophyll-a in streams remain elusive within and across ecoregions (Dodds et al. 1997, Biggs et al. 1999,

Chetelat et al. 1999, Dodds et al. 2002, Miltner 2010). Attached algal biomass is not strictly driven by nutrient availability in streams. Biofilm accumulations are regularly removed by scouring events and there can be strong top-down control of biomass via grazing activity (Power and Stewart 1987, Riseng et al. 2004). We could not explain the variation in algal biomass among sites, but periphyton biofilms had relatively high algal biomass which is consistent with values measured in other Ohio streams (Miltner 2010).

As the interface between land and streams, riparian zones play a filtering role in regulating nutrient and sediment transport from landscapes to rivers (Osborne and

146

Kovacic, 1993; Hill, 1996; Goss et al., 2014). Riparian vegetation could not explain the variation in SRP or TP concentrations in these 12 rivers. Hydrology and riparian zones work in conjunction to transport and retain nutrients and they must be assessed together in order to determine the effectiveness of riparian vegetation to sequester nutrients from the adjacent landscape (Connolly et al. 2015). At the reach or watershed scale, the nutrient filtering effects of riparian zones are variable (McDowell 2001, Castillo 2010,

Connor et al., 2013) and their effectiveness depends not only on the characteristics of the riparian vegetation and surrounding land use, but also on hydrology and soil composition

(de Souza et al. 2013). It is possible that we saw no effect of riparian vegetation on nutrient concentrations among our sites because our riparian score was too coarse grained and we did not measure other crucial aspects affecting the movement of water to the stream site. Alternatively, the relevance of riparian vegetation may have been subverted through field management practices such as tile drainage systems. Subsurface tile drains are an important conduit for P transport and they are increasingly being employed in the

Midwest (King et al. 2018).

The δ15N signature of periphyton biofilms was positively correlated with water column TP and agricultural land use. N sources have distinct isotopic values and can be used to identify anthropogenic contribution in N loads to aquatic systems. The primary fertilizers used for agriculture in Ohio are manure from livestock and anhydrous ammonia (Ohio Department of Agriculture). Anhydrous ammonia has relatively low δ15N values between -3 to 3‰, while livestock waste nitrogen is enriched in δ15N and can be as high as 10-20‰ (Valiela et al. 2000). Biogeochemical transformations of nitrogen such as denitrification, and ammonia volatilization can result in substantial isotopic

147

fractionation (Owens 1987, Kellman et al. 1998). Thus inorganic fertilizer inputs with initially low δ15N signatures result in elevated δ15N values in receiving bodies that are traceable through aquatic food webs (MClelland and Valiela 1998, Vander Zanden et al.

2005). The strong positive correlation between percent agriculture and δ15N of primary producers indicates that it δ15N was a good baseline of agriculture. Therefore, we used it as our estimation of time-integrated agricultural impact.

The strong positive correlation between leaf and periphyton δ15N, even though we sampled them nearly a year apart, suggests that δ15N signatures of primary producers reflect landscape-level effects. Periphyton had consistently higher δ15N than leaves at a given stream, which might reflect a source effect. Periphyton was collected from inorganic substrate, and thus, has two potential sources of N, the water column and biofilm . Nitrogen fixation is energetically costly relative to inorganic nitrogen uptake and is likely not favored in a saturated landscape (Vitousek et al. 2002).

Thus, the water column is the probable source of inorganic N for periphyton which likely has an elevated signature from agricultural inputs. In contrast, trees have an underground network of roots that extends deep into the soil, potentially providing trees with multiple sources of N. The inorganic N content of soil controls both the N concentration of leaves as well as the δ15N values assimilated by the roots (Fisher and Binkley, 2000, Hogberg

1997) which are both a function of soil water δ15N values (Handley et al. 1994).

Available data suggests that there is little isotopic fractionation of N during uptake by roots signifying that the observed δ15N signature in leaves reflects land use effects on

δ15N values in soil water (Karamanos and Rennie 1980).

148

Light intensity and nutrient availability determine algal elemental ratios (Hill et al. 2010). Algae obtain C by photosynthesis and take up N and P from their surroundings.

Land use and associated high nutrient loads can vary benthic nutrient stoichiometry, which in turn can affect nutrient demand and uptake (Price and Carrick 2016). Water column phosphorus concentrations did not correlate with periphyton C:P among streams in this study. Periphyton phosphorus content can be nonlinearly related with water column phosphorus concentrations and is often more closely related to periphyton growth rate (Hill and Fanta 2008). Furthermore, there is a negative effect of light on algal nutrient content due to the accumulation of carbon during photosynthesis under high light conditions (Dickman et al. 2006, Hill and Fanta 2008).

Attached algal community composition measured as fatty acid content was not affected by water column nutrients or land use gradients when individual markers were examined. However, we saw a decrease in diatom markers relative to chlorophyte and cyanobacteria markers as agricultural land use increased. Diatoms produce EPA, which is an essential fatty acid used to maintain membrane fluidity (Hodkova et al. 1999) and can positively influence invertebrate growth rates (Brett et al. 2009). Diatoms in pristine streams develop thin biofilms attached to the surface of rocks. As nutrients concentrations increase, there is often a switch from diatom-dominated biofilms to green algae-dominated biofilm communities (Biggs et al. 1998b, Suren et al. 2003). Many of the chlorophytes that do well under high nutrient concentrations are filamentous and a relatively poor supply of PUFA. This decrease in diatoms relative to other algae may have negative consequences for consumers because of the diluting factor of increased

149

green algae on relative availability of EPA, indicating lower overall food quality of the attached biofilm.

We collected common macroinvertebrates that have known ecological feeding modes to assess how they respond to variation in resource availability. The two metrics that we employed to characterize the food web structure and carbon resource use by consumers (fatty acids and SIA) indicate that autochthonous resources are important to the food webs in our study streams, irrespective of agricultural impact. However, δ13C did not provide good separation between FFGs even though periphyton δ13C was generally less negative than leaf δ13C for a given stream. One interpretation of the high degree of overlap of δ13C signatures among organisms of different FFGs is that all consumers were supported by both algal and terrestrial organic matter. However, periphyton δ13C, which was more variable within and among streams, is a function of the combined influence of dissolved inorganic carbon and diffusional effects on fractionation whereby high flow velocity has a strong negative effect on periphyton δ13C (Finlay et al.

