water contaminants of the lake erie watershed

171
Water Contaminants of the Lake Erie Watershed Dissertation Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy in the Graduate School of The Ohio State University By Michael Robert Brooker Graduate Program in Environmental Science The Ohio State University 2018 Dissertation Committee Dr. Paula Mouser, Co-advisor Dr. Jon Witter, Co-advisor Dr. Gil Bohrer Dr. Virginia Rich

Upload: khangminh22

Post on 02-May-2023

0 views

Category:

Documents


0 download

TRANSCRIPT

1

Water Contaminants of the Lake Erie Watershed

Dissertation

Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy

in the Graduate School of The Ohio State University

By

Michael Robert Brooker

Graduate Program in Environmental Science

The Ohio State University

2018

Dissertation Committee

Dr. Paula Mouser, Co-advisor

Dr. Jon Witter, Co-advisor

Dr. Gil Bohrer

Dr. Virginia Rich

2

Copyrighted by

Michael Robert Brooker

2018

ii

Abstract

Streams and rivers act as conduits, transporting pollutants from their sources to

downstream drainage basins. The Lake Erie watershed is dominated by agricultural land

use. As a result, there are many concerns over pollution sourced from upstream

agroecosystems. Among the principle issues in the region, phosphorus and other nutrient

pollutants have been faulted for stimulating and/or supporting the frequency and

magnitude of recurrent harmful algal blooms occurring in the western Lake Erie basin.

Phosphorus pollution originates from a variety of point and nonpoint sources, however

specific estimates of source contributions have proven elusive due to wide variations

between members of the same sources. Better distinguishing between the sources of

pollution, as well as an improved ability to track transport through the watershed is

essential for managing nutrient loads. One promising and new approach to elucidating

source contaminants is the organic phosphorus fraction of dissolved organic matter

(DOM). Point and nonpoint sources may exhibit unique DOM or dissolved organic

phosphorus (DOP) signatures, that allows for the differentiation between sources, either

through signature analysis or the application of marker molecules. Here, electrospray

ionization Fourier-transform mass spectrometry (ESI FT-ICR-MS) was used to analyze

the DOM and DOP signatures from nutrient pollution sources in the Lake Erie watershed.

Three marker compounds were distinct to sources were proposed for use in tracking the

iii

presence of source contamination. From this source signature analysis, differences in

DOM was next evaluated along a mixing profile for a Lake Erie tributary. The ability to

detect DOM formulae upstream to downstream sites was assessed. Compounds detected

in higher abundance upstream were more likely to be detected at downstream locations.

The mass spectra signals of merging branches appeared to be mixed linearly into several

confluence points.

In addition to nutrient sources influencing Lake Erie water quality, there are

concerns over the introduction of antibiotics to the drainage basin from use in regional

agricultural operations. Metals and antibiotics are known to co-select for antibiotic

resistance genes in agroecosystems, suggesting two possible causes in the development of

resistant microbial communities. Here, sediments from agricultural dominated channels

were analyzed for antibiotics, metals, and relevant functional microbial genes. Although

few antibiotics were detected in the sediments, some metals were found at elevated

levels. Antibiotic resistance genes were among the most abundant and diverse set of

genes detected using an environmental microbial functional microarray technology,

GeoChip. Metal homeostasis genes and the intI integrase gene, indicative of the potential

for horizonal gene transfer, were also abundant across samples. These results highlight

the prevalence of antibiotic resistant genes in sediments draining to the Lake Erie

ecosystem, with implications for downstream transport from agricultural sources.

iv

Acknowledgments

Thank you to everyone who helped support me on my way into and through graduate

school. That starts with my wife, Molly, and my family. It has been a long and difficult

path. Without Molly, I doubt that I would have ever considered going back to school. Her

encouragement is the biggest reason this document exists. Her support has gotten me

through the most difficult times. I must also thank all the faculty who guided me as my

advisors, committee members, or for giving me numerous opportunities. Dr. Paula

Mouser was my advisor for three graduate degrees, and without her I may never have

entered graduate school. She has been an inspiration and model for my own career. Dr.

Gil Bohrer and Dr. Jon Witter both have served as co-advisors on at least one of my

degrees. All three of these committee members, and my other committee member Dr.

Virginia Rich, took me under their wings and taught me the foundations of how to

conduct research, present that research, and teach courses. Their feedback on my research

and writing was crucial to my success. I am also grateful for them being understanding

during my moments of crazed panic, of which there were many. I would like to recognize

that all faculty members of my department gave me many opportunities along the way.

There are far too many people who helped me along my way to put into this section here.

Please know that I will always appreciate all you have done for me.

v

Vita

2003................................................................Northwest High School

2007................................................................B.S. Microbiology, The Ohio State

University

2013................................................................M.S. Environmental Science, The Ohio

State University

2011 to present ..............................................Graduate Teach and Research Assistant,

Department of Civil, Environmental, and

Geodetic Engineering, The Ohio State

University

Fields of Study

Major Field: Environmental Science

vi

Table of Contents

Abstract ............................................................................................................................... ii

Acknowledgments.............................................................................................................. iv

Vita ...................................................................................................................................... v

List of Tables ................................................................................................................... viii

List of Figures .................................................................................................................... ix

Preface ................................................................................................................................ xi

Chapter 1: Discrete Organic Phosphorus Signatures are Evident in Pollutant Sources

within a Lake Erie Tributary ............................................................................................... 1

Introduction ..................................................................................................................... 1

Methods........................................................................................................................... 5

Site Description and Sample Collection ..................................................................... 5

Sample Processing ...................................................................................................... 7

ESI FT-ICR-MS Data Analysis .................................................................................. 8

Results ............................................................................................................................. 9

Discussion ..................................................................................................................... 17

Chapter 2: Dissolved Organic Matter Transport and Mixing in the Portage River .......... 22

Introduction ................................................................................................................... 22

Methods......................................................................................................................... 26

Sampling Locations and Collection .......................................................................... 26

Sample Processing .................................................................................................... 28

ESI FT-ICR-MS Data Analysis ................................................................................ 29

Results ........................................................................................................................... 31

Discussion ..................................................................................................................... 45

Chapter 3: The Emerging Concern of Antibiotic Resistance Genes in Agricultural

Sediments .......................................................................................................................... 52

Introduction ................................................................................................................... 52

vii

Methods......................................................................................................................... 55

Site Description and Sample Collection ................................................................... 55

Genomic DNA Extraction......................................................................................... 56

Functional Gene Assays ............................................................................................ 56

16S rRNA Sequencing .............................................................................................. 57

Antibiotic Extraction and Analysis ........................................................................... 58

Metals Analysis ......................................................................................................... 61

Data Analysis ............................................................................................................ 62

Results ........................................................................................................................... 63

Discussion ..................................................................................................................... 82

Conclusions ....................................................................................................................... 88

References ......................................................................................................................... 92

Appendix A: Sandusky Source Material DOM Analysis ............................................... 112

Methods....................................................................................................................... 113

Collection of Mass Spectrometry Data and Peak Detection ................................... 115

Results & Discussion .................................................................................................. 116

Appendix B: Portage River DOM Mixing Analysis ....................................................... 132

Appendix C: Antibiotic Resistance Gene Analysis ........................................................ 143

viii

List of Tables

Table 3.1. Common genes between the GeoChip and Fluidigm platforms ...................... 72

Table 3.2. Shared GeoChip ARG lineages with taxonomies detected by Illumina

sequencing......................................................................................................................... 76

Table 3.3. Concentration of antibiotics in the agricultural sediments .............................. 79

Table 3.4. Concentrations of trace elements extraction from the sediments .................... 81

Table A.1. Adsorption efficiency across samples using the Bond Elut PAX solid phase

extraction resin………………………………………………………………………….122

Table A.2. ESI(-) FT-ICR-MS analysis detected a total of 14637 peaks, spread across the

samples and replicates. .................................................................................................... 123

Table A.3. ESI(-) FT-ICR-MS analysis provided peaks which were assigned formulas

with C/H/O/N/P/S elements ............................................................................................ 124

Table A.4. The Venn counts of Sandusky source material data. .................................... 125

Table A.5. List of potential marker formulas found in source and Sandusky River

samples ............................................................................................................................ 127

Table B.1. StreamStats data obtained from the four confluence sampling locations…..133

Table B.2. Nutrient concentrations and solid phase extraction (SPE) efficiencies of the

Portage River samples..................................................................................................... 134

Table B.3. QA/QC filtering of the data and the number of m/z values remaining in

samples at each step. ....................................................................................................... 135

Table B.4. The distribution of atomic composition of formula identified in each Portage

River sample. .................................................................................................................. 137

Table B.5. The molecular class distribution of formula identified in the Portage River

samples. ........................................................................................................................... 138

Table C.1. Yields and purity of DNA extracts of the sediments collected in 2016…….144

Table C.2. Methodology used in LC separation of antibiotics ....................................... 144

Table C. 3. Methodology used in LC separation of antibiotics ...................................... 145

Table C.4. Gene probe normalized signals for 99.9th percentile of detected values in the

GeoChip analysis on the sediments collected in 2016. ................................................... 147

Table C.5. Gene probe normalized signals for 99.9th percentile of detected values in the

GeoChip analysis on the sediments collected in 2015. ................................................... 150

Table C.6. Probe counts for the metal homeostasis gene probes. ................................... 153

Table C.7. The functionality of the metal genes detected across both GeoChip datasets..

......................................................................................................................................... 154

Table C.8. Fluidigm readings across samples and replicates .......................................... 155

Table C.9. Sequence reads from Illumina sequencing .................................................... 157

ix

List of Figures

Figure 1.1. Description of sampling location and watershed. ............................................. 6

Figure 1.2. Summary of atomic composition and mass to charge values by samples ...... 11

Figure 1.3. Van Krevelen diagrams of each sample, and the molecular classes of

identified formulae ............................................................................................................ 13

Figure 1.4. Comparison of samples by binary Jaccard distance matrices. ....................... 15

Figure 1.5. Relative peak heights for potential markers for detecting or tracking source-

derived DOP nutrients shared uniquely by the Sandusky River and either the (1) three

manures, (2) WWTP effluent, or (3) edge of field samples. ............................................. 17

Figure 2.1. Sampling locations in the Portage River ........................................................ 28

Figure 2.2. DOC, TDN, and TDP concentrations measured in Portage River samples ... 33

Figure 2.3. Van Krevelen plots of the 16 samples in this study ....................................... 36

Figure 2.4. Summary of elemental composition and molecular classes of assigned

formula .............................................................................................................................. 38

Figure 2.5. Percentage of shared and unique formulae between samples at the confluence

sampling locations ............................................................................................................ 39

Figure 2.6. Clustering analysis of m/z values ................................................................... 41

Figure 2.7. Quantiles of the peak heights for observed m/z values ................................... 43

Figure 2.8. Comparisons between the StreamStats and DOM mixing model contribution

estimates ............................................................................................................................ 45

Figure 3.1. Summary of the gene probe abundance and signals for functional categories

and antibiotic resistance .................................................................................................... 67

Figure 3.2. Dendrograms of the GeoChip and antibiotic resistance gene hierarchal

clustering ........................................................................................................................... 70

Figure 3.3. Fluidigm results and comparison to GeoChip observations ........................... 73

Figure 3.4. Illumina sequencing on the v4 region of the 16S rRNA gene was performed

on sediment DNA. ............................................................................................................ 75

Figure 3.5. Distribution of unmatched taxa between GeoChip lineages and taxonomies

detected by Illumina sequencing. ...................................................................................... 78

Figure A.1. Retention of carbon, nitrogen, and phosphorus by solid phase extraction

columns ………………………………………………………………………………...117

Figure A.2. Recovery of known phosphorus standards .................................................. 119

Figure A.3. Carbon, nitrogen, and phosphorus concentration of samples in Sandusky

River watershed .............................................................................................................. 121

Figure A.4. The distribution of NOSC values by molecular classes. ............................. 122

Figure A.5. Spectra captured from ESI(-) FT-ICR-MS analysis of all sample replicates

and blanks. ...................................................................................................................... 131

Figure B.1. Spectra collected by ESI(-) FT-ICR-MS analysis…………………………136

Figure B.2. Correlations between nitrogen and phosphorus concentrations and elemental

compositions ................................................................................................................... 139

Figure B.3. Hierarchal clustered dendrogram and heatmap based off the Canberra

distance matrix. ............................................................................................................... 140

x

Figure B.4. Hierarchal clustering of the binary Jaccard distance matrix between samples

collected in both Chapter 1 and Chapter 2. ..................................................................... 141

Figure B.5. The relative change in peak heights between upstream-downstream samples

in the upper reaches of the Portage River (A through E.2). ............................................ 142

xi

Preface

Agriculture dominates the Lake Erie watershed, with sources of pollution from

agroecosystems a significant concern. Nutrient pollution, primarily phosphorus and

nitrogen are loaded into Lake Erie with much of the blame on nonpoint – largely

agricultural – sources. There is a need to better understand the contributions of nonpoint

and point sources, including phosphorus loads and other emerging health threats, such as

antibiotic resistance to the region. This dissertation builds on two chapters of research in

my dual-degree Master of Science Thesis in Civil Engineering (completed October 2017)

through the detailed evaluation of pollutants in contaminant sources, sediments, and

waters of Lake Erie tributaries. The dissertation is structured with three stand-alone

chapters that are intended to be submitted directly for peer review in relevant

environmental engineering or science journals. As chapters are intended to meet the word

limit requirement for the intended journal, supplemental information for each chapter (i.e.

detailed methods, raw data, etc.) is provided in a corresponding Appendix. A short

conclusions chapter is provided after Chapter 3 to highlight significant findings and

present future research needs in this field. The following paragraphs summarize the

topics discussed in Chapters 1 through 3.

In Chapter 1, DOM signatures were described for five point and nonpoint sources

of nutrient pollution in a Lake Erie watershed. These signatures were proposed as a

means to detect the presence of phosphorus, or other nutrient pollutants, derived from

specific source materials. Chapter 1 focused on the fraction of organic phosphorus within

the DOM. Divergent DOP signatures were observed between several manures and other

xii

source materials. A high degree of similarity was detected between the Sandusky River

and edge of field runoff (from a synthetically fertilized crop field). Several marker

compounds were proposed for use in detecting and tracking source contributions through

this and other Lake Erie tributaries.

The objective of Chapter 2 was to build on the findings of Chapter 1 through a

broader analysis (in terms of geography and total number of samples) of DOM and DOP

along a second river transect draining to Lake Erie. Samples were collected along 50

linear miles from upstream tributaries through the mouth of the Portage River, with a

mixing analysis conducted at four confluence points. Samples were compared based on

their mass spectral diversity and abundance, the similarity of their organic matter

features, and their overall nutrient concentration. Apparent changes in heteroatom and

molecular classifications were observed between upstream to downstream reaches, with

CHON formulae becoming more abundant closer to the mouth of the river. The most

prominent DOM features persisted from upstream reaches to downstream locations,

suggesting a DOM mixing model analysis may provide a tool for tracking source

contributions along the watershed.

In Chapter 3, a genomic analysis of sediments accumulating in drainage ditches

from agricultural headwaters revealed a diversity and prevalence of antibiotic resistance

genes. Additionally, metal homeostasis genes were abundant. These two gene categories

are commonly co-selected in the environment. Although few antibiotics were detected

that might contribute to the prevalence of antibiotic resistant organisms, many metals

were detected at elevated levels. The data suggests that the ARGs of these sediments is

xiii

maintained through the co-selection of antibiotic resistance genes with metal homeostasis

genes.

1

Chapter 1: Discrete Organic Phosphorus Signatures are Evident in Pollutant Sources

within a Lake Erie Tributary

This chapter was submitted to the journal Environmental Science & Technology on

November 14 under the title: Discrete Organic Phosphorus Signatures are Evident in

Pollutant Sources within a Lake Erie Tributary by Michael R. Brooker, Krista

Longnecker, Elizabeth B. Kujawinski, Mary H. Evert, Paula J. Mouser. It is currently

under review.

Introduction

Freshwater lakes, such as the Great Lakes in North America, provide numerous

economic opportunities to shoreline communities in the form of tourism, recreation,

fisheries, manufacturing, and the transportation of goods across local and international

boundaries. Lake Erie is one of five Great Lakes in the United States and Canada, and as

with many freshwater resources worldwide, it has experienced recurrent harmful algal

blooms that are believed to be propagated from anthropogenically-sourced nutrients from

within its drainage basins (Conley et al., 2009; Conroy et al., 2011). Primary productivity

in freshwater systems is most often limited by phosphorus or nitrogen (Conley et al.,

2009; Conroy et al., 2011), therefore changes in the abundance and form of these

nutrients from upstream sources can have a profound effect on the ecosystem. Since the

early 2000s, increased nutrient loads have led to recurrent toxic cyanobacterial blooms

along the southern coastline of the western Lake Erie basin, while hypoxia has developed

in the hypolimnion of the central basin in the lake (Conroy et al., 2011; Michalak et al.,

2013; Steffen et al., 2017).