1999, Rasmussen and Trudeau 2007). Furthermore, the δ13C of bulk periphyton does not necessarily reflect what is actually being consumed. Experiments have demonstrated that increased periphyton productivity leads to higher δ13C at relatively small spatial scales

(Hill et al., 2008) and that consumers selectively graze on the rapidly growing cells at or near the surface of periphyton biofilms (Rezanka and Hershey, 2003). Therefore, we used

δ15N of periphyton and leaves to distinguish trophic relationships among functional feeding groups. In general, periphyton had higher δ15N signatures than terrestrial leaves.

The separation of the specialist consumer groups, shredders and grazers, using their

150

preferred food source δ15N signature (leaf and periphyton, respectively) confirmed that this was an appropriate metric of consumer resource use.

The correlation of taxon-specific δ15N signature across the agricultural gradient implies specific FFG are consuming the same thing. The change in intercept is different between periphyton and terrestrial detritus (leaf) eaters. For example shredders and collectors consistently had δ15N signatures indicating a terrestrial diet while grazers, omnivores, and filter feeders had a periphyton diet. Furthermore food web structure appears to be inflexible across the impact gradient. Collectors, which typically have mixed diets (Merritt and Cummins, 1996), fell out between grazers and shredders in δ15N space. Psephenids are considered grazers, but we only found them on the bottom of rocks and stable isotope signals indicate that they might be consuming terrestrial detritus or bacteria. We also decided to keep corbiculid clams as their own FFG because they were only found within sediments and are associated with eating FPOM, a resource likely of terrestrial origin (Vannote et al. 1980, Allan 1995). The robust nature of food web structure across the impact gradient indicates that consumers are partitioning resource space in the same way no matter the resource abundance. Our data suggest that the diets of individual taxa are not flexible across potential changes in resource pools.

The ordination plots of periphyton and macroinvertebrate FA profiles reveal two main patterns. First, FA composition varies among invertebrate taxa. This is likely a consequence of a physiological constraint (Guo et al. 2016). Invertebrates may be capable of regulating PUFA composition through converting shorter PUFAs into the longer- chained counterparts, thereby, mediating their own PUFA composition (Goedkoop et al.

151

2000, Guo et al. 2016). The extent of this ability is unknown and potentially taxon- dependent (Guo et al. 2016).

Second, the observed flexibility within taxa appears to be associated with periphyton fatty acid profiles. The plasticity within taxa-specific fatty acid profiles could be the result of age class or body size effects on fatty acid concentrations (Iverson et al.

2002). Alternatively, consumer lipid profiles varied with periphyton profiles along the

PUFA to MUFA axis (NMDS1). The observed patterns indicated that the biochemical composition of invertebrate tissue largely reflected algal FA, consistent with previous findings (Torres-Ruiz et al. 2010, Guo et al. 2016). Furthermore, when we subtracted periphyton-specific NMDS scores from consumer-specific NMDS scores, which represented the distance away from the periphyton fatty acid profile, periphyton δ15N explained the variation in invertebrate fatty acid concentrations across the NMDS1 axis.

The positive relationship between periphyton δ15N and corrected NMDS1 scores suggests that as agricultural impact increased, the PUFA content of invertebrates decreased, making them a lower quality resource for higher consumers. Again, this relationship was taxon-specific, whereby grazers and omnivores had higher PUFA content than shredders.

High retention of dietary PUFA in stream consumers means high availability for the next (Brett and Meuller-Navarra 1997, Ravet et al. 2010).

Patterns in carbon resource use by consumers in stream ecosystems is difficult to determine because it reflects the combined effect of changes in resource availability and resource quality. Temperate streams drain landscapes that are dominated by anthropogenic alterations, predominately agriculture. This study suggests that land use modifications and the removal of terrestrial riparian vegetation alters the chemical

152

composition of available food sources which in turn is reflected in consumer biochemical composition. The reduction in EPA relative to green algae markers across the agricultural gradient can potentially reduce the quality of food available to consumers. Stable isotope analysis and FA profiles gave complementary results indicating the contribution of periphyton to stream food webs. This study demonstrates that food web structure remains robust across a wide gradient of stream agricultural impact. This suggests that C flow through these streams does not differ across the range of agricultural impact and that attached algae remains the dominant source of energy supporting food webs.

Literature Cited Ahlgren, G., Vrede, T., & Goedkoop, W. (2009). Fatty Acid Ratios in ,

Zooplankton and Zoobenthos – Are There Specific Optima? In M. Kainz, M. T.

Brett, & M. T. Arts (Eds.), Lipids in Aquatic Ecosystems (pp. 147–178). Springer

New York.

Akaike, H. (1981). Likelihood of a model and information criteria. Journal of

econometrics, 16(1), 3-14.

Allan, J. D. (1995). Organic matter in lotic ecosystems. In Stream Ecology (pp. 259–281).

Springer Netherlands.

APHA-American Public Health Association. (2005). Standard methods for examination

of water and wasterwater. Washington, DC: APHA..

Arar, E. J., & Collins, G. B. (1997). Method 445.0: In vitro determination of chlorophyll

a and pheophytin a in marine and freshwater algae by fluorescence. Washington,

DC, USA: United States Environmental Protection Agency, Office of Research

and Development, National Exposure Research Laboratory.

153

Biggs, B. J. F. (2000). Eutrophication of streams and rivers: dissolved nutrient-

chlorophyll relationships for benthic algae. Journal of the North American

Benthological Society, 19(1), 17–31.

Binkley, D., Son, Y., & Valentine, D. W. (2000). Do forests receive occult inputs of

nitrogen? Ecosystems, 3(4), 321–331.

Brett, M. T., Bunn, S. E., Chandra, S., Galloway, A. W. E., Guo, F., Kainz, M. J.,

Kankaala, P., Lau, D. C.P., Moulton, T. P., Power, M. E., Rasmussen, J. B.,

Taipale, S. J., Thorp, J. H. & Wehr, J. D. (2017). How important are terrestrial

organic carbon inputs for secondary production in freshwater ecosystems?

Freshwater Biology, 62(5), 833–853.

Brett, M. T., Kainz, M. J., Taipale, S. J., & Seshan, H. (2009). Phytoplankton, not

allochthonous carbon, sustains herbivorous zooplankton production. Proceedings

of the National Academy of Sciences, 106(50), 21197–21201.

Brett, M. T., & Müeller-Navarra, D. C. (1997). The role of highly unsaturated fatty acids

in aquatic foodweb processes. Freshwater Biology, 38(3), 483–499.

Castillo, M. M., Morales, H., Valencia, E., Morales, J. J., & Cruz-Motta, J. J. (2012). The

effects of human land use on flow regime and water chemistry of headwater

streams in the highlands of Chiapas. Knowledge and Management of Aquatic

Ecosystems, (407), 09.

Chételat, J., Pick, F. R., Morin, A., & Hamilton, P. B. (1999). Periphyton biomass and

community composition in rivers of different nutrient status. Canadian Journal of

Fisheries and Aquatic Sciences, 56(4), 560–569.

154

Collins, S. M., Kohler, T. J., Thomas, S. A., Fetzer, W. W., & Flecker, A. S. (2016). The

importance of terrestrial subsidies in stream food webs varies along a stream size

gradient. Oikos, 125(5), 674–685.

Connolly, N. M., Pearson, R. G., Loong, D., Maughan, M., & Brodie, J. (2015). Water

quality variation along streams with similar agricultural development but

contrasting riparian vegetation. Agriculture, Ecosystems & Environment, 213, 11–

20.

Connor, S., Nelson, P. N., Armour, J. D., & Hénault, C. (2013). Hydrology of a forested

in an agricultural landscape of the humid tropics. Agriculture,

Ecosystems & Environment, 180, 111–122.

Cross, W. F., Wallace, J. B., Rosemond, A. D., & Eggert, S. L. (2006). Whole-system

nutrient enrichment increases secondary production in a detritus-based ecosystem.

Ecology, 87(6), 1556–1565.

Cummins, K. W., & Klug, M. J. (1979). Feeding ecology of stream invertebrates. Annual

Review of Ecology and Systematics, 10(1), 147–172.

DeForest, J. L., Smemo, K. A., Burke, D. J., Elliott, H. L., & Becker, J. C. (2012). Soil

microbial responses to elevated phosphorus and pH in acidic temperate deciduous

forests. Biogeochemistry, 109(1-3), 189–202.

Desvilettes, C., Bourdier, G., Amblard, C., & Barth, B. (1994). Use of fatty acids for the

assessment of zooplankton grazing on bacteria, protozoans and microalgae.

Freshwater Biology, 38(3), 629–637.

Dickman, E. M., Vanni, M. J., & Horgan, M. J. (2006). Interactive effects of light and

nutrients on phytoplankton stoichiometry. Oecologia, 149(4), 676–689.

155

Dodds, W. K. (2007). Trophic state, eutrophication and nutrient criteria in streams.

Trends in Ecology & Evolution, 22(12), 669–676.

Dodds, W. K., Smith, V. H., & Lohman, K. (2002). Nitrogen and phosphorus

relationships to benthic algal biomass in temperate streams. Canadian Journal of

Fisheries and Aquatic Sciences, 59(5), 865–874.

Dunstan, G. A., Baillie, H. J., Barrett, S. M., & Volkman, J. K. (1996). Effect of diet on

the lipid composition of wild and cultured abalone. Aquaculture, 140(1), 115–

127.

Falkowski, P. G., Katz, M. E., Knoll, A. H., Quigg, A., Raven, J. A., Schofield, O., &

Taylor, F. J. R. (2004). The Evolution of Modern Eukaryotic Phytoplankton.

Science, 305(5682), 354–360.

Findlay, S., & Sinsabaugh, R. L. (1999). Unravelling the sources and bioavailability of

dissolved organic matter in lotic aquatic ecosystems. Marine and Freshwater

Research, 50(8), 781–790.

Finlay, J. C. (2001). Stable-carbon-isotope ratios of river biota:implications for energy

flow in lotic food webs. Ecology, 82(4), 1052–1064.

Flecker, A. S. (1996). Ecosystem engineering by a dominant detritivore in a diverse

tropical stream. Ecology, 77(6), 1845–1854.

Fredrickson, H. L., Cappenberg, T. E., & de Leeuw, J. W. (1986). Polar lipid ester-linked

fatty acid composition of Lake Vechten seston: an ecological application of lipid

analysis. FEMS Microbiology Letters, 38(6), 381–396.

Goedkoop, W., Sonesten, L., Ahlgren, G., & Boberg, M. (2000). Fatty acids in profundal

benthic invertebrates and their major food resources in Lake Erken, Sweden:

156

seasonal variation and trophic indications. Canadian Journal of Fisheries and

Aquatic Sciences, 57(11), 2267–2279.

Goss, C. W., Goebel, P. C., & Sullivan, S. M. P. (2014). Shifts in attributes along

agriculture-forest transitions of two streams in central Ohio, USA. Agriculture,

Ecosystems & Environment, 197, 106–117.

Guo, F., Kainz, M. J., Sheldon, F., & Bunn, S. E. (2016). Effects of light and nutrients on

periphyton and the fatty acid composition and somatic growth of invertebrate

grazers in subtropical streams. Oecologia, 181(2), 449–462.

Guo, F., Kainz, M. J., Valdez, D., Sheldon, F., & Bunn, S. E. (2016). The effect of light

and nutrients on algal food quality and their consequent effect on grazer growth in

subtropical streams. Freshwater Science, 35(4), 1202–1212.

Handley, L. L., & Scrimgeour, C. M. (1997). Terrestrial Plant Ecology and 15N Natural

Abundance: The Present Limits to Interpretation for Uncultivated Systems with

Original Data from a Scottish Old Field. In M. Begon & A. H. Fitter (Eds.),

Advances in Ecological Research (Vol. 27, pp. 133–212). Academic Press.

Hansson, L.-A. (1988). Chlorophyll a determination of periphyton on sediments:

identification of problems and recommendation of method. Freshwater Biology,

20(3), 347–352.

Hill, A. R. (1996). Nitrate removal in stream riparian zones. Journal of Environmental

Quality, 25(4), 743–755.

Hill, W. R., Boston, H. L., & Steinman, A. D. (1992). Grazers and nutrients

simultaneously limit lotic primary productivity. Canadian Journal of Fisheries

and Aquatic Sciences, 49(3), 504–512.

157

Hill, W. R., & Fanta, S. E. (2008). Phosphorus and light colimit periphyton growth at

subsaturating irradiances. Freshwater Biology, 53(2), 215–225.

Hill, W. R., Fanta, S. E., & Roberts, B. J. (2008). 13C dynamics in benthic algae: Effects

of light, phosphorus, and biomass development. Limnology and Oceanography,

53(4), 1217–1226.

Hill, W. R., Smith, J. G., & Stewart, A. J. (2010). Light, nutrients, and herbivore growth

in oligotrophic streams. Ecology, 91(2), 518–527.