2

The magnitude of Lake Erie algae blooms in a given year is most strongly

correlated to spring (May-June) phosphorus loads from its tributaries (Stumpf et al.,

2012), with small blooms sometimes inflicting severe damage to the ecosystem. For

example, although it was smaller in size compared to years past, the Microcystis bloom at

the Toledo water treatment facility intake pipe in 2014 had a major impact on the

shoreline community (Steffen et al., 2017). Microcystin concentrations in the treated

water were twice as high as state guidelines (currently 1.6 μg/l in Ohio), causing the

shutdown of the drinking water treatment plant serving over 400,000 Toledo residents

and resulting in $65 million in economic damages to property values, tourism, recreation,

and emergency water handling (Steffen et al., 2017). The impact of this and other

phytoplankton blooms on the local economy has served as a call-to-action for Ohio

legislators to improve our understanding of nutrient pollutants contributing to this

problem and develop best management practices to minimize discharge.

Phosphorus pollution in drainage basins is derived from both point (e.g.,

municipal/industrial wastewater effluents or combined sewer overflows) and nonpoint

sources (sewage leaks, urban area runoff, or agricultural runoff/tile drainage) (D. B.

Baker et al., 2014; Ohio Lake Erie Phosphorus Task Force, 2013), making individual

pollutant sources difficult to isolate and manage. An extensive sampling network has

been established in select Lake Erie tributaries to monitor loads to the lake (D. B. Baker

et al., 2014), with an emphasis on reactive and total phosphorus. Reactive phosphorus, a

term used interchangeably with orthophosphate (PO43-), is readily assimilated by algae

and simple to measure (D. B. Baker et al., 2014; Baldwin, 1998). Total phosphorus has

3

been useful in forecasting harmful algal bloom severity (Stumpf et al., 2012). Models

have helped fill in the gaps between discrete sampling locations by considering local land

usage to estimate spatial contributions (D. Baker, 2011; Michalak et al., 2013). However,

despite efforts made toward monitoring and modeling source contributions to Lake Erie,

distinguishing between specific pollutant sources to mitigate the most impactful loads to

the lake has proven difficult.

In order to gain further insight into pollutant sources and phosphorus pool

dynamics, researchers are applying new mass spectrometry tools. One method, analysis

of oxygen isotopic fractionation, has allowed for partial source tracking of phosphate

entering Lake Erie from its tributaries (Elsbury et al., 2009). Isotopic fractionation arises,

in part, from the enzymatic hydrolysis of dissolved organic phosphorus (DOP). The

results of this isotopic analysis in Lake Erie suggested a non-riverine source of phosphate

was supplying the algal bloom, but could not establish whether DOP was the source of

this phosphorus (Elsbury et al., 2009). DOP is rarely analyzed on environmental samples

because (1) concentrations are low and are indirectly quantified (Baldwin, 1998;

Monaghan, E. J., Ruttenberg,K.C., 1999; Ruttenberg KC, 2012), and (2) the low

elemental abundance of phosphorus within dissolved organic matter (DOM) makes

detection using mass spectrometry difficult (Cooper et al., 2005; D. M. Karl, 2014; Kruse

et al., 2015). Analysis of DOP has been disregarded in favor of measuring total dissolved

phosphorus (TDP), as TDP effectively defines bioavailable phosphorus (Ohio Lake Erie

Phosphorus Task Force, 2013; Ruttenberg and Dyhrman, 2005). However, TDP obscures

the diversity of DOP formulae elucidated through mass spectrometry (Cooper et al.,

4

2005; Minor et al., 2012) which may aid in source identification and provide a better

understanding of biogeochemical controls in the system.

Electrospray ionization Fourier-transform ion cyclotron resonance mass

spectrometry (ESI FT-ICR-MS) can provide new insight into the molecular composition

of environmental samples through non-target identification of phosphorus in dissolved

organic matter (DOM). To date, ESI FT-ICR-MS has rarely been used to investigate

DOP, and, in some cases, phosphorus has been excluded from these (Kujawinski and

Behn, 2006; Kujawinski et al., 2009) due in part to a low elemental abundance of

phosphorus (~0.3%) in organic matter. However, organophosphorus compounds can be

concentrated for mass spectrometry analysis with solid phase extraction (SPE), which

removes background interferences (i.e., desalts) while retaining organic constituents that

resemble the original sample (Ohno and Ohno, 2013; Raeke et al., 2016). Even in the

absence of selective concentration, ESI FT-ICR-MS analysis revealed an abundance of

organic phosphorus-containing compounds in Lake Superior and its tributaries (Minor et

al., 2012). Based on the frequency of harmful algal blooms, we expected Lake Erie would

be replete in unique organic phosphorus compounds that could be related to tributary

sources.

The objective of this study was to characterize organic-bound phosphorus from

select point and nonpoint pollutant sources in a Lake Erie tributary. We analyzed organic

matter and organophosphorus signatures in three different nonpoint source fertilizer

materials (hog, chicken, and dairy manures), runoff from the edge of a synthetically

fertilized agricultural field, a point source discharge location from a municipal

5

wastewater treatment plant (WWTP), and the Sandusky River using ultrahigh resolution

ESI FT-ICR-MS. Molecular masses, molecular classes, sample similarity and unique

marker formulae were identified in these samples. Our analysis identified a diverse

organic phosphorus pool that is obscured by the single phosphorus measurement typically

used to represent these sources. These data provide signatures of pollutant that can used

to monitor their movement through tributaries, and gives consideration to the

understudied pool of organic phosphorus.

Methods

Site Description and Sample Collection

Sampling was performed in the Sandusky River tributary system, which drains

into the Western Lake Erie Basin at Sandusky Bay. The Sandusky River is dominated by

nonpoint phosphorus pollution (90%) with smaller contributions from point (9%) and

atmospheric (1%) sources (Figure 1.1A) (Ohio Lake Erie Phosphorus Task Force, 2010).

The primary land use in the watershed is agricultural, with the vast majority of fertilizer

application derived from inorganic (66%) forms rather than manure (27%) or biosolids

(7%) (Ohio Lake Erie Phosphorus Task Force, 2010). Most of the manure applied in the

Lake Erie basin originates from cattle (50%), hog (34%), and poultry (5%) sources

(Figure 1.1B) (Ohio Lake Erie Phosphorus Task Force, 2010). Sampling was conducted

on March 14, 2016 following a precipitation event (Figure 1.1C). At the time of

collection, flows were high (>90th percentile) and corresponded with a high total

phosphorus load (www.heidelberg.edu/NCWQR) (D. B. Baker et al., 2014). Six samples

were collected from the Sandusky River tributary network (Figure1.1D), including (1) an

6

edge of field site, (2) hog manure, (3) chicken (poultry) manure, (4) dairy (cattle) manure,

and (5) wastewater treatment plant (WWTP) effluent. Downstream of these sampling

locations, another sample was collected from the (6) Sandusky River.

Figure 1.1. Description of sampling location and watershed. (A) The Ohio Lake Erie

Phosphorus Task Force has estimated the nonpoint contribution from point and nonpoint

sources for the Sandusky River (Ohio Lake Erie Phosphorus Task Force, 2010). (B) This

group has also detailed the contributions of various manures, as elemental P, to the

Western Lake Erie basin (Ohio Lake Erie Phosphorus Task Force, 2010). (C) Flow, total

phosphorus, and soluble reactive phosphorus were reported in the 2015-2016 water year

by Heidelberg University (www.heidelberg.edu/NCWQR) (D. B. Baker et al., 2014). The

arrow shows the flow conditions at the time of sampling. (D) Six samples were collected

from the Sandusky River watershed situated in north-central Ohio. The chicken and hog

samples were collected on the same property.

Sampling equipment was pre-conditioned by triple rinsing sampling devices and

storage containers with Milli-Q water. The chicken manure sample was retrieved from

the center of an open-air stockpile following excavation by landowner equipment. The

7

hog manure sample was sampled from a hog manure pit using a PVC sampling device.

Dairy manure was collected from a secondary lagoon using the PVC sampling device. An

edge of field sample was collected from the mouth of a tile drainage pipe flowing into the

connected stream. Wastewater effluent was collected following chlorination but prior to

discharge from the Tiffin Water Pollution Control Center. Finally, the Sandusky River

sample was collected from the faucet of the USGS station (USGS 04198000). All

samples were collected in pre-rinsed (DI water) polyethylene containers, transported on

ice to the OSU Environmental Biotechnology Laboratory, and held at 4°C. Wet samples

were processed within 24 hours.

Sample Processing

The dry weight of manure samples was determined by weighing subsamples into

porcelain dishes and heating for 24 hours at 70°C. Following the dry weight

determination, duplicate manure samples were suspended at equivalent ratios of water to

dry weight ratios (15:1) using Milli-Q water (Ohno et al., 2016). The manure-water

mixtures were equilibrated overnight at 4°C. Combusted glassware (30 min at 500°C)

was used for the remainder of sample preparation. All samples were vacuum-filtered

through pre-rinsed (methanol and DI water) 0.7-µm glass fiber filters (Whatman GF/F).

The concentrations of dissolved organic carbon (NPOC) and nitrogen (TDN) were

determined using a Shimadzu TOC-V/TNM-1 analyzer. Phosphorus (TDP)

concentrations were measured using an Agilent ICP-AES (Figure A.3). Samples were run

as previously described for NPOC/TDN (Kekacs et al., 2015), while TDP was measured

8

at wavelength 213.648nm (Bartos et al., 2014). Each sample was prepared in duplicate at

a concentration of 6.5 mg L-1 NPOC in preparation for solid-phase extraction.

We previously determined that the Plexa-PAX solid phase extraction columns

were most efficient at retaining organic phosphorus compounds used as laboratory

standards (Appendix A). Thus, Plexa-PAX SPE columns were used for the concentration

of DOM. The 6 samples we collected were prepared in duplicate along with two

reference standards (Pony Lake Fulvic Acid [PLFA], Suwanee River Fulvic Acid

[SRFA]) for a total of 14 samples. Briefly, columns were prepared by wetting with 3 mL

100% HPLC grade methanol, and were then rinsed with 2L DI water. While still wet,

275mL of each sample were gravity filtered through the conditioned SPE columns to

collect and concentration the organic contents. The binding efficiency of samples (C/N/P)

was calculated from the concentrations measured before and after SPE filtration (Table

A.1).

Samples were eluted from the columns with 5mL of HPLC grade methanol,

followed by 5 mL of methanol+ 5% formic acid. These elutions were combined into

amber glassware and stored at -20°C. The samples were shipped on dry ice to the Woods

Hole Oceanographic Institution for ESI (-) FT-ICR MS analysis.

ESI FT-ICR-MS Data Analysis

Mass spectrometry data was collected as previously described (Appendix A)

(Minor et al., 2012). Peaks were detected in the range of 200-1000 Da. Molecular

formulae assignments were made with the Compound Identification Algorithm

(Kujawinski and Behn, 2006; Kujawinski et al., 2009). A total of 14,637 unique peaks

9

were detected under this analysis. Quality controls were used to quality filter the dataset:

peaks observed in DI water or solvent blanks, and singletons were removed (Table A.2).

Only m/z values with an assigned formula were considered for further analysis.

Additional data analyses were performed using R Statistics (version 3.1.1). The

distributions of peak heights and m/z values were compared among replicates and

samples. Then the peak heights were normalized to the sum of peaks for each replicate.

Replicates were combined through averaging of these normalized peak heights. Sample

similarity was compared based on presence/absence of the formulae using Venn Euler

diagrams, and based on relative peak heights using a Bray-Curtis dissimilarity matrix

generated by the ‘vegan’ package (Oksanen et al., 2015). A list of organic phosphorus

formulae, shared between the Sandusky River and at least one source material, were

filtered as a subset from the data. Putative tracers were further screened from this list with

the stipulation that the maximum relative peak height was observed in a source sample.

Results

The amount of carbon, nitrogen, and phosphorus varied across samples (Figure

A.3), with manure-extracted DOM having considerably higher concentrations relative to

other aqueous samples. The manure samples had nutrient concentrations in the range of

30-76 mg C L-1, 12-60 mg N L-1, and 4.8-9.6 mg P L-1 as compared to WWTP effluent,

edge of field, and Sandusky River samples (6.5-8.8 mg C L-1, 2.6-11mg N L-1, and <0.03-

0.09 mg P L-1). The influent and effluent concentrations for these samples were measured

to estimate the amounts that were retained by PAX columns. PAX extraction efficiency

varied considerably, with 8-44% C, 6-41%N, and <0-100% P retained by the columns

10

(Table A.1). At the low P concentrations observed in several of our samples, extraction

efficiencies were near our analytical quantitation limits and reported as estimates.

ESI FT-ICR-MS analysis was used to characterize the molecular properties of

organic matter isolated from the six sample materials. A total of 7250 formulae were

identified in the dataset following quality filtering and formulae assignment (Table A.2).

Reproducibility between replicates were generally high (>80% shared formulae) for all

six samples (Table A.2), which allowed for a combination of replicate data by averaging

the normalized signal between duplicates as well as include any detected formulae in the

final representative sample (Table A.3). The number of identified formulae ranged from

1803 to 4522 across the samples (Figure 1.2A and Table A.3). Within these data, the

number of formulae containing a P atom ranged from 132 to 313 for these six samples,

representing between 3.3% and 12.8% of detected formulae (Figure 1.2B). The manure

samples contained the greatest proportion of DOP formulae (10.9% to 12.8%), double to

triple what was detected in the edge of field, WWTP effluent, and Sandusky River

samples (3.3% to 5.3%). Manure samples also had a greater abundance of formulae with

N or S atoms compared to the other samples, with CHON representing 30-40% of the

manure formulae versus 16-21% in the other three samples.

11

Figure 1.2. Summary of atomic composition and mass to charge values by samples. (A)

The number of assigned formulae representing DOM (full bar) and DOP (red bar) varied

across the six watershed samples. Actual values are printed within their respective bars,

with the total noted above. Note that any formulae containing a P atom was considered to

be DOP. (B) The proportional distribution of major atom classes for each sample shown

with pie charts, with percentages indicating the total proportion of DOP. The distribution

of (C) DOM and (D) DOP m/z values were visualized using kernel-based cumulative

density plots (violin plots). The width of each band indicates the kernel-based density of

m/z values relative to total number, with white bands representing sample quartiles.

12

The manure samples were composed of a greater number of low molecular mass

formulae compared to the other samples (Figure 1.2C). Specifically, the median

molecular mass of observed m/z values for hog (420 Da), chicken (387 Da), and dairy

(419 Da) manures were on average 70 Da lower than that of the WWTP effluent (474

Da), edge of field (485 Da) and Sandusky River (481 Da). DOP formulae generally

followed this trend, with the chicken (341 Da) and dairy (355 Da) manures having a

lower median mass than the WWTP effluent (414 Da), edge of field (412 Da), and

Sandusky River (404 Da) samples (Figure 1.2C). The hog manure sample had the highest

median molecular mass of DOP m/z values (424 Da). Furthermore, unlike the other 5

samples, which were shifted toward lower molecular mass of DOP relative to DOM, the

hog manure DOP molecular mass distribution resembled that of its overall DOM.

Sample signatures were visualized using Van Krevelen diagrams, which relate the

C:H to the O:C molar ratios for all observed formulae (Figure 1.3A). The relative

placement of each formulae provides an estimation of molecular class, which we refer to

as “–like” types. The overall scatter had apparent differences, with manure samples

exhibiting a greater diversity of molecular type classes (i.e., more scatter) compared to

the Sandusky River, edge of field, and WWTP effluent samples which more tightly

clustered around the lignin-like features. To further highlight these differences in overall

scatter, we tallied the relative abundance of formulae in each of 7 different molecular

classes (Figure 1.3B). Although the majority of formulae across all samples were lignin-

like (51-80%), the WWTP effluent, edge of field, and Sandusky River samples were

especially dominated by lignin- and tannin-like features (86-88%) compared to manures.

13

In contrast, the three manure samples consisted of a higher proportion of most other

molecular classes, notably protein-, lipid, and carbohydrate-like features.

Figure 1.3. Van Krevelen diagrams of each sample, and the molecular classes of

identified formulae. (A) Van Krevelen diagrams showing the molar ratio of

hydrogen:carbon versus oxygen:carbon for each assigned formula, color-coded based on

atomic composition. Lipid-, protein-, carbohydrate-, unsaturated hydrocarbon-, lignin-,

tannin-, and uncondensed hydrocarbon-like molecular class ranges are represented by

boxes. (B) The relative abundance of molecular classes was summarized for each sample.

To further probe sample similarity, we compared DOM and DOP signatures based

upon the abundance of shared formulae using Euler diagrams (Figure 1.4A and 1.4B,

14

respectively). Although they were collected from 12 to 41 miles apart, the WWTP

effluent, edge of field, and Sandusky River samples shared 54% of all assigned DOM

formulae. When we calculated the intersection between the Sandusky River and either the

WWTP effluent or the edge of field sample, over 84% of assigned formulae were shared

for both data sets. This level of similarity is comparable to our replicates (81-90% shared

formulae, Table A.2). Interestingly, the edge of field sample shared a considerably

greater number (124) and percentage (75%) of DOP formulae with the Sandusky River as

compared to the WWTP effluent sample (98, 59%).