Hodková, M., Šimek, P., Zahradnı́čková, H., & Nováková, O. (1999). Seasonal changes

in the phospholipid composition in thoracic muscles of a heteropteran,

Pyrrhocoris apterus. Insect Biochemistry and Molecular Biology, 29(4), 367–376.

Högberg, P. (1997). Tansley Review No. 95 15N natural abundance in soil-plant systems.

New Phytologist, 137(2), 179–203.

Homer, C.G., Dewitz, J.A., Yang, L., Jin, S., Danielson, P., Xian, G., Coulston, J.,

Herold, N.D., Wickham, J.D., & Megown, K. (2015). Completion of the 2011

National Land Cover Database for the conterminous United States-Representing a

decade of land cover change information. Photogrammetric Engineering and

Remote Sensing, 81 (5), 345-354.

Iverson, S. J. (2009). Tracing aquatic food webs using fatty acids: from qualitative

indicators to quantitative determination. In M. Kainz, M. T. Brett, & M. T. Arts

(Eds.), Lipids in Aquatic Ecosystems (pp. 281–308). Springer New York.

Karamanos, R. E., & Rennie, D. A. (1980). Changes in natural 15N abundance associated

with pedogenic processes in soil. II. Changes on different slope positions.

Canadian Journal of Soil Science, 60(2), 365–372.

158

Kellman, L., & Hillaire-Marcel, C. (1998). Nitrate cycling in streams: using natural

abundances of NO3--δ15N to measure in-situ denitrification. Biogeochemistry,

43(3), 273–292.

King, K. W., Williams, M. R., LaBarge, G. A., Smith, D. R., Reutter, J. M., Duncan, E.

W., & Pease, L. A. (2018). Addressing agricultural phosphorus loss in artificially

drained landscapes with 4R nutrient management practices. Journal of Soil and

Water Conservation, 73(1), 35–47.

Matthews, W. J., Stewart, A. J., & Power, M. E. (1987). Grazing as components of

North American stream ecosystems: effects of Campostoma anomalum.

Community and of North American stream fishes, 128-135.

McClelland, J. W., & Valiela, I. (1998). Changes in food web structure under the

influence of increased anthropogenic nitrogen inputs to . Marine Ecology

Progress Series, 168, 259–271.

McClelland, J. W., Valiela, I., & Michener, R. H. (1997). Nitrogen-stable isotope

signatures in estuarine food webs: A record of increasing urbanization in coastal

watersheds. Limnology and Oceanography, 42(5), 930–937.

McDowell, W. H. (2001). Hurricanes, people, and riparian zones: controls on nutrient

losses from forested Caribbean watersheds. Forest Ecology and Management,

154(3), 443–451.

Merritt, R. W. & K. W. Cummins, 1996. An Introduction to the Aquatic Insects of North

America. Kendall/Hunt Publishing Co. Dubuque, Iowa.

Miltner, R. J. (2010). A method and rationale for deriving nutrient criteria for small rivers

and streams in Ohio. Environmental Management, 45(4), 842–855.

159

Nakano, S., & Murakami, M. (2001). Reciprocal subsidies: Dynamic interdependence

between terrestrial and aquatic food webs. Proceedings of the National Academy

of Sciences, 98(1), 166–170.

Niyogi, D. K., Koren, M., Arbuckle, C. J., & Townsend, C. R. (2007). Stream

communities along a catchment land-use gradient: Subsidy-stress responses to

pastoral development. Environmental Management, 39(2), 213–225.

Oksanen, J., Blanchet, F. G., Kindt, R., Legendre, P., Minchin, P. R., O’Hara, R. B.,

Simpson, G.L., Solymos, P., Stevens, M.H.H., Wagner, H. & Oksanen, M.J.

(2012). vegan: Community Ecology Package: R package version 2.1-13/r2115.

Osborne, L. L., & Kovacic, D. A. (1993). Riparian vegetated buffer strips in water-

quality restoration and stream management. Freshwater Biology, 29(2), 243–258.

Owens, N. J. P., Woodward, E. M. S., Aiken, J., Bellan, I. E., & Rees, A. P. (1990).

Primary production and nitrogen assimilation in the North Sea during July 1987.

Netherlands Journal of Sea Research, 25(1), 143–154.

Polis, G. A., Anderson, W. B., & Holt, R. D. (1997). Toward an Integration of Landscape

and Food Web Ecology: The dynamics of spatially subsidized food webs. Annual

Review of Ecology and Systematics, 28(1), 289–316.

Pollero, R. J., Brenner, R. R., & Gros, E. G. (1981). Seasonal changes in lipid and fatty

acid composition of the freshwater mollusk, Diplodon patagonicus. Lipids, 16(2),

109–113.

Power, M. E., & Rainey, W. E. (2000). Food webs and resource sheds: towards spatially

delimiting trophic interactions. In: Ecological consequences of habitat

160

heterogeneity, Eds M.J. Hutchinson, E.A. John & A.J.A. Stewart, 291-314.

Blackwell Scientific, Oxford, U.K.

Power, M. E., & Stewart, A. J. (1987). Disturbance and recovery of an algal assemblage

following flooding in an Oklahoma stream. The American Midland Naturalist,

117(2), 333–345.

Price, K. J., & Carrick, H. J. (2016). Effects of experimental nutrient loading on

phosphorus uptake by biofilms: evidence for nutrient saturation in mid-Atlantic

streams. Freshwater Science, 35(2), 503–517.

Rasmussen, J. B., & Trudeau, V. (2007). Influence of velocity and chlorophyll standing

stock on periphyton δ13C and δ15N in the Ste. Marguerite River system, Quebec.

Canadian Journal of Fisheries and Aquatic Sciences, 64(10), 1370–1381.

Ravet, J. L., Brett, M. T., & Arhonditsis, G. B. (2010). The effects of seston lipids on

zooplankton fatty acid composition in Lake Washington, Washington, USA.

Ecology, 91(1), 180–190.

Reid, D. J., Lake, P. S., Quinn, G. P., & Reich, P. (2008). Association of reduced riparian

vegetation cover in agricultural landscapes with coarse detritus dynamics in

lowland streams. Marine and Freshwater Research, 59(11), 998–1014.

Rezanka, K. M., & Hershey, A. E. (2003). Examining primary producer–consumer

interactions in a Lake Superior tributary using 15N-tracer, grazer-reduction, and

nutrient-bioassay experiments. Journal of the North American Benthological

Society, 22(3), 371–387.