15

Figure 1.4. Comparison of samples by binary Jaccard distance matrices. Sample

similarity based on presence/absence data visualized using Euler diagrams for (A) DOM

and (B) DOP. The centroid is marked by a small circle with numbers indicating the

number of formulae shared within an intersection. Not all numbers are indicated but may

be found in Table A.4. The number of unique formulae for each sample is color-coded

and placed adjacent to that sample’s ring. Hierarchal clustering dendrograms for (C)

DOM and (D) DOP prepared from Bray-Curtis dissimilarity matrices generated from

relative peak heights for assigned formulae. Numbers along the top reflect the level of

dissimilarity between samples at the branch point.

Across the three manure samples, only 33% of the m/z values were shared with

the Sandusky River. In pairwise comparisons, the intersection between the Sandusky

River and individual manures ranged from 31-36%. Moreover, these shared formulae

were not unique among manures and the Sandusky River; all but one was also present in

16

the WWTP effluent and edge of field runoff samples. Only five m/z values were uniquely

shared between the manures and Sandusky River sample.

We expanded our analysis to also consider the similarity of two NOM standards

(PLFA and SRFA) based on relative peak heights of all assigned formulae using

hierarchal clustering analysis to visualize relationships between all eight samples (Figure

1.4C-D). Dendrogram clustering patterns for DOM (Figure 1.4C) reaffirmed the

similarities between the Sandusky River, edge of field, and WWTP effluent samples also

based on presence/absence analysis (Figure 1.4A). These three samples and the NOM

standards formed a separate branch from the manure samples, which exhibited greater

dissimilarity between one another and the rest of the samples (Figure 1.4C). When

considering only DOP formulae, dissimilarity grew between all samples, although the

two major branches remained the same (Figure 1.4D). We found that the SRFA clustered

among our samples for DOM, yet when we only considered DOP, both SRFA and PLFA

were separated from the WWTP effluent, edge of field, and Sandusky River samples.

These NOM standards are primarily derived from a terrestrial origin, with the Pony Lake

(Antarctica) more geographically remote and less anthropogenically-impacted then the

Suwanee River (Georgia, USA).

In an effort to identify phosphorus formulae originating from our point source

(WWTP effluent) and nonpoint sources (all others) in the Sandusky River, we generated

a list of DOP formulae present in at least one sample and the Sandusky River. The list

was further screened to remove formulae that increased in peak abundance from the

source to the Sandusky River, as this could indicate origination of these m/z values within

17

the river. Our filtering resulted in 72 formulae, which we propose could serve as markers

for detecting or tracking source-derived nutrients (Table A.5). We next identified

formulae from this list that were unique to the (1) edge of field, (2) WWTP effluent, or

(3) the three manure samples (Figure 1.5). The relative peak height for the manure

marker was an order of magnitude higher than was observed in the Sandusky River,

while peak heights for edge of field and WWTP effluent markers were comparable

between the source and Sandusky River sample.

Figure 1.5. Relative peak heights for potential markers for detecting or tracking source-

derived DOP nutrients shared uniquely by the Sandusky River and either the (1) three

manures, (2) WWTP effluent, or (3) edge of field samples.

Discussion

Regulatory agencies and research institutions in the Great Lakes region are

collectively working to understand the sources and sinks of nutrient pollution associated

with eutrophication-induced hypoxia and reduce the recurrence of harmful algal blooms

through nutrient management strategies. In Lake Erie tributaries where land use is

dominated by agriculture, the majority of phosphorus is thought to be derived from

inorganic fertilizer applied to fields (D. B. Baker et al., 2014; Ohio Lake Erie Phosphorus

18

Task Force, 2010; Ohio Lake Erie Phosphorus Task Force, 2013). However, this finding

relies upon models which have considered bulk phosphorus analyses of total or dissolved

reactive P, measurements that cannot be used to discriminate between point and nonpoint

pollution sources within the watershed. Our ultrahigh resolution MS analysis showed that

DOM and DOP signatures collected from drainage tiles at the edge of an agricultural

field in the Sandusky River were highly similar (84% DOM, 75% DOP) to that of the

river itself collected 41 miles downstream. This level of similarity is remarkable

considering Sandusky River replicates shared 85% of m/z values. Closer in hydrologic

proximity (12 miles between sampling locations), the Sandusky River and WWTP

effluent sample also had similar DOM (84% shared formulae), but were more dissimilar

in their DOP (59% shared m/z values). It is notable that the edge of field and Sandusky

River are most alike in their DOP character, as this finding is consistent with the type of

nutrient pollution, primary land use, and fertilizer form previously reported for the

Sandusky River (Ohio Lake Erie Phosphorus Task Force, 2010).

The signatures of the three manure samples were vastly different from all of the

other samples. Manures account for 27% of total P applied as fertilizer to agricultural

systems for the Lake Erie basin (Ohio Lake Erie Phosphorus Task Force, 2010) serving

as a rich source of natural fertilizer despite challenges associated with their handling. Our

analysis shows manure samples are abundant in N- (> 30%) and P- (>10%) containing

organic molecules that are easily liberated from the solids by water. The DOM that was

extracted from these manures in our labs had higher relative phosphorus and nitrogen

concentrations than our other samples, and this can likely explain the high abundance of

19

DOP formulae. Manure DOM also consisted of lower molecular mass m/z values, relative

to the other samples, which may represent more labile compounds that are easily

assimilated into the landscape (Ohno et al., 2007). Future studies should consider the

signatures associated with manure-applied field runoff.

DOP and DOM signatures from point and nonpoint sources would be altered by

abiotic (i.e., photodegradation) or biotic (i.e., biodegradation) processes in soils,

groundwater and surface waters as it moves through the watershed. In particular, the

transport of biosolids and manure derived organic compounds through porous media into

the water column could be retarded by adsorption to solid materials (Dodd and Sharpley,

2015; Sharma et al., 2017). Sorption affinity of phosphorus is specific to each compound

and is also affected by soil type (Berg and Joern, 2006) therefore we would expect

hydrophilic compounds to be more prominent in manure-derived runoff. Although the

molecular masses in our analysis provide little information about hydrophobicity, the

extraction method we used to obtain our manure DOM was likely to have selected for the

more hydrophilic compounds

Inorganic phosphate can be readily assimilated by most plants and organisms,

while organic phosphorus requires enzymatic cleavage (D. M. Karl, 2014). Natural

organic phosphorus exists in the P(V) (organophosphates) or P(III) state

(organophosphonates), with the latter requiring enzymatic oxidation to phosphate

following liberation of the phosphonate groups (D. M. Karl, 2014; Pasek et al., 2014).

Conversely, organophosphates are directly hydrolyzed into inorganic phosphate by

enzymes such as alkaline phosphatase (D. M. Karl, 2014; Ruttenberg and Dyhrman,

20

2005). Organophosphonates therefore require a greater investment of activation energy

that has been found to slow microbial growth, leading to a buildup of these compounds in

natural systems (Adams et al., 2008; D. M. Karl, 2014). Organophosphorus can be

utilized concurrently with inorganic P, but its nutrient value is greater when total

phosphorus supplies are limited (Bjorkman and Karl, 2003; Ruttenberg and Dyhrman,

2005). The Lake Erie tributary network has relatively high phosphate concentrations

compared to other aquatic systems suggesting organic phosphorus turnover will be

slower relative to inorganic forms in its rivers and streams (D. Baker, 2011; D. B. Baker

et al., 2014). Whether these compounds persist and accumulate in the lake remains to be

seen.

Certain organic molecules in these samples are more susceptible to chemical

transformations and would be more readily assimilated by microorganisms. The WWTP

effluent sample was enriched with microbial-derived (e.g., lipid- and protein-like)

features from the activated sludge process while the edge of field and Sandusky River,

like the SRFA standard, were greatly dominated by lignin. Tannin- and lignin-like

features are regarded to have a terrestrial (plant-derived) origin compared to protein-,

carbohydrate-, and lipid-like features which instead originate from endogenous

microorganisms (Feng et al., 2016; Minor et al., 2012). In addition to indicating the

source material, these molecular classes also correlate to the nominal oxidation state of

carbon (NOSC, Figure A.4), which describes the redox potential of the formulae.

Specifically, tannin-like features are oxidized; lipid-like features are reduced; and lignin-

like features have an average oxidation state around zero (Boye et al., 2017). This

21

suggests that the reduced lipid- and protein-like features more common to the manure

formulae can be expected to oxidize during the transport in aerobic surface waters. We

would therefore expect manure derived DOM to undergo the greatest amount of signature

change relative to other samples. A targeted analysis of similar samples (e.g., using LC

MS/MS of authentic standards to validate these proposed compounds) would be useful in

elucidating these structures of m/z values shared between our samples, which would

provide greater insights to their molecular properties (Lee and Kerns, 1999).

The similarity between edge of field sample and Sandusky River supports the

previously reported data that nutrient loads are predominantly sourced from agricultural

fields in this Lake Erie tributary. However, the edge of field, WWTP effluent, and

Sandusky River samples are hydrologically connected and would be expected to share

some background DOM signature received from rainwater, runoff, and/or groundwater in

the watershed. Still, the DOP signatures were highly divergent between manures and

other source materials, which should allow us to detect the presence of these nonpoint

and point sources in the tributary network. Formulae shared by the Sandusky River and

other samples were identified, and could serve as source markers in the watershed.

Additionally, we elucidated unique DOM and DOP signatures which could be used by

regulatory agencies to detect and monitor the presence of nutrient pollutant sources in the

tributaries.

22

Chapter 2: Dissolved Organic Matter Transport and Mixing in the Portage River

This chapter is currently being prepared for submission to the Journal of Great Lakes

Research with authors and title to be determined.

Introduction

Among the Great Lakes, Lake Erie has experienced the greatest degree of

eutrophication. Nitrogen and phosphorus loads from Lake Erie tributaries have

contributed to recurrent harmful algal blooms in the lake for much of the last half century

(D. Baker, 2010; Steffen et al., 2017). Since the early 2000s, increased nutrient loads

have led to recurrent toxic cyanobacterial blooms along the southern coastline of the

western Lake Erie basin, while hypoxia has developed in the hypolimnion of the lake's

central basin (Conroy et al., 2011; Michalak et al., 2013; Steffen et al., 2017). This

nutrient pollution is derived from both point (e.g., municipal/industrial wastewater

effluents or combined sewer overflows) and nonpoint sources (sewage leaks, urban area

runoff, or agricultural runoff/tile drainage) (D. B. Baker et al., 2014; Ohio Lake Erie

Phosphorus Task Force, 2013), with the loads from nonpoint sources being particularly

difficult to manage. The magnitude of the algal bloom for a given year is most strongly

correlated to spring (May-June) phosphorus loads from its tributaries (Stumpf et al.,

2012). Eutrophication became a crisis when, in 2014, microcystin remained in the

finished water of the Toledo water treatment plant, disrupting the service of residents and

incurring millions in economic losses (Steffen et al., 2017). These impacts have catalyzed

23

the need for better understanding nutrient pollution sources in the Lake Erie watershed

and developing source management strategies that mitigate pollutants to remedy the

problem.

Since 1975, the National Center for Water Quality Research has led monitoring

efforts for phosphorus and other nutrient pollutants in the region. Samples are collected at

an extensive network of stations to measure nutrient loads derived from the lake's

tributaries (D. B. Baker et al., 2014). Data collected from this network is used with

models that consider upstream land use in order to assess contributions from point and

nonpoint sources in the region (Michalak et al., 2013; Ohio Lake Erie Phosphorus Task

Force, 2010). However, these estimates are more reflective of riverine-scale contributions

(e.g., the Sandusky vs Maumee river) rather than those of specific sources (Michalak et

al., 2013). Efforts have been made to collect similar data from individual units, such as

the tile drainage of agricultural drainage (King et al., 2015), and combine these

hydrologic units to describe observations of the whole (i.e., tributary). However,

monitoring at the field scale would require intensive sampling because measurements can

vary significantly between different or even within the same field (e.g., hotspots). Other

manners of source identification and pollutant tracking are needed to estimate nutrient

loads to the basin and identify leading sources for targeted reductions.

Conservative tracers (e.g., concentration of inorganic ions, isotopes) have long

been used to estimate contributions from hydrologic sources (Barthold et al., 2011;

Doctor et al., 2006; Elsenbeer et al., 1995). For example, stable isotopes (13CDIC, 18O, 2H)

were used to estimate the mixing ratios of well water, river water, and anthropogenic

24

sourced waters at the border of Italy and Slovenia (Doctor et al., 2006). However,

isotopic fractionation does not always provide sufficient resolution for distinguishing

between point and nonpoint sources in watersheds. To this end, signatures of DOM

generated from fluorescent emission spectroscopy (Larsen et al., 2015; L. Yang et al.,

2015) or electrospray ionization Fourier-transform ion cyclotron resonance mass

spectrometry (ESI FT-ICR-MS) (Arnold et al., 2014; Kujawinski et al., 2009), have

proven useful in differentiating between distinct sources in some environments. For

example, indicator species of terrestrial and marine DOM were identified in surface water

of Atlantic Ocean samples, allowing for the discrimination between terrigenous and

autogenously produced organic carbon materials (Kujawinski et al., 2009). ESI FT-ICR-

MS has also been used to differentiate between forest or pasture-dominated headwaters

for a freshwater system (Lu et al., 2015). In addition to the two studies described above

which examine broad signatures of DOM, other, end member mixing analysis - using a

variety of inorganic tracers (isotopes, salinity, silica, potential temperature, etc.) - has

been used to model DOM components from mixtures of several sources (Hudson et al.,

2007; Medeiros et al., 2015; Wilson and Xenopoulos, 2009; L. Yang et al., 2015). For

example, end member mixing analysis was capable of modelling ESI FT-ICR mass

spectra of northern Atlantic Ocean samples from their four major sources of water

(Hansman et al., 2015). Conversely, several properties of DOM, elucidated through

electron emission spectroscopy, were found to be capable of acting as tracers in end

member mixing analysis (L. Yang et al., 2015). In other words, some conserved features

of DOM may be suitable as tracers in end member mixing analysis. Identifying conserved

25

components of DOM is crucial in any effort to use it for source tracking during transport

and mixing.

Changes during hydrologic transport complicates our ability to identify and track

DOM sources during as it moves downstream toward Lake Erie. DOM is susceptible to

changes from biological processing, photolysis, and abiotic reactions (e.g., hydrolysis or

oxidation) during transport. Its signature may also change from the autogenous

production of similar or unique compounds in the water column (Kellerman et al., 2015;

Medeiros et al., 2015; Stubbins et al., 2010). Certain, more calcitrant components of

DOM are more likely to persist along a river flow path. If these compounds are unique to

pollutant sources, they represent possible tracers in the hydrologic system. For example,

as much as 60% of DOM from marine samples could be attributed to DOM introduced

from the Amazon River. The terrestrial DOM remained present after mixing along the

continental shelf (Medeiros et al., 2015). In Swedish lakes, many N-containing DOM

formulae identified using ESI-FT-ICR-MS persisted in the water column, with minor

changes in peak abundance over time (Kellerman et al., 2015), suggesting limited

biological processing of N-containing features in the cold, submerged system.

Understanding how mass spectra change during transport from mixing, dilution, and

internal processing is critical to understanding the fate of DOM features.

Given the need to expand the tools available to assess nutrient pollutant source

loading to Lake Erie, the objective of this study was to evaluate how complex signatures

of dissolved organic matter change along a tributary due to branch mixing and other

hydrologic controls. Samples collected from upstream reaches to the mouth of the

26

Portage River were analyzed using ultrahigh resolution ESI FT-ICR-MS, which allows

for non-target analysis of dissolved organic matter. ESI FT-ICR MS is capable of

resolving thousands of DOM features to their elemental composition. The characteristics

of the assigned molecular formulae were compared between samples, with a focus on

shared and unique DOM features at confluence points. Linear mixing models were

applied to mass spectra data to test whether expected mixing ratios were conserved in

DOM signatures. The analysis showed that DOM is highly similar throughout the Portage

River, with some evidence that the similarities are due to the downstream transport of

DOM and linear mixing at merging branches. Organic matter originating from pollutant

sources could be monitored by watershed managers at downstream locations to detect the

presence of key pollutant sources in the watershed.

Methods

Sampling Locations and Collection

Samples were collected from 16 locations in the Portage River on April 6-7, 2017

using a Teflon-coated sample container attached to a rope (Figure 2.1A). Among the

Lake Erie tributaries, the drainage area of the Portage River accounts for 973 mi2 of the

nearly 14,000 mi2 total drainage area for Lake Erie (approximately 7%) (Ohio Lake Erie

Phosphorus Task Force, 2010). As it is among the smaller tributaries in the drainage

basin, it was selected for the ease of sample collection. Due the size of the watershed,

only one USGS gauge (#04195500) is used to collected real time water data. This gauge

is also the sampling site monitored by the Center for National Water Quality Research

(D. B. Baker et al., 2014). Therefore, significant gaps exist for the source of nutrient

27

loads and the contributions of flows from the upper reaches. The majority of the land use

in the watershed is dedicated to agriculture (76%), but with contributions from urban

(13%) and natural sources (11%) (Ohio Lake Erie Phosphorus Task Force, 2010). The

container was conditioned at each site by rinsing several times. A sample was collected

into HDPE containers for initial elemental (C/N/P) analyses. Acid-rinsed amber

glassware was baked at 550°C overnight to remove residual organic matter. This

glassware was used to collect the samples for mass spectrometric analysis. Duplicates

were collected at two locations, to confirm reproducibility. All samples were stored on

ice during collection and subsequently stored at 4°C until processing within 14 days.

Following sampling, Milli-Q water, and Suwannee River Fulvic Acid (SRFA) dissolved

organic matter standards were prepared and processed with the samples as

methodological controls.

In addition to the collection of water samples, hydrologic data, including drainage

area and predicted recurrent intervals for storm water flows was gathered for ungauged

sites using USGS StreamStats version 4 (Ries III et al., 2017). Return interval flows were

used to estimate volumetric flows originating from upstream tributaries at the four

confluence sampling points. The contributions from these tributaries were calculated

using the 2-year return interval (Figure 2.1B).

28

Figure 2.1. Sampling locations in the Portage River. (A) Sixteen locations in the Portage

River watershed were sampled from the east branch near to the mouth of the river. Each

site was identified by the road on which it was sampled, but assigned an alphabetical

letter in order from upstream (A) to downstream (H) locations. There were four locations

that included a confluence point (E-H) at the intersection of two tributaries with primary

tributaries labeled with 1 (e.g., E.1), secondary tributaries labeled with 2, and the

confluence being labeled with 3. (B) Drainage information for confluence sample points.

Contributions from the two tributaries were calculated from the 2-year return interval

flows.

Sample Processing

The samples for dissolved nutrient analysis were pre-processed by vacuum-

filtration using 0.7-µm glass fiber filters (Whatman GF/F). Phosphorus was measured in

three different forms: dissolved reactive phosphorus (DRP), dissolved hydrolysable

phosphorus (DHP), and total dissolved phosphorus (TDP) using standard colorimetric

methods (EPA 365.3) with automated analysis (Seal Analytical Autoanalyzer III). The

concentrations of dissolved organic carbon (NPOC) and total dissolved nitrogen (TDN)

were determined using a Shimadzu TOC-V/TNM-1 analyzer. NPOC was determined

according to EPA method 415.1 while TDN was determined (ASTM D8083) (Kekacs et

29

al., 2015). Calibration curves were generated between 3-20 mg C/N L-1 using potassium

hydrogen phthalate or potassium nitrate, respectively, with limits of detection at 2 mg C

L-1 and 0.01 mg N L-1. In order to estimate solid phase extraction efficiency, TDP was

measured on an Agilent ICP-OES at 213.648 nm according to EPA method 3051 (Bartos

et al., 2014).

Combusted glassware (30 min at 550°C) was used for mass spectrometric

analysis. All samples were vacuum-filtered through pre-rinsed (methanol and DI water)

0.7-µm glass fiber filters (Whatman GF/F). Solid phase extraction was performed with

Plexa-PAX columns using 325 mL of undiluted samples adjusted to pH 10 with sodium

hydroxide. Initial NPOC concentrations ranged between 6-14 mg C L-1 (Table B.1).

Readings of NPOC, TDN, and TDP were made on solid phase extraction influent and

effluent samples to estimate the amount retained by PAX columns (Table B.1). PAX

columns were primed using Milli-Q water and methanol per the manufacturer’s

instructions. DOM was eluted from SPE columns using 10 mL of HPLC-grade methanol,

followed by 10 mL of HPLC-grade methanol +5% formic acid. The elutions were

combined and stored at -20°C for ESI FT-ICR-MS analysis at Woods Hole

Oceanographic Institute.

ESI FT-ICR-MS Data Analysis

Mass spectrometry data was collected as has been previously described

(Appendix A) (Minor et al., 2012). The samples were analyzed with electrospray

ionization under the negative ionization mode on a 7T FTICR mass spectrometer

(Thermo Fisher Scientific, Waltham, MA USA). The instrument settings were optimized

30

by tuning on the SRFA standard. The samples were infused into the ESI interface at 4 μL

min-1, and the instrumental and spray parameters were optimized for each sample. The

capillary temperature was set at 250°C, and the spray voltage was between 3.7 and 4 kV.

For each sample, 200 scans were collected spanning the 150-1000 Da m/z range.

Molecular formula assignments were made using the Compound Identification Algorithm

with an error of 1 ppm (Kujawinski and Behn, 2006; Kujawinski et al., 2009). A total of

29,273 unique peaks were detected. Two quality control measures were used to filter the

dataset (1) peaks observed in DI water or solvent blanks were removed from all samples,

and (2) any singletons were removed from the dataset (Table B.3). Additionally, only m/z

values with an assigned formula were considered in further analysis.

Data analyses were performed using R Statistics (version 3.1.1). Samples were

compared based on (1) presence or absence and (2) relative peak height. For relative peak

height comparisons, peak abundances were normalized to maximum peak height for each

sample. Cluster analysis was performed using the ‘vegan’ package distance methods,

with the Jaccard method used for analysis of presence/absence data and Canberra method

used with relative peak heights (Oksanen et al., 2015).

We considered the potential of downstream transport as the detection of the same

m/z value between upstream and downstream samples. The more prominent peaks, or

those with the greatest height, were expected to be detected in downstream samples. The

m/z values were binned by peak height working in 0.01th quantile intervals. The

probability of positive detection in the nearest downstream neighbor was calculated for

each of those quantile bins. End member mixing analysis models were developed using

31

the ‘quadprog’ package to reveal whether the DOM spectra were linearly mixed at the

expected ratios according to the StreamStats estimates. The model considered the m/z all

m/z values detected in the confluence and at least one of the branch samples. The

solve.QP command was used to develop models for the confluence according to:

𝐶𝑜𝑛𝑓𝑙𝑢𝑒𝑛𝑐𝑒 = 𝑎 × 𝑇𝑟𝑖𝑏𝑢𝑡𝑎𝑟𝑦1 + 𝑏 × 𝑇𝑟𝑖𝑏𝑢𝑡𝑎𝑟𝑦 2

where a>0, b>0, and a+b=1. Peak heights we m/z values were used to solve the equation

across the factorized (Cholesky decomposition method) set to account for covariance

between variables. A random sample of 500 m/z values were used to solve the mixing

equation. Bootstrapping (n=1,000) was used to generate a 95% confidence interval for the

estimated mean contribution of each branch to its confluence with the assumption that

predictions followed a normal distribution.

Results

Carbon, nitrogen, and phosphorus concentrations varied across the Portage River

Table B.1, Figure 2.2). NPOC ranged between 6 and 47 mg C L-1 with the highest value

observed near the Fostoria WWTP collection site (Location D). The NPOC measured at

the WWTP site were more than 3-fold higher than all other samples, which were <12.6

mg C L-1. Carbon concentrations were higher in the upper reaches of the east branch

where all five samples were measured at >10.3 mg C L-1. The south branch and all

samples downstream of the first confluence sampling area (Location E) were measured at

<10 mg C L-1. The TDN concentrations of the samples ranged between 5.9 and 10.2 mg

N L-1. The concentrations in the east branch were between 7.2 and 7.4 mg N L-1 until the

first confluence point (Location E.2) which showed the highest measured N levels (10.2

32

mg N L-1). Concentrations dropped to 8.0 mg N L-1 or less after this sample location.

TDP ranged between 124 and 132 μg P L-1 in the upper reaches (A-E), dropping to

between 101 and 113 μg P L-1 leading up to the final mixing area (F-G). Toward the

mouth of the river (H), the concentration fell to 83 μg P L-1 or less as the river reached

Lake Erie. On the order of 87-94% total phosphorus was measured as DHP, while 71-

86% was measured as DRP. Generally, DRP and DHP accounted for a lower proportion

of TDP as samples moved downstream (A through H).

33

Figure 2.2. DOC, TDN, and TDP concentrations measured in Portage River samples.

Points are labelled to their respective samples, with the color indicating the dominant

(darker) and secondary (lighter) branches at each confluence. Sample D, near the Fostoria

WWTP, was detected at nearly 47 mg C L-1.

A total of 23 samples were analyzed with ESI(-) FT-ICR-MS analysis including

blanks, standards, and replicates (Figure B.1). Over 29,000 unique m/z values were

34

observed across the set. Following quality filtering, a total of 11,344 m/z values remained

in the dataset between the mass range of 150-1000 Da (Table B.3). While these

preprocessing steps removed 72% of detected values, at least 74% of the detected m/z

values were retained for each sample. Almost all these m/z values were assigned a

formula (11,064, 99%), and these data were then normalized and used for all analyses

described in the following sections.

The data were summarized based on elemental composition of assigned formulae

(Figure 2.4A, Table B.4). The CHO compounds dominated all samples (57% to 73%),

but there was also an abundance (22.6-36.2%) of CHON compounds throughout the river

(Table B.3). The lowest abundance of CHON was observed in the upstream reaches of

the east branch (A-C), near Fostoria, Ohio, while the highest abundances were detected in

the samples near the mouth of the Portage River (H.3). However, this patttern was not

reflected across the range of TDN concentrations and CHON% values (p=0.28, Figure

B.2A). A noticeable spike in the percentage of CHON formulae wasdetected near the

Fostoria WWTP (D), yet increased organic nitrogen diversity was not manifested as an

increase in TDN concentrations. Throughout the hydrologic system, the percentage of

CHON formulae fluctuated substantially, and usually this corresponded to the change in

CHO formulae. Despite an overall reduction in TDP from upstream to downstream

locations, the relative abundance of phosphorus containing formula also increased from

upstream reaches toward the mouth of the Portage River (Figure B.2B). The sample

closest to the mouth of the river (H.3) had the lowest abundance of CHO formulae

35

compared to any other sample, with organic matter becoming more nutrient-laden (i.e.,

higher in organic-N/P/S abundance).

Van Krevelen diagrams were used to visualize the elemental composition of m/z

values collected from the 16 locations (Figure 2.4A). River samples were clearly

dominated by a cluster of CHO and CHON formulae in the lignin-like and tannin-like

region of the plot. The CHOP formulae were primarily evident in the lignin-like and

protein-like regions. There were a few noticeable differences in the prevalence of CNON

heteroatoms in the carbohydrate-like region (e.g., E.1 or H.3 vs A). In addition, CHOS

formula showed notable differences in the lignin/condensed hydrocarbon-like regions

(e.g., H.3 vs A or F.3 vs B).

36

Figure 2.3. Van Krevelen plots of the 16 samples in this study. The hydrogen:carbon molar ratio is plotted against the oxygen:carbon

molar ratio for each formula identified in the sample. The seven boxes indicate the molecular class features of each individual

formula.

37

Next, we summarized the data according to the molecular classifications of m/z

values indicated by their molar ratios to compare the sample similarities based on

molecular features. All samples were dominated by lignin-like features (≥73%, Table

B.5, Figure 2.4B). In general, tannin-, protein-, and lipid-like features were the next most

abundant features for many of the samples. Again, a major shift was observed between

samples C and D, near the Fostoria WWTP as tannin-like features increased and lignin-

like features decreased in relative abundance. There were similarities between the

samples A through C with protein-like features becoming more abundant (7.1-8.3%)

largely at the expense of lignin-like features (80.3-79.3%). In sample D, the shift became

more dramatic with lignin-like features decreasing by nearly 4% (75.6%), but in this case

tannin-like features became more predominant. However, the nearest downstream sample

(E.2) did not share these characteristics, and sample D in fact shared more similarities

with the other branch (E.1) at the confluence E. Thus, the effects from the WWTP

effluent may have diminished during downstream transport. Carbohydrate-like features

ranged from 0.5-2.3%, and these were highest in the H.2 and H.3 samples that were

collected near the mouth of the river. Again, the samples nearest to the mouth had

striking differences with the upstream samples in that lignin-like features were much less

abundant and tannin-like and unclassified (‘other’) features became more prominent.

38

Figure 2.4. Summary of elemental composition and molecular classes of assigned

formula. (A) Pie charts show the allocation of elemental classes by each sample. (B) Bar

charts illustrate the relative abundance of the molecular classifications based off the

windows assigned in the Van Krevelen plots. Lignin-like features accounted for >50% of

these classes for all samples.

To first consider mixing at the confluence points, we used Venn diagrams to show

the proportion of shared m/z values between samples (Figure 2.5). The majority of peaks

detected were shared across all three samples except in the upstream most confluence

location E). Location E1 at this confluence location had the greatest uniqueness across all

three samples with nearly 28% of formulae detected in only one of the samples on an

individual basis, sample G.1 was the most unique of any from the other two samples in its

sampling area, with 18.9% unique m/z values. The larger tributary (E.1, F.1, G.1, and

H.1) shared between 61.5-71.9% of their formulae at the four confluence locations,

whereas these proportions ranged from 56-76% for the smaller tributary (E.2, F.2, G.2,

and H.2). Both site E and G had a greater degree of similarity between the smaller

tributary than the larger tributary with the confluence sample. This was surprising as

higher DOM loading was expected from the larger of two tributaries.

39

Figure 2.5. Percentage of shared and unique formulae between samples at the confluence

sampling locations. The data is plotted in Venn representations but scaled to the relative

number of m/z values detected in each sample. The *.1 sample is expected to provide the

greater proportion of flow, relative to *.2, at the confluence point, *.3.

The similarity between samples were compared using the binary (Jaccard)

distance matrix combined with hierarchal clustering (Figure 2.6). Notably, the m/z values

with the greatest peak heights clustered together and were detected across all samples in

40

this study. Between confluence samples, F.1 clustered closely to F.3, and G.2 clustered

with G.3. Alternatively, there was greater separation between the samples from

confluence E, and even greater separation between confluence H samples. One of the far

downstream sites, H.1, clustered with sample D, near the Fostoria WWTP. Outside

sample D, the upstream samples collected around the outskirts of Fostoria were all

clustered closely together, suggesting that there was a substantial change to the DOM

signature near the Fostoria WWTP. Hierarchal clustering was also performed using the

Canberra distance matrix (using noramlized peak heights). This analysis resulted in

improved clustering between the confluences and their two tributaries (Figure B.3),

suggesting low abundant peaks represent noise in the dataset. Additionally, cluster

analysis using Canberra distances resulted in more similarity in transects closer in

distance (e.g., D and E.1), rather than similarity to samples collected further downstream

(D and H.1). Otherwise, there were few differences in the clustering of samples using

presence/absence as compared with normalized peak heights.

41

Figure 2.6. Clustering analysis of m/z values. Dendrograms were prepared for the

samples and for the m/z values (not shown) using the binary Jaccard distance matrices.

Heatmap values are scaled according to the normalized peak heights for each m/z value.

To characterize the DOM being transported, the five samples (A through E.2)

collected in the upstream reach of the east branch were compared based upon the shared

m/z values between the nearest neighbors. The normalized peak heights were binned

along the quantiles for each sample at 0.01th increments. Specific m/z values were then

sought in the downstream sample location (Figure 2.7A). Notably, the probability of

detecting a m/z value at the nearest downstream sample increased with greater peak

height magnitudes. At the 0.25th quantile, the probability of downstream detection was

42

typically greater than 75%. This relationship held for sample pairs close in proximity

(within 2 miles). In the case of the longest flow path distance (D through E.2, ~10 miles),

the m/z value had to be above the 0.5th quantile to have a 75% chance of being detected

downstream. Next, we compared the distribution of normalized peak heights by different

elemental composition (Figure 2.7B). The peak heights were considered for all 16

Portage River samples, and formulae detected in multiple samples were counted that

many times. It is evident from this analysis that the CHO and CHOP formula were higher

in abundance than other formula classes. Many of the CHONP and CHOPS formula had

smaller peak heights (<0.25th quantile). Still, CHO and CHON formula accounted for

>93% of the detected formula in each sample, so it is unlikely that other elemental

compositions will be cross-detected following transport from upstream to downstream

locations.

43

Figure 2.7. Quantiles of the peak heights for observed m/z values. (A) The probability of

detecting the same m/z value at the downstream sampling location in the east branch

reach was determined for each 0.01 quantile increment. The quantile value for the

upstream sampling location (e.g., A) is plotted on the x-axis against the detection

probability at the downstream sampling location (e.g., B). (B) The distribution of these

quantiles by heteroatom groups is shown in violin plots.

To demonstrate linear mixing between tributaries at the confluences, DOM was

used in the mixing analysis to estimate the contributions of each intersecting branch. For

a reference of comparison, StreamStats was used to estimate the predicted mixing

between these two branches (Table B.1). Quadratic programming was used to solve the

confluence mixing equation using the peak heights of each m/z value providing a

contribution from each branch. These DOM mixing models were performed by

bootstrapping to provide a range of predictions, and calculate the 95% confidence

interval around the mean of these models (Figure 2.8). The prediction ranges and means

of these DOM models were then compared to the StreamStats derived estimates. The

estimates between these two approaches were similar at confluence E, H, and F.

44

Although there was no overlap between the 95% confidence interval and the StreamStats

estimate, the prediction range did include the StreamStats estimates at these sites. Most

notably, the estimated contributions at confluence E were ±3% of each other. Confluence

G samples had the greatest disagreement between StreamStats and our mixing model

estimates, but this confluence also had the greatest variation from the DOM mixing

models. Notably, of the samples collected in this study, G.1 and G.2 were the furthest

distance away from their respective confluence point. Based on the similarity between the

DOM mixing model and StreamStats estimates, these results suggest that DOM mass

spectra signals mix linearly within the watershed.