Riseng, C. M., Wiley, M. J., & Stevenson, R. J. (2004). Hydrologic disturbance and

nutrient effects on benthic community structure in midwestern US streams: a

161

covariance structure analysis. Journal of the North American Benthological

Society, 23(2), 309–326.

Souza, A. L. T. de, Fonseca, D. G., Libório, R. A., & Tanaka, M. O. (2013). Influence of

riparian vegetation and forest structure on the water quality of rural low-order

streams in SE Brazil. Forest Ecology and Management, 298, 12–18.

Stewart, A. J. A., John, E. A., & Hutchings, M. J. (2000). The world is heterogeneous:

ecological consequences of living in a patchy environment. In: The Ecological

Consequences of Environmental Heterogeneity. The 40th Symposium of the

British Ecological Society. Eds Hutchings MJ, John EA, Stewart AJA, 9–32.

Blackwell Science, Oxford, U.K.

Suren, A. M., Biggs, B. J. F., Duncan, M. J., Bergey, L., & Lambert, P. (2003). Benthic

community dynamics during summer low‐flows in two rivers of contrasting

enrichment 2. Invertebrates. New Zealand Journal of Marine and Freshwater

Research, 37(1), 71–83.

Taipale, S., Strandberg, U., Peltomaa, E., Galloway, A. W. E., Ojala, A., & Brett, M. T.

(2013). Fatty acid composition as biomarkers of freshwater microalgae: analysis

of 37 strains of microalgae in 22 genera and in seven classes. Aquatic Microbial

Ecology, 71(2), 165–178.

Torres-Ruiz, M., Wehr, J. D., & Perrone, A. A. (2007). Trophic relations in a stream food

web: importance of fatty acids for macroinvertebrate consumers. Journal of the

North American Benthological Society, 26(3), 509–522.

162

Valiela, I., Tomasky, G., Hauxwell, J., Cole, M. L., Cebrián, J., & Kroeger, K. D. (1998).

Operationalizing sustainability: Management and risk assessment of land-derived

nitrogen loads to estuaries. Ecological Applications, 10(4), 1006–1023.

Vander Zanden, M. J., Vadeboncoeur, Y., Diebel, M. W., & Jeppesen, E. (2005). Primary

consumer stable nitrogen isotopes as indicators of nutrient source. Environmental

Science & Technology, 39(19), 7509–7515.

Vannote, R. L., Minshall, G. W., Cummins, K. W., Sedell, J. R., & Cushing, C. E.

(1980). The river continuum concept. Canadian Journal of Fisheries and Aquatic

Sciences, 37(1), 130–137.

Vitousek, P. M., Cassman, K., Cleveland, C., Crews, T., Field, C. B., Grimm, N. B.,

Howarth, R. W., Marino, R., Martinelli, L., Rastetter, E. B., & Sprent, J. I. (2002).

Towards an ecological understanding of biological nitrogen fixation. In The

Nitrogen Cycle at Regional to Global Scales (pp. 1–45). Springer, Dordrecht.

Wang, L., Robertson, D. M., & Garrison, P. J. (2007). Linkages between nutrients and

assemblages of macroinvertebrates and fish in wadeable streams: Implication to

nutrient criteria development. Environmental Management, 39(2), 194–212.

White, D. C., Davis, W. M., Nickels, J. S., King, J. D., & Bobbie, R. J. (1979).

Determination of the sedimentary microbial biomass by extractible lipid

phosphate. Oecologia, 40(1), 51–62.

Zelles, L. (1999). Fatty acid patterns of phospholipids and lipopolysaccharides in the

characterisation of microbial communities in soil: a review. Biology and Fertility

of Soils, 29(2), 111–129.

163

Table 5.1. Calculated watershed land use and mean measured water column characteristics for the study streams. Land use categories include agriculture (hay, pasture, and cultivated crops), developed land, forest, and wetlands (%). Total phosphorus (TP), soluble reactive phosphorus (SRP) and water column chlorophyll-a (Chl-a) are in µg/L. Total suspended solids (TSS) are in mg/L. Riffles refers to the number of riffles sampled in each stream. Values in parentheses represent standard errors.

Stream Riparian Stream Abbreviation Agriculture Developed Forest Wetlands TP SRP Chl-a TSS Score Riffles Scioto Brush Creek SBC 23.9 6.87 67.83 0.01 8.2 (1.1) 16.3 (0.3) 1.5 (0.3) 3.4 (0.7) 28.7 3 Little Beaver Creek LBC 39.4 9.94 50.46 0 15.6 (0.6) 24.0 (0.8) 57.7 (4.9) 3.3 (0.2) 23.7 3 Vermilion River VR 63.3 7.09 27.59 1.24 66.3 (11.6) 23.4 (2.4) 2.5 (0.1) 8.3 (0.5) 5.5 2 Nimishillen Creek NC 63.6 24.5 11.19 0.26 96.3 (4.7) 7.1 (0.3) 1.1 (0.3) 4.3 (0.6) 11.5 2 East Fork White Oak Creek EFWOC 67.9 6.15 25.97 0 30.4 (1.3) 26.2 (2.4) 1.5 (0.9) 5.6 (1.4) 5.3 3 East Fork Little Miami River EFLMR 77.9 6.34 14.5 0.05 102 (7.0) 30.0 (1.0) 54.2 (13.5) 5.4 (1.1) 10.3 3 Little Miami River LMR 80.5 9.5 9.37 0.15 27.9 (3.4) 30.9 (2.1) 1.4 (0.5) 8.9 (0.5) 31.7 3 Twin Creek TC 81.1 7.37 11.15 0.22 44.0 (4.8) 2.0 (0.2) 3.6 (0.1) 4.6 (0.3) 18.7 3 Blanchard River BR 82.9 7.82 8.61 0.35 171 (31.6) 30.8 (0.3) 24.8 (9.2) 6.2 (2.1) 13.7 3 Little Darby Creek LDC 86.9 6.55 6.11 0.3 45.4 (0.4) 20.6 (0.7) 6.1 (2.6) 3.2 (0.8) 26.0 3 Paint Creek PC 88.4 7.84 3.56 0.03 95.3 (6.6) 31.8 (0.9) 2.0 (0.2) 2.7 (0.1) 16.7 3 Middle Branch Portage River PR 88.7 8.28 2.64 0.06 38.5 (6.1) 3.1 (0.4) 5.6 (1.4) 3.7 (0.3) 1.5 3

164

Table 5.2. Classification of organisms into functional feeding groups based on Merritt and Cummins (1996) and Cummins and Klug

(1979). Organisms were used in stable isotope and total lipids analyses. a denotes functional feeding groups analyzed for stable- isotopes and stoichiometry only. Stream abbreviations as in Table 1. Organism Functional Feeding Tissue Used Stream Group Chironomidae Collector a Whole Body BR, EFLMR, EFWOC, LBC, LDC, LMR, NC, PC, PR, SBC, TC, VR Hydropsychidae Filter Feeder Whole Body BR, EFLMR, EFWOC, LBC, LDC, LMR, NC, PC, PR, SBC, TC, VR Ambloplites rupestris PC Pimephales notatus BR, LBC, LDC, PR, SBC, TC, VR Compostoma anomalum EFLMR Corydalus spp. Predator Whole Body EFLMR, LDC, LMR, PR, SBC, TC Pleuroceridae Grazer Whole body, BR, EFLMR, EFWOC, LDC, LMR, NC, PC, removed from shell SBC, TC, VR Tipulidae Shredder Whole Body BR, EFLMR, EFWOC, LBC, LDC, LMR, NC, PC, PR, SBC, TC, VR Orconectes spp. Omnivore Tail tissue BR, EFWOC, LBC, LDC, LMR, NC, PC, PR, SBC, TC, VR Psephenidae Psephenid a Whole Body BR, EFLMR, EFWOC, LBC, LDC, LMR, NC, PC, SBC, TC, VR Corbicula spp. Clam a Whole body, BR, EFLMR, EFWOC, LDC, LMR, PC, SBC, removed from shell VR

165

Table 5.3. Mean (SE) of periphyton and terrestrial detritus metrics measured in 12 streams. Abbreviations: Chl-a = chlorophyll-a in mg/m2, AFDM = ash free dry mass in g/m2, Ash = ash content in g/m2, P = phosphorus content in g/m2, P% = percent phosphorus,

N% = percent nitrogen, C% = percent carbon. Streams are ordered by increasing agricultural land use and abbreviations listed in Table

5.1. A “--“ indicates no data.

15 13 Resource Stream Chl-a AFDM Ash P P% N% C% C:N C:P δ N δ C Periphyton SBC 37.4 (17.7) 21.7 (5.1) 66.4 (17.3) 0.05 (0.01) 0.23 (0.10) 2.0 (0.43) 14.5 (4.9) 20.1 (4.0) 135.0 (32.1) 5.9 (0.4) -26.9 (1.8) LBC 28.0 (13.0) 21.6 (7.1) 110 (77.3) 0.10 (0.05) 0.53 (0.16) 4.4 (0.24) 30.6 (7.2) 16.8 (5.5) 63.9 (21.1) 7.7 (0.3) -27.1 (1.9) VR 122 (11.7) 46.3 (25.2) 222 (36.0) 0.19 (0.09) 0.47 (0.05) 4.0 (0.21) 28.5 (1.4) 17.9 (5.9) 86.6 (31.1) 9.6 (0.4) -26.4 (0.6) NC 40.1 (8.8) 42.6 (3.3) 83.1 (24.1) 0.32 (0.03) 0.73 (0.10) 4.6 (0.37) 28.9 (2.4) 11.8 (3.2) 46.6 (16.0) 11.4 (0.4) -33.2 (0.6) EFWOC 30.2 (6.9) 20.3 (4.1) 52.8 (31.1) 0.15 (0.02) 0.74 (0.23) 3.8 (0.45) 29.5 (3.5) 15.1 (4.2) 74.8 (26.0) 7.0 (0.4) -26.8 (1.5) EFLMR 87.8 (30.9) 36.7 (11.4) 178 (63.3) 0.31 (0.15) 0.74 (0.15) 4.3 (0.57) 33.1 (2.8) 6.6 (0.4) 19.2 (4.5) 10.6 (0.2) -26.8 (1.4) LMR 114 (22.3) 47.0 (12.7) 285 (63.5) 0.31 (0.09) 0.60 (0.13) 4.0 (0.54) 52.0 (11.7) 17.8 (3.3) 73.5 (21.1) 7.3 (0.3) -18.2 (1.4) TC 103 (22.7) 69.0 (12.9) 214 (78.8) 0.22 (0.06) 0.32 (0.03) 5.0 (0.69) 47.2 (6.7) 13.7 (2.8) 86.2 (16.7) 9.1 (0.2) -20.2 (0.9) BR 93.2 (11.8) 65.8 (9.6) 141.9 (38.8) 0.19 (0.04) 0.26 (0.06) 2.7 (0.54) 36.3 (3.3) 16.5 (2.6) 88.0 (16.4) 13.2 (0.3) -16.8 (2.8) LDC 95.6 (23.2) 58.9 (9.6) 313 (49.1) 0.26 (0.05) 0.47 (0.06) 4.3 (0.37) 54.3 (8.1) 13.4 (1.4) 75.0 (14.4) 8.0 (0.2) -18.2 (0.8) PC 47.7 (15.3) 18.5 (8.8) 38.7 (24.9) 0.21 (0.05) 0.56 (0.07) 5.0 (0.42) 43.5 (5.1) 14.9 (3.7) 65.1 (18.9) 9.4 (0.3) -18.8 (0.8) PR 50.6 (29.1) 21.5 (8.0) 90.3 (46.9) 0.11 (0.08) 0.50 (0.09) 3.6 (0.33) 33.4 (3.8) 9.1 (0.7) 57.8 (17.2) 11.3 (0.6) -23.9 (1.6) Terrestrial Detritus SBC ------0.08 (0.02) 1.35 (0.03) 49.3 (0.3) 31.4 (0.5) 256.1 (31.7) -1.0 (0.1) -31.1 (0.1) LBC ------0.11 (0.01) 0.92 (0.03) 51.3 (0.4) 47.1 (1.6) 155.4 (2.3) 0.9 (0.2) -29.2 (0.1) VR ------0.10 (0.01) 1.09 (0.05) 52.1 (0.01 41.2 (2.1) 206.5 (7.4) 1.7 (0.1) -29.2 (0.2) NC ------0.19 (0.03) 1.78 (0.07) 50.9 (0.4) 11.4 (0.1) 127.7 (27.2) 6.1 (0.01) -28.4 (0.01) EFWOC ------0.11 (0.01) 1.31 (0.02) 48.6 (1.0) 31.5 (0.2) 175.2 (2.1) 2.1 (0.1) -30.1 (0.2) EFLMR ------0.13 (0.01) 1.48 (0.02) 50.7 (0.9) 29.8 (0.6) 135.2 (2.5) 3.9 (0.1) -30.4 (0.1) LMR ------0.13 (0.03) 1.32 (0.01) 49.6 (1.7) 33.6 (0.3) 190.3 (15.7) 2.0 (0.01) -28.4 (0.1) TC ------0.15 (0.03) 1.74 (0.03) 50.3 (0.9) 24.5 (0.4) 128.9 (20.3) 3.2 (0.1) -30.6 (0.1) BR ------0.13 (0.01) 1.66 (0.16) 49.6 (0.8) 25.6 (2.6) 138.9 (13.7) 4.8 (0.2) -29.6 (0.1) LDC ------0.14 (0.03) 2.10 (0.02) 51.7 (0.3) 21.1 (0.3) 154.6 (39.0) 3.3 (0.1) -29.2 (0.1) PC ------0.14 (0.03) 1.38 (0.05) 51.0 (0.7) 24.3 (0.9) 98.1 (13.0) 3.3 (0.4) -29.4 (0.1) PR ------0.14 (0.03) 3.42 (0.07) 46.7 (1.8) 32.2 (1.3) 119.4 (5.3) 5.1 (0.1) -32.3 (0.1)