45

Figure 2.8. Comparisons between the StreamStats and DOM mixing model contribution

estimates. The range of predicted values from bootstrapped DOM mixing models are

displayed as violin plots. The 95% confidence interval of predicted means are shown as

the yellow band. The StreamStats estimates for each sample is shown as the purple

diamond.

Discussion

Concentrations of organic carbon, total nitrogen, and total phosphorus were

highly variable throughout the Portage River watershed. Notably, as the samples

approached the mouth of the river, the values of total dissolved phosphorus decreased.

This could be due to high phytoplankton activity and assimilation of this nutrient

resulting in the drop of phosphorus concentrations as has been observed along marine

46

coastlines (Ruttenberg and Dyhrman, 2005). It could also be related to the loss of

phosphorus due to particulate adsorption. Corresponding with this decrease in TDP, the

relative proportion of P-formulae increased in those samples, possibly indicating an

enrichment of DOP at these locations. However, with the use of PAX-SPE we expected

to concentrate DOP, so these two variables are not directly comparable. While the

amount of DOP (TDP-DHP) was not higher at these locations, the proportion of DOP of

the TDP was measured at the highest ratio in the sample closest to the mouth and higher

at these marginal locations. Phytoplankton has a general preference for the inorganic

forms of phosphorus which does seem to enhance the proportion of DOP relative to TDP

in coastal sites (Monaghan, E. J., Ruttenberg,K.C., 1999; Ruttenberg KC, 2012).

Throughout the Portage River, the percentage of P- and N-formulae fluctuated between

samples. We expected this to relate to the concentrations of both total nitrogen and

phosphorus, but only found a negative correlation between total dissolved phosphorus

and percentage of CHOP* formulae.

The Portage River drains to the western Lake Erie Basin, and has 76% of its

watershed area is dedicated to agricultural land use (Ohio Lake Erie Phosphorus Task

Force, 2010). In that sense, the Portage River is similar to the Sandusky River with

slightly more area dedicated to urban land use (13% vs 10%) (Ohio Lake Erie

Phosphorus Task Force, 2010). The samples collected between these two watersheds

were compared to determine the similarity of source signatures to the samples collected

throughout the Portage River. Comparisons were made between the two sets of DOM

data using shared molecular formulae. Many of the formulae (4567, 63%) identified in

47

the Sandusky dataset were also detected in the Portage River dataset. Of those shared

formula, only 26 included a phosphorus atom, meaning <4% of the DOP formula

identified in the Sandusky dataset were also detected in the Portage River. Additionally,

four of our proposed markers were detected within the Portage River: C10H23O3N4P,

C17H25O9P, C18H23O9P, and C17H21O9P. However, the Portage River samples were

collected during a high flow event. Possibly, the proposed markers were diluted and

obscured from detection due to diffuse, background DOM. Such background DOM could

also explain the high degree of similarity between the 16 Portage River samples.

The Portage River samples were dominated by lignin-like and tannin-like

features, similar to the Sandusky River sample analyzed in the previous set. However,

there were a greater number of carbohydrate-, protein-, and lipid-like features with the

Portage River samples compared to the Sandusky River. A greater number of CHON

formula were detected in the Portage River samples compared to the Sandusky River

(>23.8% vs 20.8%). Although CHON were in greater abundance in Portage River

samples, the number of P-containing formula were lower in all but three of the samples:

Bridge St (3.7%), Chet’s Place (3.5%), and Portage River Retreat (3.6%); while Bierly

Ave (3.1%) was within the range of P formula detected in the Sandusky River (3.3%).

Our dataset was also relative enriched in CHON content, but lower in P containing

formula then samples collected in the tributaries of Lake Superior (Minor et al., 2012).

Compared to an urban (3% agriculture) and mixed-land use (28% agriculture) stream in

Florida, our samples were also more enriched in CHON formula (Lusk and Toor, 2016).

In fact, the relative abundance of CHON formulae was twice as high as that of the mixed-

48

land use stream (11.6%) (Lusk and Toor, 2016).. Another potential source of CHON

formula, however, is sewage (e.g., septic contaminated groundwater) which also has high

CHON content (Arnold et al., 2014). Treated wastewater effluent may also be enriched in

sulfur formulae (Gonsior et al., 2011). Our sample (D) collected nearest to the Fostoria

WWTP had a notable increase in N- (25 to 35%) and S-formulae (2.1 to 2.4%) from the

sample upstream, location C. Both S- and N-formulae (3.6 and 37%) became more

enriched in the DOM spectra of the samples closest to the mouth reaching proportions

that exceeded those next to the WWTP. The character of the Portage River samples

appeared similar to other natural riverine systems, in terms of lignin-like features, but had

a relatively high abundance of CHON compound.

The samples from the Portage River clustered together often in relation to their

physical proximity to one another. Notably, the dominant peaks detected in this study

were present in most if not all samples, and normalized peak heights were nearly the

same as well. For example, the upper reach samples of locations A through C cluster

together in Jaccard and Canberra distance matrices. However, the next nearest site to

these, location D sampled by the WWTP, clustered near other samples further

downstream. We hypothesize that this may reflect DOM influenced by sewage from

WWTP or septic inputs (Maizel and Remucal, 2017). Hierarchal clustering analysis

between the Portage River and Sandusky watershed samples was also performed on

Jaccard distances using only the formula that were cross-detected (Figure B.3). While we

expected the Fostoria WWTP sample (D) to cluster closely to the WWTP effluent sample

collected in the Sandusky sampling set, this did not turn out to be the case. This

49

difference though may be on account of differences between the final treatment processes

used at the Tiffin (chlorination) and Fostoria (UV irradiation) wastewater treatment

plants. Wastewater treatment plants expose influent organic matter to physical (e.g.

filtration), biological (e.g. activated sludge), and chemical (e.g. chlorination) processes

that affect the organic matter as it proceeds through these treatment processes (Maizel

and Remucal, 2017). The biological processes at one such facility was associated with an

increased production of CHON, CHOS, and CHOP formulae (Maizel and Remucal,

2017). The observed increase of these features near the Fostoria WWTP may highlight

their ability to detect sewage contamination. Further, their abundance at the mouth of the

Portage River may highlight the amount of contamination originating from sewage

sources in this watershed.

Changes to the DOM through production or degradation does not stop with the

discharge from point and nonpoint sources to riverine systems. Rather, internal riverine

processing continues within the waterways due to indigenous biota, photolysis, and

exposure to abiotic elements (Mesfioui et al., 2012; Stubbins et al., 2010). Therefore, we

sought to identify the compounds which persisted during their transport through the east

branch of the Portage River. The samples collected from the upstream reaches of the

east branch of the Portage River were used to elucidate the downstream transport of m/z

values. We found that the probability of detecting a m/z value corresponded to the

magnitude of its peaks. In other words, the most prominent peaks may be the most

reliable markers for tracking sources of pollution. As DOM is transported, it may have

been diluted due to unaccounted headwaters merging with our stem of the Portage River

50

between sampling locations (Medeiros et al., 2015). These inputs may also have

contributed to relatively stable levels of certain m/z values, especially if these

components are originating from diffuse sources.

Another consideration in the transport of the m/z values is the molecular class

features which may affect biodegradability. For example, bioassays looked at the

susceptibility of N containing formula and found that protein- and lipid-like features were

more reactive than lignin-like features (Lusk and Toor, 2016; Mesfioui et al., 2012). Still,

there may not be dramatic changes as only 5-7% of the DON was removed during a five-

day bioassay (Lusk and Toor, 2016). Thus, at the upper reach sampling locations, we also

considered the change in peak heights of m/z values as a function of their molecular

classification (Figure B.4). We found that many of lignin-like features were less likely to

change between our sampling points (i.e. median change for these features was 0 between

upstream and downstream neighbors). Like the CHON bioassay studies, the peak heights

of the protein- and lipid-like features were more likely to decrease in the corresponding

downstream sample. While we did not further consider this factor, it may be worthwhile

to select for only those features which show limited degrees of change in peak heights

over the course of downstream transport, as these may serve as more robust markers to

track sources of pollution.

The constrained, linear mixing models developed from DOM spectra were in

good agreement with the estimates we calculated using StreamStats predicted flows. Only

the third confluence point (G) yielded poor results, whereas the other confluence models

performed well with differences of <9%. An important note is that sample G.1 was

51

sampled just upstream of the local WWTP and was also the furthest distance from its

confluence sample (G.3) of any sample collected. If, as we hypothesize, WWTP effluent

has a significant impact on DOM, the WWTP situated between G.1 and G.3 sampling

locations may partially explain the failure of this model. To our knowledge, ESI FT-ICR-

MS has not been used within end member mixing analysis to provide source

contributions. However, mixing analysis has been used to reproduce the ESI FT-ICR-MS

spectra from the Atlantic Ocean from several riverine sources (Hansman et al., 2015).

Additionally, this technique was used to track the loading of DOM from the Amazon

River into the Atlantic Ocean finding that many of m/z values persisted and were diluted

from mixing with autogenous DOM (Medeiros et al., 2015). These studies suggest that

the mass spectra elucidated with ESI FT-ICR-MS, specifically the more conserved or

recalcitrant features may follow a linear mixing model in certain hydrologic systems. As

ionization efficiencies of different compounds can vary considerably and the inability to

acquire standards limits the quantification of individual compounds, non-target ESI-FT-

ICR-MS methods are semi-quantitative at best. However, recent studies support the

notion that relative peak heights provide an indication of relative abundance for m/z

values (Banerjee and Mazumdar, 2012; Kamga et al., 2014; Lu et al., 2015). The

feasibility of using DOM for source tracking or within end member mixing analysis relies

on better defining the quantification limits of ESI FT-ICR-MS. Our results suggest that

there is some potential for DOM to be used in end member mixing analysis pending

further evaluation.

52

Chapter 3: The Emerging Concern of Antibiotic Resistance Genes in Agricultural

Sediments

This chapter will be submitted under the title: The Prevalence of Antibiotic Resistance

Genes in Agricultural Channel Sediments by Michael R. Brooker, Julia Beni, Timothy

LaPara, Jill Kerrigan, Bill Arnold, Paula J. Mouser. This manuscript will be submitted to

Water Research. This work was supported by the National Integrated Water Quality

Program Award Number 2012-51130-20255 from the USDA National Institute of Food

and Agriculture.

Introduction

In the Midwestern United States, agricultural drainage is managed by constructing

trapezoidal ditches optimized to remove excess water from the field. Recently, there have

been efforts to restore these ditches to resemble more natural streams by incorporating

floodplains by widening the drainage corridor (Powell et al., 2007). This practice of two-

stage channel construction captures sediments and establishes macrophyte communities

which support a productive microbial habitat. Such microbial communities provide

ecosystem services, for instance, these floodplains are strongly associated with

denitrification thus attenuating downstream eutrophication (Rabalais et al., 2002; Roley

et al., 2012). Due to the novelty of this two-stage practice, there has been minimal

characterization of the microbial ecology where much of the emphasis has been on

ecosystem services provided (e.g. denitrification) (Arango et al., 2007; Roley et al.,

2012). Additionally, the microbial ecosystems of floodplains have been recognized for

their role in treating other agricultural pollutants like herbicides and pesticides (Douglass

et al., 2015; Vymazal and Březinová, 2015). Microbial characterization of two-stage

channels must be performed to fully understand the potential benefits and drawbacks of

this practice.

53

One method for assessing the functional potential of uncultivated microorganisms

in these systems is through metagenomic techniques. The GeoChip microarray is a

metagenomic tool with the capability to semi-quantitatively detect thousands of

environmentally relevant functional genes equally across samples (Zhou et al., 2010).

These genes span a range of categories, but notably include those involved in nutrient

cycling (i.e. organic remediation) or aiding in microbial stress (i.e. antibiotic resistance or

metal homeostasis). This technique has been used to characterize the microbial

ecosystems of many environments; wastewater treatment, grasslands, forests, and riverine

sediments (Cong et al., 2015; Low et al., 2016; Sun et al., 2016; Y. Yang et al., 2014).

Among the many systems characterized by the GeoChip, a relatively high

abundance of antibiotic resistance genes (ARGs) were noted in urban watershed

sediments (Low et al., 2016). The ARGs in these urban sediments were predominantly

efflux pumps, which may confer multidrug resistance (Low et al., 2016). However, these

efflux pumps can provide resistance to other toxins, may be co-selected for with metal

resistance, or used in intercellular signaling (Baker-Austin et al., 2006; Low et al., 2016).

ARGs are maintained or proliferated through in situ selection by antibiotics or other

toxins, microbial migration, or horizontal gene transfer (HGT) (Niehus et al., 2015). In

HGT, mobile genetic elements (i.e., integrons, transposons) may be transmitted through

plasmids, and integrons were implicated in the enrichment of ARGs in agricultural soils

amended with hog manure (Johnson et al., 2016). ARGs have been detected in

groundwater, tile drainage, and downstream watersheds connected to agricultural fields

54

(Frey et al., 2015; Storteboom et al., 2010). Their presence would be expected in the

sediments collecting within drainage corridors.

Antibiotic resistance is a naturally occurring phenomenon that occurs

ubiquitously across all environments. In recent times, the use of antibiotics for medical

and agricultural purposes has increased the prevalence of this functionality in

anthropogenically-impacted environments (Hobman and Crossman, 2015; Martinez,

2009). In the United States, a majority of antibiotic usage is dedicated to agricultural

operations (Center for Disease Dynamics, Economics & Policy, 2015). Thus, there are

emerging concerns over the prevalence and persistence of antibiotic resistance genes

(ARGs) in agroecosystems (Rothrock et al., 2016). Antibiotic resistance in agricultural

fields occurs through several routes of exposure, for instance manure spreading (Ghosh

and LaPara, 2007; McManus et al., 2002; Munir et al., 2011; Negreanu et al., 2012;

Schmitt et al., 2006; Udikovic-Kolic et al., 2014). Antibiotic resistant organisms are not

confined to these fields, but are rather transported through ground and surface water

flows (Chee-Sanford et al., 2001; Chee-Sanford et al., 2009; Frey et al., 2015). While the

presence of ARGs at agricultural sites and their downstream transport in surface water

flows have been studied, little is known about these genes in agricultural channel

sediments.

Agricultural channels, with natural floodplains developing within their banks

(Powell et al., 2007), were selected for our analysis. Three sites located in the Western

Lake Erie Basin (WLEB) were selected to characterize microbial ecosystem services and

taxonomy. Our previous research had determined that the chemical and physical

55

properties of these sites were strongly associated with their designated Ecoregion

(Brooker, 2018). Due to the variations in soil properties, we hypothesized that the

microbial communities would be diverse from one another across these sites. We used the

GeoChip 5.0 to assess the functional genomic diversity, while 16S rRNA gene

sequencing was used for taxonomic analysis. Specific classes of functional genes were

also assessed using Fluidigm quantitative PCR (qPCR). Our results identified ARGs,

metal homeostasis genes, organic remediation genes, and an integrase gene to be

prevalent in the sediments of Lake Erie drainage corridors that indicates these sediments

are a considerable reservoir for antibiotic resistance in the environment.

Methods

Site Description and Sample Collection

Sampling locations were chosen in three U.S. EPA defined Ecoregions of the

Western Lake Erie Basin: Clayey, High-Lime Till Plains (CHLP), Oak Openings (OO),

and Paulding Plains (PP) (Omernik, 1986). Criteria included in the site selection included

1) the presence of self-formed floodplains, 2) adjacency to agricultural row crop fields,

and 3) greater than 70% agriculture land use in the watershed. Samples were collected on

October 1, 2015 and again on October 18, 2016 to assess ecosystem services using

GeoChip and 16S rRNA gene sequencing. The samples collected in 2016 were further

analyzed using Fluidigm qPCR as well as assessed for a suite of common antibiotics and

metals.

Sediment cores were extracted from the surface (0-20cm) of floodplains using a

soggy bottom sampler device with sterilized PVC liners (AMS). The dry weights for

56

sediments were calculated after drying more than 10g of sediment at 70°C until no further

changes in weight were observed (approximately 24 hours). For quantification of

antibiotic concentrations, samples were collected in combusted glassware (450°C, 2

hours) using ceramic spoons. Samples were transported on ice to the laboratory where

they were stored at -20°C overnight. Samples used for antibiotic quantification were

freeze-dried over the course of a week, while DNA extracts commenced on the day

following sampling.

Genomic DNA Extraction

Community DNA was extracted using the MoBio PowerSoil DNA extraction kit

as previously described (Brooker et al., 2014). For the 2015 samples, duplicate cores

were collected from each sampling location. Each core was homogenized separately, and

two 0.5-mg (wet weight) aliquots were taken from each core for extraction (n=12). After

extraction, DNA was pooled for each core (n=2), resulting in six total samples for

analysis (duplicates from three different sampling locations, n=6). In 2016, DNA was

extracted using 0.25-mg (wet weight) of homogenized sediment from triplicate cores at

each sampling location, resulting in nine total samples for analysis. Triplicate cores were

pooled by site for GeoChip and 16S rRNA gene sequencing (n=3). All replicates were

analyzed using Fluidigm qPCR (n=9). DNA was stored at -80°C until shipment on dry

ice to sequencing and analysis facilities.