166

Table 5.4. Top model regression statistics for water column total phosphorus (TP, µg/L),

15 15 leaf δ N and periphyton δ N. TP data were log10 -transformed. Agriculture and developed land use percentage data were arcsine square-root transformed.

2 Response Predictor Intercept Slope df F adj. r p TP Agriculture 0.5339 1.127 10 8.145 0.3938 < 0.05 TP Developed 1.3661 1.025 10 0.403 -0.05739 0.5398 15 TP Peri δ N 0.30185 0.1488 10 24.97 0.6854 < 0.001 15 Peri δ N Agriculture 4.315 4.843 10 3.466 0.2831 < 0.05 15 Leaf δ N Agriculture -3.283 6.127 10 9.577 0.4381 < 0.05

167

Table 5.5. Mean (SE) δ13C isotope ratios of consumer groups for each study stream.

Stream are ordered by increasing agricultural land use and abbreviations are in Table 5.1.

Filter Stream Collector Omnivore Feeder Grazer Shredder Predator Psephenid Corbiculid -27.9 -27.4 -31.2 -28.1 -27.9 -30.1 -26.7 SBC (-) (0.2) (0.9) (0.6) -- (0.5) (-) (0.85) -27.0 -25.8 -29.0 -27.0 -26.6 -29.5 LBC (0.3) (0.2) (0.7) -- (0.3) (-) (1.3) -- -27.8 -28.5 -31.1 -28.4 -35.2 -31.1 -32.0 VR (0.1) (0.9) (0.3) (0.2) (0.1) -- (1.2) (1.5) -25.1 -25.2 -25.4 -28.1 -27.2 NC (0.5) (0.4) (0.9) (0.2) -- -- (-) -- -26.8 -28.4 -27.8 -26.7 -34.6 -27.9 EFWOC (1.5) -- (0.8) (0.6) -26.8 (-) (-) (0.6) (4.1) -26.9 -29.5 -28.0 -27.3 EFLMR (0.4) -- (0.7) (1.6) ------(2.1) -26.5 -28.3 -30.1 -28.9 -31.1 -28.0 -32.9 -28.2 LMR (1.1) (1.0) (0.5) (0.6) (0.1 (-) (2.0) (10) -28.1 -27.3 -28.6 -28.3 -29.7 -29.8 -30.1 TC (2.0) (-) (0.3) (0.4) (1.3) (0.5) (2.1) -- -25.5 -25.7 -26.8 -25.4 -28.7 -28.4 -27.1 BR (0.6) (0.3) (0.4) (0.3) (0.3) -- (0.8) (1.1) -27.9 -28.5 -31.3 -28.8 -27.2 -29.2 -32.3 -27.7 LDC (0.8) (0.4) (1.1) (0.7) (6.0) (-) (0.7) (1.1) -25.0 -25.5 -26.4 -25.8 -28.0 -27.3 -25.4 PC (0.7) (0.4) (0.2) (0.1) (0.2) -- (0.7) (0.60) -24.4 -25.8 -27.0 -25.5 -27.0 PR (0.7) (0.3) (0.2) -- (0.1) (-) -- --

168

Table 5.6. Mean (SD) δ 15N isotope ratios of consumer groups for each study stream.

Streams are ordered by increasing agricultural land use and abbreviations are in Table

5.1.

Filter Stream Collector Omnivore Grazer Shredder Predator Psephenid Corbiculid Feeder SBC 7.2 (-) 9.3 (0.1) 2.8 (0.3) 5.7 (0.2) -- 8.4 (0.3) 8.0 (-) 7.1 (0.29) LBC 10.5 (1.1) 5.5 (0.2) 6.3 (0.3) -- 1.9 1.0 (-) 10.9 (1.1) -- VR 10.8 (0.2) 13.2 (0.1) 5.3 (0.1) 13.0 (0.4) 3.5 -- 12.2 (0.2) 10.6 (0.49) NC 11.9 (0.4) 14.1 (0.3) 12.4 (0.1) 5.6 (0.1) -- -- 10.0 (-) -- EFWOC 6.6 (0.9) -- 3.1 (2.6) 8.6 (0.5) 1.0 1.0 (-) 7.6 (0.6) 8.8 (1.0) EFLMR 11.6 (0.7) -- 11.9 (0.3) 6.5 (0.9) ------11.3 (0.88) LMR 9.1 (1.9) 4.2 (0.3) 4.0 (0.8) 10.8 (0.7) 3.2 (1.4) 3.2 (-) 7.9 (0.4) 9.9 (2.9) TC 11.0 (2.0) 12.8 (-) 4.4 (0.5) 6.6 (0.1) 3.1 (2.5) 12.6 (0.7) 9.0 (0.6) -- BR 13.3 (0.4) 8.6 (0.9) 8.5 (0.2) 14.8 (0.3) 11.7 (0.2) -- 15.2 (0.1) 12.3 (0.19) LDC 10.7 (1.0) 5.1 (0.3) 9.7 (0.9) 11.1 (0.3) 2.7 (0.3) 12.5 (-) 8.4 (1.5) 9.7 (0.32) PC 12.1 (0.7) 6.3 (0.4) 7.6 (0.3) 13.1 (0.1) 3.3 -- 4.8 (0.4) 11.9 (0.11) PR 13.9 (1.2) 6.0 (0.2) 3.6 (-) -- 3.2 1.0 (-) -- --

169

Table 5.7. ANCOVA results of periphyton and leaf δ15N as predictors of consumer δ15N.