Functional Gene Assays

The GeoChip 5.0 analysis was performed at Glomics, Inc. (Norman, OK) and has

been described in detail (Cong et al., 2015) The GeoChip 5.0 microarray consists of

57

167,044 probes covering about 1500 gene families of functional genes commonly

observed in the environment. Briefly, purified DNA was labelled with Cy3 fluorescent

dye with a random priming method and hybridized to the GeoChip 5.0 array slide. Slides

were washed and scanned at 633nm using a laser with a NimbleGen MS200 Microarray

Scanner. Data was preprocessed by the microarray analysis by removal of poor quality

spots (SNR <2.0).

A combination of antibiotic resistance, metal homeostasis, and integrase genes

were selected for qPCR analysis using an integrated fluidic circuit (Fluidigm

Corporation, San Francisco, CA) (Johnson et al., 2016). Briefly, 48 primer sets and nine

samples (2016) were input into a 48.48 Access Array. EvaGreen dye was used at the

fluorescent marker, which allowed for real-time quantification of amplification products.

The amplicon pool was prepared by the Fluidigm FL1 and FL2 workflow. Several genes

were targeted with multiple primer sets, and these were annotated with a gene suffix (e.g.,

aadA and aadA5), and primer sets used can be found in Johnson et al. (2016). Threshold

cycle values were quality checked by Fluidigm software. Standard curves for each gene

were used to estimate the number of copies detected in each sample. The number of gene

copies per gram of sediment (dry weight) was calculated using the amount of eluent

applied in DNA extraction and the amount of sediment added to the extraction tube

(Table C.1).

16S rRNA Gene Sequencing

PCR amplicon libraries targeting the 16S rRNA gene were produced using a

barcoded primer set adapted for the Illumina platform (Caporaso et al., 2012). Each 25

58

µL PCR reaction contained 9.5 µL of MO BIO PCR Water (Certified DNA-Free), 12.5

µL of QuantaBio’s AccuStart II PCR ToughMix (2x concentration, 1x final), 1 µL

Forward Primer (5 µM concentration, 200 pM final), 1 µL Golay barcode tagged Reverse

Primer (5 µM concentration, 200 pM final), and 1 µL of template DNA. The conditions

for PCR were as follows: 94 °C for 3 minutes to denature the DNA, with 35 cycles at 94

°C for 45 s, 50 °C for 60 s, and 72 °C for 90 s; with a final extension of 10 min at 72°C to

ensure complete amplification. Amplicons were then quantified using PicoGreen

(Invitrogen) and a plate reader (Infinite 200 PRO, Tecan). Once quantified, volumes of

each of the products were pooled in equimolar amounts, cleaned up using AMPure XP

Beads (Beckman Coulter), and quantified using a fluorometer (Qubit, Invitrogen). After

quantification, the molarity of the pool is determined and diluted to 2 nM, denatured, and

then diluted to a final concentration of 6.75 pM with a 10% PhiX spike for sequencing on

the Illumina MiSeq. Amplicons were sequenced on a 151bp x 12bp x 151bp MiSeq run

using customized sequencing primers and procedures (Caporaso et al., 2012). The

sequencing was performed at the Joint Genome Institute (San Francisco) or Argonne

National Laboratories (Chicago, IL) for the 2015 and 2016 sets, respectively.

Antibiotic Extraction and Analysis

Antibiotic extraction and quantification was carried out according to Kerrigan et

al. (2017). All glassware was triple-rinsed with a dilute Alconox solution, tap water, and

DI water before being baked at 550°C for more than 5 hours to remove organic matter.

Labware unable to be baked was triple-rinsed with acetonitrile, ethyl acetate, and

methanol following the DI wash. Stainless steel accelerated solvent extraction cells were

59

cleaned using the non-baking approach. Endcaps were rinsed without the use of Alconox,

and then disassembled. The frit, cap insert, and snap fitting were soaked in a water bath

and then sonicated in an acetone bath for 10 min. Following re-assembly, the organic

solvent rinses were repeated.

Freeze-dried sediments were thawed and sieved prior to the extraction of

antibiotics. Sediments (1 g) were spiked with surrogates (20 ng nalidixic acid and 100 ng

13C6-sulfamethazine) in a methanol solution prior to ASE extraction. The ASE cells were

assembled with 2 glass fiber filters, a thin layer of Ottawa sand, the sediment sample,

filled with Ottawa sand, and covered with another glass fiber filter. A 50:50 methanol to

50 mM phosphate buffer (pH=7) was applied at 100°C for 5 min, allowed to sit for 5 min,

with the process repeated twice and using a rinse volume of 150%. Methanol was

removed from the ASE extract using a rotary evaporator in a 35°C water bath.

Solid phase extraction (SPE) was adapted from Meyer et al. (2000). Oasis HLB (6

mL, 200 mg, 30 µm) and MCX (6 mL, 150 mg, 30 µm) columns were used in tandem,

with the HLB column stacked on top of the MCX. Both columns were preconditioned

with 10 mL of methanol and ultrapure water. Samples were loaded and passed through

the column under a vacuum (<15 mm Hg). The HLB column was washed with 6 mL of

40:60 methanol:water while MCX was washed with water (ultrapure). Antibiotics were

eluted from the columns in tandem; first applying 3-mL of the extracts to the HLB

column, and then applying 2x5 mL methanol to the MCX stacked on the HLB column.

An addition elution of 3 mL 5% ammonium acetate in methanol was applied separately to

the MCX column. The elution was initiated with a vacuum manifold, but allowed to drip

60

by gravity once started with the eluent collected in a 15-mL glass centrifuge tube.

Internal standards (100 ng each of clinafloxacin, 13C2-erythromycin, 13C2-erythromycin-

H2O, simeton, and 13C6-sulfomethoxazole) in methanol were spiked into the eluent. The

eluents were dried under industrial grade N2 in a 40°C water bath. Samples were

dissolved into 200 µL of 20 mM ammonium acetate, and syringe-filtered (GHP, 0.4 µm)

to remove suspended particles prior to liquid chromatography tandem mass spectrometry

analysis.

Samples were analyzed on a Thermo Dionex ultimate 3000 RSLCnano system

equipped with a Thermo TSQ Vantage triple quadrupole tandem mass spectrometer

(MS/MS) in positive electrospray ionization mode. Separation of antibiotics (8 µL

injection volume) were achieved with a XSelect CSH C18 (3.5 µm, 130 Å, 50 × 2.1 mm)

column at a flow rate of 0.5 mL/min and temperature of 35 °C. The elution buffer

consisted of 0.1% formic acid in water or methanol and were applied at two gradients

(Table C.2). From 0 to 1.5 min and 5.5 to 20 min, flow was diverted to waste. Due to the

number of analytes included in the study, each sample was analyzed by three LC-MS/MS

methods that monitored for: (1) sulfonamides, 13C6-sulfamethazine, and others; (2)

tetracyclines, fluoroquinolones, and nalidixic acid; and (3) macrolides.

Analytes were detected and quantified using single reaction monitoring (SRM)

transitions (Table C.3). Confirmation SRMs were used to corroborate the identity of

quantified peaks. The mass spectrometer sensitivity varied between analyses, and thus

parameters were optimized with the infusion of 5µM simeton in 50:50 20 mM

ammonium acetate:methanol prior to each analysis. Typical values for mass spectrometer

61

parameters were: scan time 0.02 sec; scan width: 0.15; Q1/Q3: 0.7; spray voltage: 3300 V;

sheath gas pressure: 18 psi; capillary temperature: 300°C; collision pressure: 1.5 mTorr;

declustering voltage: -9 V; and tube lens: 95.

Several quality assurance and control measures were taken to assure the precision

of reported antibiotic concentrations. Antibiotic extraction efficiency from sediment was

determined for each collection site. This was achieved by spiking a methanolic solution

of antibiotics (100 ng) onto the sediment prior to ASE and calculating the mass loss due

to the extraction process. Method blanks (comprised of Ottawa sand spiked with

surrogates) were subjected to the entire extraction process and were extracted at least

every eight samples to monitor for any carryover contamination. Limits of quantification

(LOQs) and detection (LODs) were defined as S/N ratio of 10 and 3, respectively.

Antibiotic concentrations above LOQ were calculated using internal standard

methodology and were corrected according to percent recovery. Reported LOQs and

LODs were also corrected according to percent recovery.

Metals Analysis

Sediments were shipped to the Service Testing and Research Laboratory

(Wooster, Ohio) for metals testing. Microwave-assisted acid extraction (EPA 3051A) was

used with acid to extract the metals from air dried soil. Briefly, approximately 0.5 g of

the sediment was weighed into a vessel and 10 mL of 3:1 nitric to hydrochloric acid was

added. The solution was microwaved to 175°C and digested at this temperature for 10

min. Elemental analysis was conducted on the digest using an Agilent 5110 inductively

coupled plasma optical emission spectrometry (ICP-OES).

62

Data Analysis

Analysis of 16S rRNA gene sequences was performed using the QIIME pipeline

(Caporaso et al., 2010). For Illumina sequencing, paired ends were joined using the EA

utils toolkit (Aronesty, 2013). OTUs were picked at a 97% similarity using the BLAST

algorithm against the Greengenes database (version 13_5), with all taxonomic

assignments made using this database (Altschul et al., 1990; DeSantis et al., 2006).

While the GeoChip utilizes species-specific probes, the propensity of these genes

for mobilization make these assignments unreliable. To determine whether the organisms

assigned to of the GeoChip 5.0 ARG probes were present, an in-silico 16S rRNA dataset

was generated from the detected species. The unique set of organisms were filtered into a

list for each GeoChip sample (n=6). These organisms were searched for within the

SILVA SSU database (release 128) (Yilmaz et al., 2013). To remove unknown species

(e.g. ‘uncultured archaeon’ or ‘sediment metagenome’), any organism which generated

over 100 hits in the SILVA database was not included. QIIME was used to filter the

SILVA fasta file, separating out one sequence for each of the organisms. Sequences

within these files were truncated between the 515F (5’-GTGCCAGCMGCCGCGGTAA-

3’) and 806F (5’-ATTAGAWACCCBNGTAGTCC-3’) primer positions (~291 bp)

corresponding to the sequences obtained by our actual 16S rRNA sequencing runs.

Sequences with no match for either of these primers were removed from the set. The

PhyloToAST toolkit was used to identify the shared and unique OTUs related to the in-

silico ARG sequences compared with Illumina 16S rRNA gene sequences (Dabdoub et

63

al., 2016). An OTU-level biom file was prepared and used for all other analysis of the

Illumina sequences.

Further data analysis was completed using R Statistics (3.1.1) and its packages.

The normalized signals from sample replicates were averaged across individual gene

probes. Relative signals were determined as the proportion of the signal for each gene

family per the mean of the signals for all gene families of each sample. Following a brief

comparison, the duplicates included in the 2015 GeoChip set were combined by

averaging the signals detected in each sample. Relative abundance of probes was

calculated for each gene probe detected belonging to gene families and their categorical

assignments. Mean signals were calculated across the three samples. Clustering analysis

was performed using the binary Jaccard dissimilarity method in the ‘vegan’ package

(Oksanen et al., 2015). Fluidigm values were averaged across the triplicate readings for

each sediment sample. The values determined for probes common to the Fluidigm and

GeoChip were correlated (Pearson) to one another considering probe counts, mean

signals, and sum of signals (count × mean). The results from 16S rRNA analysis were

compared to the in-silico GeoChip 16S rRNA results. The number of taxa (species) and

OTUs (species clustered to a 97% similarity) that were not detected in the Illumina set

were calculated. Further, we identified the organisms which had a genus-member in the

Illumina set, but were not detected themselves.

Results

We used the GeoChip 5.0 microarray to characterize the functional metagenomes

of sediments collected from agricultural drainage channels in three Ecoregions of the

64

Western Lake Erie Basin. Two measures of site homogeneity were used in GeoChip

analysis: (1) replicates were compared for individual sites in 2015, and (2) sites were

compared in 2016. In the duplicate cores, 67-92% of probes were shared, suggesting a

high measure of similarity at specific sample locations. Once duplicates were combined,

81-88% probes were shared across sampling locations analyzed in 2015. These combined

replicates were used for all further analysis. When comparing across sites without the use

of replicates in 2016, 72-75% probes were shared. Combined, these data suggest a high

degree of functional similarity across agricultural drainage channels and that site

replicates may not be necessary to capture variations in functional diversity for this

system.

Probes detected in the carbon cycling, nitrogen cycling, metal homeostasis, and

antibiotic resistance categories were among the most abundant (Figure 3.1, Table 3.1).

The number of probes detected, and the patterns between the categories of genes

remained consistent between years and across the samples ranging from 48,762-60,564

probes. Metal homeostasis genes were the most commonly detected category, followed

by stress and carbon cycling genes. However, these patterns closely followed the number

of gene probes which are included for those gene categories. Therefore, the abundance of

gene probes may be more reflective of the GeoChip arrangement than a property of the

microbial ecosystem. Antibiotic resistance genes accounted for 12-13% (whereas

antibiotic resistance gene probes account for 9% of a GeoChip plate) of the detected

probes in all samples for both years, and these probes had some of the highest signals in

the array (Tables C.4 and C.5). The major facilitator superfamily (MFS) of antibiotic

65

transporters were prominent in both our datasets, with 12-13 of these probes having a

signal above the 99.9th percentile. In fact, eight of those probes had signals in the 99.9th

percentile for both years. Between 2015 and 2016, though, the MFS antibiotic

transporters continued to have the greatest number of detected gene probes while the

signal intensities became lower relative to other ARGs (Figure 3.1). The patterns of gene

category diversity remained relatively unchanged, indicating a continued presence of

antibiotic resistance in these sediments.

66

Table 3.1.Probe counts in the GeoChip samples by category. The number of probes in

each of these categories is provided from the manufacturer’s data for the GeoChip 5.0

GeoChip

5.0

2015 2016

CHLP OO PP CHLP OO PP

Metal Homeostasis 43,432 14,655 15,511 12,814 15,595 15,916 15,668

Carbon Cycling 26,922 8,951 9,430 7,820 9,720 9,818 9,578

Stress 26,306 8,569 9,073 7,501 9,427 9,495 9,409

Antibiotic Resistance 15,850 6,919 7,202 6,262 7,166 7,250 7,171

Organic Remediation 11,591 4,854 5,085 4,272 5,195 5,282 5,184

Virulence 6,493 1,430 1,523 1,248 1,530 1,577 1,550

Other 5,302 2,672 2,824 2,291 2,885 3,008 2,931

Nitrogen 10,380 2,397 2,516 2,092 2,627 2,663 2,634

Sulfur 4,739 1,582 1,680 1,390 1,741 1,773 1,769

Secondary Metabolism 4,032 1,517 1,574 1,365 1,583 1,571 1,576

Phosphorus 3,260 1,182 1,270 1,035 1,250 1,299 1,278

Virus 2,857 472 507 376 551 587 541

Electron Transfer 797 296 312 260 303 325 332

Total 161,961 55,496 58,507 48,726 59,573 60,564 59,621

67

Figure 3.1. Summary of the gene probe abundance and signals for functional categories

and antibiotic resistance. Relative abundance of the GeoChip categories were based off

the number of probes detected for that category divided by the total number of probes in

each sample (left). The number of probes detected in these categories were summarized

and compared across the two years of sampling (right). We further analyzed these probes

by comparing the mean signal intensities for the gene families across the two years

(inset). Error bars represent the standard deviation for the samples collected from our

three sites.

Metal homeostasis outnumbered the next most abundant gene category, carbon

cycling, by at least 5,000 probes in each sample, about 1.5-fold higher than carbon

cycling. Among these, iron and nickel gene probes were the most abundant (Table C.6).

Notably, genes associated with zinc, arsenic, copper, chromium, mercury, and silver

regulation were also abundant. The majority of metal genes were associated with

transport (either uni- or bi-directional), but with considerable numbers of detoxification

genes (Table C.7). There were few examples of metals genes used for storage or

sequestration in the data. Detoxification genes were only detected for arsenic, copper,

tellurium, chromium, and mercury. The transport genes for metals, however, may also

play a role in regulating the toxicity of metals depending on the concentrations within the

68

cytoplasm. The functional use of the metals in these communities specifically cannot be

inferred simply from their abundance.

Across both years, many similar probes yielded signals which were in the 99.9th

percentile of the data. For example, gene probes for ompR, which is involved in

osmoregulation, yielded some of the highest signals. Additionally, a single fnr gene

associated with oxygen limitation yielded a high signal in the dataset for each year.

Within the stress category, high-affinity phosphorus transporters (pst) showed high

signals for 2015 samples, but not in 2016 samples. In the carbon cycling gene category, a

pectin lyase (pel Cdeg) gene probe yielded one of the higher signals in 2015, while amyA

(amylase) and mcrA (methanogenesis) showed higher signals in 2016. Between the years,

the most prominent metal homeostasis genes changed considerably. The nikA gene

(nickel) had several probes in the 99.9th percentile of signals in 2015 samples, but these

signals were not as rich in 2016 samples. Rather, genes were present for tellurium,

potassium, manganese, and chromium. While all ARGs in the 99.9th percentile of the

2015 samples were MFS transporters, some MATE, Mex, and β-lactamase genes also

yielded high signals in the 2016 sample set. Many of these probes with the high signals

were detected between both years, but rarely were the signals as enriched in both sets.