Both models were significant at 0.05 level. Significant p-values in bold.

Predictor Organism FFG Estimate SE t value p

Periphyton δ15N Hydropsychid Filter Feeder -4.0939 1.1884 -3.445 0.000976 Snail Grazer -1.6758 1.2518 -1.339 0.015065 Corbiculid Clam -0.9035 1.3314 -0.679 0.499638 Crayfish Omnivore -2.1354 1.2465 -1.713 0.041177 Psephenid Psephenid -1.3438 1.2464 -1.078 0.284712 Tipulid Shredder -6.768 1.2844 -5.269 1.48E-06 Chironomid Collector 0.2334 0.2007 1.163 0.272007

15 Leaf δ N Hydropsychid Filter Feeder -4.09392 1.3458 -3.042 0.00332 Snail Grazer -1.08265 1.41894 -0.763 0.44806 Corbiculid Clam -0.42149 1.51801 -0.278 0.78211 Crayfish Omnivore -2.22047 1.41174 -1.573 0.12033 Psephenid Psephenid -1.27049 1.41469 -0.898 0.37227 Tipulid Shredder -6.99353 1.45431 -4.809 8.58E-06 Chironomid Collector 0.4831 0.1539 3.139 0.0105

170

)" Scioto Brush Creek " Little Miami River " Little Beaver Creek )" Twin Creek )" Vermilion River )" Blanchard River )" Nimishillen Creek )" Little Darby Creek )" )" East Fork White Oak Creek Paint Creek 010 20 40 60 )" East Fork Little Miami River )" Middle Branch Portage River Kilometers

Lake Erie

)"

)"

)" )" )"

)" )" )"

)"

)" )" )"

Figure 5.1. Map of sampling locations and land use from the National Land Cover Database (2011). Land use categories: crop agriculture (brown), pasture (yellow), forest (green), developed (red), water (blue). Stream abbreviations: TC = Twin Creek, PR = Middle Branch Portage River, NC = Nimishillen Creek, EFWOC = East Fork White Oak Creek, BR = Blanchard River, LDC = Little Darby Creek, SBC = Scioto Brush Creek, PC = Paint Creek, LBC = Little Beaver Creek, EFLMR = East Fork Little Miami River, LMR = Little Miami River, VR = Vermilion River.

171

N

Chl P

AFDM

NP PC2 (28%)PC2 3- 10123 2 1 0 -1 -2 -3

-2 -1 0 1 2

PC1 (45%)

Figure 5.2. The first two axes of the principal component analysis (PC1, PC2) synthesizing periphyton characteristics for the 12 streams sampled. Each point represents a study stream and are scaled by agricultural land use. Abbreviations: Chl = chlorophyll a in mg/m2, AFDM = ash-free dry mass in g/m2, N = percent nitrogen, P = percent phosphorus, NP = nitrogen: phosphorus ratio.

172

15 a b

8

10

4 N(‰)

5 15 N(‰)  15  Leaf

0 0

-5 0 5 10 15 0 20 70 100 15 Agricultural land use (%) Periphyton N (‰) Figure 5.3. Mean basal resource δ15N as a function of watershed agriculture (%, open symbols = periphyton, closed symbols = leaf), (b) mean leaf and periphyton δ15N signature for the 12 study streams. Bars represent standard error of the mean.

173

Collecter Filter Feeder Grazer Shredder Mussel Omnivore Psephenid

15 15

10 10 N of Consumers (‰) Consumers of N N of Consumers (‰) Consumers of N 15 15

 5  5

0 0

0 5 10 15 0 5 10 15 15 15  N of Periphyton (‰)  N of Periphyton (‰)

Collecter Filter Feeder Grazer Shredder Mussel Omnivore Psephenid

15 15

10 10 N of Consumers (‰) Consumers of N N of Consumers (‰) Consumers of N 15 15

 5

 5

0 0

-5 0 5 10 -5 0 5 10 15 15  N of Leaf (‰)  N of Leaf (‰) Figure 5.4. Consumer δ15N as a function of primary producer (leaf and periphyton) δ15N

(a-d). ANCOVA statistics in Table 5.7.

174

Fungi

MUFA PC2 (25%)PC2 SAFA

HUFA w3 EPA BAFA 15-. 05000510152.0 1.5 1.0 0.5 0.0 -0.5 -1.0 -1.5 -1 0 1 2

PC1 (58%)

Figure 5.5. The first two axes of the principal component analysis (PC1, PC2) synthesizing a subset of the periphyton total lipid profiles for the 12 streams sampled.

Each point represents a study stream and are scaled by agricultural land use (%).

Abbreviations for biomarkers: SAFA = saturated fatty acids, MUFA = monounsaturated fatty acids, HUFA = highly unsaturated fatty acids, PUFA = polyunsaturated fatty acids,

EPA = eicosapentaenoic acid, EFA = essential fatty acids, w6 = omega-6 fatty acids,

BAFA = bacterial fatty acids, and Fungi.

175

10.0

7.5

5.0

2.5 Diatom:Green&Cyanobacteria

0.0

0 20 70 100 Agricultural land use (%)

Figure 5.6. Ratio of diatom fatty acid biomarkers (16:1ω7 and 20:5ω3) to green and cyanobacteria fatty acid biomarkers (18:2ω6 and 18:3ω3) as a function of agricultural land use. Bars represent standard error of the mean.

176

Crayfish Fish Hydropsychid 0.6 Megalopteran Periphyton Snail Tipulid 0.4 w6

0.2 PUFA MUFA BAFA 0.0

NMDS2 SAFA

-0.2 w3

-0.4

w3_w6 -0.6

-0.5 0.0 0.5 1.0 NMDS1

Figure 5.7. First two axes of the non-metric multidimensional scaling (NMDS) ordination of fatty acid compositions of food-web components in stream ecosystems (two- dimensional stress = 0.11). Abbreviations: PUFA = polyunsaturated fatty acids, BAFA = bacterial fatty acids, MUFA = monounsaturated fatty acids, SAFA = saturated fatty acids,

ω3 = ω3 fatty acids, ω6 = ω6 fatty acids, ω3_ω6, ω3:ω6 ratio.

177

FilterFeeder Grazer Omnivore Shredder

0.5

0.0

-0.5 NMDS1 -1.0

-1.5

-2.0

0 5 10 15 15N of Periphyton (‰)

Figure 5.8. Periphyton-corrected NMDS scores as a function of periphyton δ15N.

178