69

The other families of antibiotic transporters were also highly abundant compared

to the MFS transporters The Mex (RND), ABC transporters, and SMR probes accounted

for the next greatest number of antibiotic resistant probes detected. MATE transporters

were also abundant, but less so than two of the β-lactamase resistance genes. The average

signals of these other transport families were also considerably high relative to many

other genes in the GeoChip. Although few probes were detected, Tet, Van, and the β-

lactamase resistance genes yielded high signals. This likely indicates low diversity, but

high abundance of these genes. However, the evidence suggests that the major

mechanism of antibiotic resistance in the sediments was from efflux-mediated protection.

Like ARGs, nearly half of the organic remediation probes included with the

GeoChip 5.0 were detected across the samples. A vast majority of these probes are

associated with the remediation of aromatic compounds (Figure C.1). Within the

aromatics group, probes were most abundant in the nitroaromatics, carboxylic acid, and

other subgroups. Outside of aromatics, probes were also abundant for the degradation of

herbicide related compounds and chlorinated compounds. Fewer probes were detected for

the pesticide related compounds, but this could also be reflective of the fact that there are

fewer probes for that subcategory. Few notable differences emerged between samples

collected from the two years in terms of the number and type of organic remediation

probes detected. The general trend showed that more probes were detected in 2016

samples in category, which is the opposite of what was observed for antibiotic resistance.

Hierarchal clustering analysis of GeoChip data revealed samples grouped by the

year in which they were collected, not by Ecoregion location (Figure 3.2). This suggests

70

temporal changes dominated differences in ecosystem services as opposed to site

differences, which is remarkable based on expected variation in physical and biotic

features between EcoRegions. In order to assess whether this trend was shared for

individual gene groups, a similar analysis was conducted on ARGs. Of the 9182 total

ARG probes in this analysis, 6689 ARGs were detected across the two years of sampling.

Within the ARGs, 85-91% of probes were shared for 2015 samples; 89-90% were shared

for 2016 samples; while 80-86% were shared by the same site across the two years. Like

the whole microarray analysis, hierarchal clustering of ARG-designated probes revealed

samples clustered by collection year, not location (Figure 3.2). This high degree of

similarity suggests the persistence of many species-specific ARGs in agricultural

drainage channel sediments.

Figure 3.2. Dendrograms of the GeoChip and antibiotic resistance gene hierarchal

clustering. Hierarchal clustering of our samples was performed based on the binary

Jaccard distance matrix for the GeoChip analysis. Clustering was performed based on the

71

entire GeoChip set (black) and with only the ARG probes (red). The relative level of

dissimilarity is printed above the dendrogram.

DNA extracted from 2016 samples were split and analyzed more in-depth by

applying the qPCR Fluidigm platform. Our Fluidigm array successfully quantified 22

genes belonging to antibiotic resistance or metal homeostasis mechanisms (Table C.8).

Gene abundance was expressed per gram of dry sediment per the amount of wet sediment

used for extraction and the volume of elution (Table C.1). The most abundant gene

detected by Fluidigm was the integrase I gene, a marker of mobile genomic elements that

commonly feature ARGs. This gene was present at more than 105 genes per gram of

sediment for each site (Figure 3.3A). Not all the genes detected by GeoChip were

included in the Fluidigm or vice versa, and integrase I was one example which was not

included in the GeoChip. The abundance of genes varied slightly across the set; for

instance, a greater abundance of blaOXA (class D), tetA, and aacA genes was detected in

the PP samples. These represent resistance to three major class of antibiotics – β-lactams,

tetracyclines, and aminoglycosides. However, there were also eight genes detected in the

OO and CHLP samples which were not detected for the PP sample. The merA gene,

providing mercury resistance, was the most abundant of metal homeostasis genes

detected with the Fluidigm analysis.

Also abundant in the Fluidigm set were the mexB and blaSHV (β-lactamase class

A) genes. While the high abundance of mexB is in agreement with the GeoChip data in

terms of probe counts and average signal intensities, the blaSHV appeared consistent with

the signal intensity of β-lactamase A. Several other genes quantified by the Fluidigm were

72

also detected by the GeoChip, which allowed for a more detailed comparison between

platforms (Table 3.1). The average abundance of common ARGs were compared against

the signal and abundance of their respective measurements (number of probes, mean signal,

and sum of signals) from the 2016 GeoChip assay (Figure 3.3B-D). While the correlation

between Fluidigm and GeoChip readings had a positive trend, none of these relationships

proved to be significant. Thus, the GeoChip results are unable to be stated as quantifiable

differences. The lack of agreement in common genes and their abundance calls to question

which tool (GeoChip, Fluidigm) is preferable for detecting and tracking changes in ARGs

in this environment.

Table 3.2. Common genes between the GeoChip and Fluidigm platforms. Several genes

included in the GeoChip and Fluidigm analyses shared a common function, although not

all were a perfect match (e.g., beta-lactamase D consists of more than just blaOXA).

GeoChip ID Fluidigm ID Function

β-lactamase A blaSHV β-lactamase class A

β-lactamase D blaOXA β-lactamase class D

chrA chrA chromate transporter

copA copA copper transporter

Mex mexB RND transporter

merA merA mercuric reductase

Qnr qnrB quinolone resistance determinant

Van vanB D-alanine--D-lactate ligase

73

Figure 3.3. Fluidigm results and comparison to GeoChip observations. (A) A Fluidigm

qPCR assay was used to quantify the gene abundance of a suite of metal homeostasis and

antibiotic resistance genes. Of 48 genes included, only 22 were successfully quantified.

DNA extractions were performed on three separate cores collected at each site, with the

standard deviation between these extractions illustrated by the error bars. The Fluidigm

gene abundances were then compared to the 2016 GeoChip results by the (B) probe

counts, (C) mean signal intensities, (D) sum of signals. Points are illustrated as the gene

name label with the coordinates for the point corresponding to the bottom-left of each

label. A zero value represents a non-detect for either platform.

Illumina sequencing of the v4 region of the 16S rRNA gene detected more than

6695 sequences per sample in 2015 and more than 15672 sequences per sample in 2016

(Table C.4). Across both years, Proteobacteria (36.8-44.3%), Bacteroidetes (5.0-14.5%),

Acidobacteria (7.3-18.0%), and Actinobacteria (1.5-13.9%) were among the dominant

groups (Figure 3.4A). Between both years, the OO samples had the lowest relative

abundance of Actinobacteria than the other two samples, but had a greater abundance of

74

other, minority phyla. Within their respective years, the OO samples also had a slightly

higher abundance of β-proteobacteria and Nitrospirae than the other samples.

Alternatively, the α-proteobacteria were more prevalent in the CHLP samples, while the

Chloroflexi were more dominant in the PP samples.

Despite the greater number of sequences in 2016, 5377 OTUs were assigned for

the 2016 set compared to 6235 OTUs for 2015. Only 41% (3363/8249) of the OTUs were

shared among years. Either this indicates a dramatic shift in communities between the

years, or is a result of using different sequencing facilities. Sample β-diversity was

compared using Jaccard matrices and hierarchal clustering, revealing 2016 samples

clustered closely together with greater dissimilarity in 2015 microbial community

structure (Figure 3.4B). The lower number of sequence may partially explain this pattern

reads in 2015. There were also differences between these data in terms of the relative

abundance of dominant OTUs (Figure 3.4C). Notably, the more dominant OTUs in the

2016 set were detected at a higher relative abundance compared with the 2015 set. Based

on the number of OTUs compared to sequences detected (Table C.9), the 2015 set did not

reach a sufficient sampling depth to detect all species present in the microbiome. Still,

there were many similarities between the phylum-level structure detected between years

at the same sampling site.

75

Figure 3.4. Illumina sequencing on the v4 region of the 16S rRNA gene was performed

on sediment DNA. (A) The taxonomic composition of the samples was compared at the

phyla level, differentiating between the subphyla of the Proteobacteria. (B) Hierarchal

clustering of these sequences was performed using the binary Jaccard distance matrix

with the relative dissimilarity shown above the dendrogram. (C) A rank abundance curve

was generated with the average abundance and standard deviation of the ranks for the

three samples collected from each year of sampling.

The GeoChip uses species-specific probes that revealed phylogenetic information

about detected ARGs. This allowed for an in-silico comparison between resistant taxa

detected in GeoChip and 16S rRNA gene analyses. The number of organisms ranged

between 1,175-1,301 per GeoChip sample (Table 3.3). However, the in-silico sets of

these organisms were condensed down to 743-795 OTUs clustered at 97% similarity. In

76

other words, many of the species included in the GeoChip could not be differentiated

between related member at the OTU97 level. By number of organisms in these ARG sets,

~17% had a matching OTU in the 2015 16S samples, while ~29% had a matching OTU

in the 2016 16S samples. However, only 10% or 19% of the ARG OTUs corresponded to

a 16S rRNA gene sequence OTU in the 2015 and 2016 samples, respectively. In other

words, many of the organisms tagged to the GeoChip ARG probes were not detected by

16S rRNA gene sequencing. Most of those community members detected through

targeting the 16S rRNA gene analysis could account for multiple probes in the GeoChip

data. Or in other words, multiple species included in the GeoChip are classified to the

same OTU at a 97% similarity.

Table 3.3. Shared GeoChip ARG lineages with taxonomies detected by Illumina

sequencing. The number of GeoChip taxa associated with the ARGs, having a match in

the SILVA database, are shown. An in-silico sequence file was developed for each

sample. The number of taxa with a match to the Illumina set for each sample was

calculated, as well as the percent of those with a match. Many of these taxa were

clustered into the same OTU, and again the number of matched OTUs were calculated.

2015 2016

CHLP OO PP CHLP OO PP

GeoChip Taxa 1230 1258 1175 1279 1301 1268

Match in 16S rRNA gene 213 216 206 372 368 371

Percent Matched 17% 17% 18% 29% 28% 29%

GeoChip OTU97 778 795 743 763 781 766

Match in 16S rRNA gene OTU97 77 78 76 146 148 148

Percent Matched 10% 10% 10% 19% 19% 19%

77

Many of the unmatched taxa belonged to genera that were detected by Illumina

sequencing. Between 41.5-43% of the unmatched OTUs in 2015 samples and 28.8-30.0%

in 2016 were related to a genus present in the Illumina set. The abundance of unmatched

taxa was predominantly found in the Gammaproteobacteria, Alphaproteobacteria,

Actinobacteria, Betaproteobacteria, and Deltaproteobacteria (Figure 3.5). Many of these

groups were better represented in the more deeply sequenced 2016 sample set, which

may explain the drop in unmatched taxa. The fact that many of these taxa had a related

genus member detected in the Illumina sequences makes it surprising that they

themselves were not detected, although this discrepancy may also have been imparted

from amplification and sequencing errors leading to the assignment of closely-related

genera.

78

Figure 3.5. Distribution of unmatched taxa between GeoChip lineages and taxonomies

detected by Illumina sequencing. There were 13 phyla (subphyla for the Proteobacteria)

which were observed in both the in-silico and 16S rRNA datasets. The relative number of

these unmatched taxa were plotted by their phylum and averaged across the three samples

in each year. The error bars represent the standard deviation for these unmatched taxa.

To further understand environment conditions in which the ARGs were detected,

antibiotics were extracted from 2016 sediments and analyzed using LC-MS/MS.

Detection limits were determined for each combination of sediment and antibiotic

ranging from 0.01-10.3 µg kg-1 dry sediment. Few antibiotics were detected in any of the

samples, with the CHLP sediments testing positive for the presence of 5 antibiotics in at

least one of its replicates, the most of any sediments (Table 3.4). Only one of the

replicates of the PP sediments tested positive for an antibiotic (ofloxacin), while the OO

sediments tested positive for 4 antibiotics in at least one of the replicates. In these OO

sediments, the quantity of trimethoprim and erythromycin were above the limit of

79

quantification, meaning that these antibiotics were detected at 1.37 and 9.2 µg kg-1 dry

sediment, respectively. However, the other replicate yielded readings below the

detectable limit for both cases, meaning antibiotics concentrations may be highly

localized within the sediment matrix.

Table 3.4. Concentration of antibiotics in the agricultural sediments. Antibiotics were

extracted from duplicate cores from the three sites, and quantified using LC-MS/MS with

detection limits specific to each sediment and antibiotic. Readings below the detection

limit are highlighted in blue. Values above the detection limit but less than the limit of

quantification are provided as estimates highlighted in blue. Readings that were

quantifiable are presented in white.

Antibiotic

Site

CHLP OO PP

µg kg-1 dry sediment

sulfapyridine <0.3 est. 0.4 <0.4 <0.4 <0.3 <0.3

sulfadiazine <0.02 <0.02 <0.02 <0.02 <0.02 <0.02

sulfamethoxazole <0.3 est. 0.5 <0.4 <0.4 <0.4 <0.4

sulfamethazine <0.1 est. 0.1 <0.2 <0.2 <0.1 <0.1

sulfachloropyridazine <0.03 <0.03 <0.04 <0.04 <0.03 <0.03

sulfadimethoxine <0.1 <0.1 <0.3 <0.3 <0.2 <0.2

carbadox <0.06 <0.06 <0.06 <0.06 <0.07 <0.07

trimethoprim <0.1 <0.1 <0.1 1.37 <0.1 <0.1

lincomycin <0.01 <0.01 <0.01 <0.01 <0.8 <0.8

tetracycline <0.04 <0.04 <0.02 <0.02 <0.02 <0.02

oxytetracycline <0.02 <0.02 <0.02 <0.02 <0.04 <0.04

chlortetracycline <0.02 <0.02 <0.03 <0.03 <0.03 <0.03

norfloxacin <3.1 <3.1 <10.3 <10.3 <4.1 <4.1

ciprofloxacin <1.3 <1.3 <4.5 <4.5 <1.7 <1.7

enrofloxacin est. 0.8 est. 1.0 <0.8 <0.8 <0.7 <0.7

ofloxacin est. 0.4 est. 0.4 <1.3 <1.3 <0.5 est. 0.8

erythromycin <0.5 est. 1.1 9.19 <0.7 <1.1 <1.1

Along with antibiotic analysis, samples from 2016 sediment were further

analyzed for metal concentrations using a microwave-assisted acid digestion reaction.

80

ICP-OES analysis was performed to quantify the concentration of 28 metal/metalloids

(Table 3.4). Nine of these elements were present at the order of g kg-1 dry sediment,

while the others were in the range mg kg-1 dry sediment. Several of the analyzed elements

harbor some known level of toxicity to microorganisms (e.g. Zn, Ni, Chromium, etc.).

Genes involved in the detoxification or transport of these metals, as well as some that

were not quantified (e.g. Hg and Ag), were detected by the GeoChip (Table C.3). Direct

detoxification, or the transport of these metals out of cell could provide resistance to

these toxic effects. Notably, the concentrations of nickel, zinc, chromium, and copper

were relatively elevated compared to the other metals, while the number of probe counts

specific to the detoxification/transports of these metals were similarly high (Figure C.2).

Detoxification genes were detected for arsenic, copper, tellurium, chromium, and

mercury, with the highest number detected in arsenic, tellurium, and mercury. While

these metals had a low concentration, and mercury was not detected by our analysis, the

abundance of detoxification genes could indicate that these are toxic levels for those

metals.

81

Table 3.5. Concentrations of trace elements extraction from the sediments. Sediments

from each site were run in duplicate and readings were performed using ICP-OES

following microwave-assisted digestion. Values below the detection limit are presented

as <DL, and values below the limit of quantification are listed as estimated values.

Analyte

Site

CHLP OO PP

g kg-1 dry sediment

Ca 60.6 81.4 22.9 22.5 31.3 29.4

Al 18.2 17.3 26.0 24.9 32.6 34.2

Fe 18.6 16.6 22.9 22.4 23.2 23.7

Mg 22.7 30.5 11.4 11.3 12.2 12.0

K 6.5 6.6 5.6 5.4 12.3 13.0

Si 1.6 1.5 1.6 1.4 1.7 1.6

Mn 0.6 0.6 1.8 1.6 0.7 0.7

P 0.5 0.6 1.5 1.4 0.7 0.7

S 0.5 0.6 1.4 1.4 0.4 0.4

mg kg-1 dry sediment

Na 249.4 295.3 228.0 238.1 382.1 405.0

Ba 115.1 130.9 251.6 241.4 171.3 181.1

Zn 67.6 93.0 131.2 125.2 69.9 72.2

Sr 112.8 129.6 55.9 54.6 65.1 65.0

V 45.7 42.2 57.7 56.2 70.0 72.6

Cr 20.9 19.9 33.1 32.1 37.0 38.6

Ni 24.5 23.9 27.3 26.2 34.9 35.9

Cu 20.4 43.6 27.8 26.9 26.2 22.8

Li 19.5 19.7 25.7 24.6 36.3 38.2

B 17.0 25.6 14.1 13.2 44.0 47.4

Pb 15.2 13.6 20.5 20.7 13.9 14.0

Co 10.9 9.0 11.4 10.9 14.7 15.5

As 10.6 7.9 9.1 8.9 7.0 7.4

Mo 7.0 6.0 2.1 2.4 5.1 4.6

Se 2.0 1.8 3.0 3.5 <0.9 <0.9

Tl est. 1.0 2.3 4.6 1.8 <0.05 1.2

Cd 1.7 1.5 2.5 2.4 1.9 1.9

Sb est. 0.7 est. 0.8 est. 0.7 <0.4 est. 1.0 1.5

Be <0.9 <0.9 <0.9 <0.9 <0.9 <0.9

82

Discussion

Agricultural drainage channels frequently fail due to the accumulation of

sediments that impede hydrological flows necessary for productive crop yields. The

formation of floodplains by these sediments have been envisioned as the basis for

reengineering drainage channels that incorporates these floodplains by widening of the

channel. The benefits from these two-stage channels include microbial processes, like

denitrification, that help to attenuate nutrient loads (Roley et al., 2012). We characterized

the functional potential of these microbial ecosystems more thoroughly using the

GeoChip 5.0 microarray, which has been used to broadly characterize the diversity of

functional genes in many environments, including urban/forested rivers (Low et al.,

2016) and wastewater bioreactors (Sun et al., 2016). It should be noted that in both of the

above-mentioned environments, the antibiotic resistance genes had many positive gene

probes and relatively high signals that made them the focus of their respective studies

(Low et al., 2016; Sun et al., 2016). The similarities between our results and those of Low

et al. (2016) are striking in that MFS antibiotic resistance genes were the most diverse

probe in the array. While Low et al. (2016) observed 4887 MFS probes (6.6% relative

abundance), we observed 4077 (6.9% relative abundance) and 3911 (6.4% relative

abundance) for our 2015 and 2016 datasets, respectively. This shows a high degree of

similarity between the ARGs of agricultural drainage sediments in Ohio with the river

sediments of the urban and forested environment of Singapore. Our GeoChip analysis

revealed that the sediments of agricultural drainage floodplains host a high diversity of

antibiotic resistance, primarily related to MFS transporters.

83

Antibiotic resistance can be propagated through migration and horizontal gene

transfer in addition to selective pressure (Niehus et al., 2015). ARGs are notorious for

being transferred, mediated through conjugation or phage, or carried on plasmids with

metal homeostasis genes (Li et al., 2015; Niehus et al., 2015). The persistence of ARGs at

the study sites described in this chapter as well as the potential for their spread to

downstream locations can be better understood through future exploration into the

mechanisms of selection, horizontal gene transfer, and migration of ARGs in sediments.

Not only were ARGs detected in sediments for three EcoRegions draining to Lake

Erie, but they remained present the year following their initial detection. In many cases,

the same ARG probes were detected in both years, indicating a persistence of these genes

at specific sampling locations. Yet, despite recurring detection of ARGs, there was little

evidence of antibiotics accumulated in sediment samples. Antibiotics detected in one or

more sediment samples (e.g. ofloxacin, enrofloxacin) were often below the limit of

quantification, on the order of 1 g/kg. Therefore, we cannot ascribe persistence of ARGs

to active selective pressure by antibiotics across the three sites. However, antibiotic

resistance can be co-selected with metal homeostasis genes, which were also highly

abundant in our sediments (Baker-Austin et al., 2006). Therefore, selection might be due

to the presence of metals in the sediment which led to co-selection of these ARGs.

In our survey of metals and metalloids in these sediments, several were detected

in all three sampling sites at concentrations approaching toxic levels for soil

microorganisms (Giller et al., 1998). For example, case studies have found cadmium,

zinc, lead, and nickel to exert some measurable toxicity (i.e., depressed soil respiration) at

84

levels of ≥ 10 mg kg-1 dry soil (Giller et al., 1998). Zinc, lead, and nickel were all well

above this level in the site sediments. In the case of zinc, concentrations were 6 to 13-fold

higher than levels know to exhibit measurable toxicity, suggesting considerable metal

toxicity may exist in these sediments. Gene probes for nickel, copper, and zinc

detoxification/transport, capable of providing resistance to metal toxicity in soil

microorganisms (Akhtar et al., 2013), were abundant in the dataset. Similarly, arsenic and

silver detoxification/transport genes were also abundant, although it is hard to define the

toxicity of these compounds without further distinguishing the forms of these elements

(e.g., redox of As; particle size of Ag) (Akhtar et al., 2013; Schlich et al., 2013). The

quantity of the metals, and the presence of detoxification/transport genes combine to

support the notion that metals could be one reason for the selection for antibiotic

resistance genes in the microbial community.

Combining our analysis of metals with antibiotics, the internal pressure within the

system for ARG selection may be applied by the metals. While antibiotics have been

shown to quickly degrade in soils, metals persist as there is no manner to remove them

from the system (Chee-Sanford et al., 2009; Hu et al., 2016). A common pathway in

which ARGs are introduced to agroecosystems is through manure application (Davies

and Davies, 2010; Udikovic-Kolic et al., 2014). To our knowledge, the fields adjacent to

our streams were not using manure-based fertilizers making this pathway dubious.

However, since metals may be introduced through inorganic fertilizers or pesticides

(Gimeno-García et al., 1996), the detected concentrations of metals are a much more

likely explanation for the abundance of ARGs in our samples. Recently, Hu et al. (2016)

85

demonstrated that nickel additions to agricultural soils resulted in the enrichment of

ARGs, including multidrug resistant genes, in fields which had no known contact with

organic fertilizers. Notably, the nickel concentrations applied in that experiment far

exceeded the concentrations of nickel measured in our sediments (>50 compared to <40

mg Ni kg-1 soil). Thus, further evaluation is needed to elucidate the toxicity levels of

metals in regard to selecting for ARGs.

In addition to metals and antibiotics, other organic compounds may be responsible

for the selection of antibiotic resistance. Organic residues may act as biocides for some

microorganisms, with antibiotic resistant organisms holding some advantage to

susceptible organisms (Pal et al., 2015; Romero et al., 2017). Recently, the mutation of

bacteria with a pesticide degrading gene revealed the selection of multidrug resistant

organisms following the exposure of soils to toxic levels of pesticides (Rangasamy et al.,

2017). Although in this case the degradation gene provided resistance to the

microorganisms, it could still be rationally argued that transport genes may have the

potential to interact with toxic levels of pesticides or other organic residues (Pal et al.,

2015; Romero et al., 2017). The abundance of organic remediation probes detected by the

GeoChip may also indicate regular exposure to these compounds in the sediments. Like

metals, the organic components of our sediments should be studied in more detail to

elucidate their potential to co-select for antibiotic resistant organisms.

While selection may be occurring, there is also evidence that could support the

case for horizontal gene transfer of the ARGs. Specifically, the intlI gene was detected at

the highest abundance of all genes included in the Fluidigm qPCR analysis. Class I

86

integrases provide the functionality for recombination for ARGs carried by plasmids or

other mobile genomic elements (Goldstein et al., 2001). Thus, there should be some

concern over the spread of ARGs within the agricultural floodplains. Due to this high

abundance of intlI, it was important to determine whether the organisms identified by the

GeoChip would be also be detected by 16S rRNA sequencing.

The same DNA extracts used in the GeoChip and Fluidigm analyses were also

used for lllumina sequencing of the 16S rRNA gene sequencing for both years. Notably,

analysis of the 16S gene sequences revealed a much more disparate community compared

to the GeoChip microarray. In other words, the functionality between our sampling

locations and between years were much more similar to one another than the community

members revealed by Illumina sequencing. To compare the phylogenetic information

contained in the GeoChip to the 16S analysis we first had to generate an in-silico set of

16S DNA from the SILVA database and process that through QIIME to account for the

0.97 clustering of our reference database used to pick OTUs. Through this simulation,

less than 30% of the GeoChip taxa were identified in the actual 16S analysis results, a

rather poor overlap. Eighteen phyla (including subphyla for Proteobacteria) were shared

between both the 16S and GeoChip OTUs. Notably, three Proteobacteria subphyla (α-/β-

/γ-), Actinobacteria, and Firmicutes were the phyla which contained the greatest number

of unexplained taxa matched between the GeoChip and 16S data. Known biases have

found the v4 region – covered by our sequencing primers - to limit the detection of

Actinobacteria and Verrucomicrobia through our Illumina sequencing method (Guo et

al., 2015). While we could account for the lack of matches in Actinobacteria from primer

87

bias, we are unable to explain the lack of matches for the other phyla. Since ARGs are

known to be highly mobile, the lack of matches between Illumina sequences and the in-

silico GeoChip set may be accounted for due to other organisms harboring these genes

than those that are assigned to the probes. Combining the abundance of intI with the lack

of matches between the 16S community members and GeoChip taxa may reveal that

ARGs were highly mobilized in these sediments or the ecosystem from which the

sediments originated. Therefore, future analyses should consider screening the cultivable

microbial community to identify those which demonstrate antibiotic resistance.

The presence, persistence, and fate of ARGs in our agricultural drainage ways is

of major concern because of the threat that antibiotic resistance poses as a health concern

to downstream human communities. However, it is important to also consider that the

majority of the ARGs exposed by our analysis were the MFS transport systems. Efflux-

mediated resistance does not necessarily mean that an organism has resistance to

antibiotic at a clinical level, but in these systems, can provide resistance to a greater range

of compounds (Kumar et al., 2013). Cultivation methods could be applied to decipher the

level of resistance of the organisms in our sediments, such as was performed by Low et

al. (2016). Another important fact to consider is that the development of two-stage

channels is to attenuate the sediment loads reaching downstream waters (Jayakaran et al.,

2010). Perhaps then these systems could limit the transport of ARGs to downstream

waters, if they are capturing antibiotic resistant organisms attached to sediments. The

continued presence of ARGs at these sites, though, also means that these sediments could

act as a reservoir of these genes in the environment.

88

Conclusions

The purpose of the research covered by this dissertation was to gain a better

understanding of pollutants in the Lake Erie watershed. Although the drainage basins

sampled in this dissertation represent a minor fraction draining into this important Great

Lake, the data and analyses performed here have applicability to world issues and could

be used in other watersheds experiencing eutrophication and harmful algal bloom issues.

Agricultural pollution is a problem across the world and the issues presented here are not

isolated to Lake Erie or North America.

The objective of Chapters 1 and 2 was to examine develop a protocol that could

be used to distinguish between sources of nutrient pollution and monitor for their

presence in the Lake Erie watershed. New information is need by regional managers to

identify the leading sources so that management practices can be targeted at the leading

sources. In Chapter 1, disparate organic matter and organic phosphorus signatures were

characterized for several point and nonpoint sources. Marker m/z values were proposed to

help discriminate between these sources. In Chapter 2, we analyzed the transport of DOM

through the tributary network. DOM was highly similar throughout the watershed,

although broad characteristics changed considerably. We found that the most prominent

features were more consistently detected during their transport from upstream to

downstream samples. Additionally, the DOM spectra appeared to mix linearly at several

confluence points. The data from these chapters suggest that DOM can differentiate the

source of its origin. However, we discovered that the most prominent m/z values were

reliably detected through its transport. Since the original markers proposed for tracking

89

pollutant sources were typically detected with low peak magnitudes, tour selection

process of markers should be reevaluated to seek marker DOM more likely to be detected

throughout the watershed during transport. Marker compounds could be expanded to

include compounds other than just DOP, which should provide more options through the

inclusion of the abundant CHO and CHON formulae.

In Chapter 3, sediments collected from agricultural floodplains developing within

drainage channels contained many ARGs. It is difficult to define the abundance of these

genes as different technologies showed a disagreement between each other. It appears

that there is a high diversity of MFS antibiotic transport genes in these sediments. There

must be further work to elucidate the abundance of these genes, but that will require

greater research into the interpretation of results obtained through different technologies.

The presence of integrase genes suggested the potential for HGT or mobilization of

ARGs in our floodplains, and the lack of matching taxa in the 16S to GeoChip datasets

suggests these genes have already been transferred to other microbial community

members. Few antibiotics were detected in our sediments, and so we argue that selection

must be occurring through some other mechanism. It is reasonable to believe that MFS

transporters may instead be co-selected for due to the presence of metals at putatively

toxic concentrations. Seemingly high abundance of metal resistance genes further

supports this theory. Concern should be given to the presence of these ARGs in

agricultural channels, but we must gain a better understanding over how these genes

arrived at these sediments, why they may persist, and how they may be transported

further downstream.

90

Our analyses were used to better define a well-known and established issue

focused on the eutrophication of Lake Erie through nutrient loads, but also may have

exposed that the tributaries may act as conduits for antibiotic resistance. The analysis of

DOM shows some promise in being able to identify the presence of source

contamination. Continued development of this approach could hold value to the Lake

Erie watershed, but also to coastal areas like the Gulf of Mexico or Chesapeake Bay

when distinguishing sources of nutrient pollution is necessary. Alternatively, we present

evidence that antibiotic resistance must be monitored for its transport through waterways

as these genes present a rising health concern.

Future research needs to improve upon the results and conclusions drawn from the

research presented here. In terms of the mass spectrometric analysis, the signatures were

characterized for only a few of the many sources in the regions. The analysis could be

expanded to include other sources of concern (CSOs, septic tanks, lawns, etc.) and runoff

from manure-fertilized fields. This research could improve our knowledge over the

signatures to look for in this, or other nutrient-impaired waterways. Additionally, the

understanding of transport and mixing of DOM could be improved through controlled

laboratory and field experiments. For example, direct mixing of sample could be

performed in the laboratory while organic nitrogen and phosphorus compounds could be

added as tracers in the natural environment. For antibiotic resistance, value would be

added from the cultivation and phenotyping of the microorganisms in the sediments.

Specifically, antibiotic resistant pathogens should be targeted as they pose the greatest

threat to human health. Finally, the origin, transport, and fate of antibiotic resistance in

91

agricultural waters should be studied more in depth. For example, while we demonstrate

that the sediments in the channels were enriched with antibiotic resistance we do not

know whether this is acting as a sink or source for further downstream transport. Field

studies should be designed to answer whether these systems are removing antibiotic

resistant organisms from downstream transport, the localized selection for or against

antibiotic resistance within the sediments, and the potential for antibiotic resistance to be

dispersed during flow events. My future research will explore many of these areas under

the common theme of exploring agricultural contaminants.

92

References

Adams M.M., M.R. Gomez-Garcia, A.R. Grossman and D. Bhaya. 2008. Phosphorus

deprivation responses and phosphonate utilization in a thermophilic Synechococcus

sp. from microbial mats. J. Bacteriol. 190:8171-84.

Akhtar M.S., B. Chali and T. Azam. 2013. Bioremediation of arsenic and lead by plants

and microbes from contaminated soil. Research in Plant Sciences 1:68-73.

Altschul S.F., W. Gish, W. Miller, E.W. Myers and D.J. Lipman. 1990. Basic local

alignment search tool. J. Mol. Biol. 215:403-10.

Arango C.P., J.L. Tank, J.L. Schaller, T.V. Royer, M.J. Bernot and M.B. David. 2007.

Benthic organic carbon influences denitrification in streams with high nitrate

concentration. Freshwat. Biol. 52:1210-1222.

Arnold W.A., K. Longnecker, K.D. Kroeger and E.B. Kujawinski. 2014. Molecular

signature of organic nitrogen in septic-impacted groundwater. Environ. Sci. :

Processes Impacts2400.

Aronesty E. 2013. Comparison of sequencing utility programs. The Open Bioinformatics

Journal 7:.

93

Baker D. 2011. The sources and transport of bioavailable phosphorus to lake erie. 1-30.

Baker D. 2010. Trends in bioavailable phosphorus loading to lake erie. 1-21.

Baker D.B., R. Confesor, D.E. Ewing, L.T. Johnson, J.W. Kramer and B.J. Merryfield.

2014. Phosphorus loading to lake erie from the maumee, sandusky and cuyahoga

rivers: The importance of bioavailability. J. Great Lakes Res. 40:502-517.

Baker-Austin C., M.S. Wright, R. Stepanauskas and J.V. McArthur. 2006. Co-selection

of antibiotic and metal resistance. Trends Microbiol. 14:176-182.

Baldwin D.S. 1998. Reactive “organic” phosphorus revisited. Water Res. 32:2265-2270.

Banerjee S. and S. Mazumdar. 2012. Electrospray ionization mass spectrometry: A

technique to access the information beyond the molecular weight of the analyte.

International Journal of Analytical Chemistry 2012:.

Barthold F.K., C. Tyralla, K. Schneider, K.B. Vaché, H. Frede and L. Breuer. 2011. How

many tracers do we need for end member mixing analysis (EMMA)? A sensitivity

analysis. Water Resour. Res. 47:.

Bartos J.M., B.L. Boggs, J.H. Falls and S.A. Siegel. 2014. Determination of phosphorus

and potassium in commercial inorganic fertilizers by inductively coupled plasma-

optical emission spectrometry: Single-laboratory validation. J. AOAC Int. 97:.

94

Berg A.S. and B.C. Joern. 2006. Sorption dynamics of organic and inorganic phosphorus

compounds in soil. J. Environ. Qual. 35:.

Bhatia M.P., S.B. Das, K. Longnecker, M.A. Charette and E.B. Kujawinski. 2010.

Molecular characterization of dissolved organic matter associated with the greenland

ice sheet. Geochim. Cosmochim. Acta 74:3768-3784.

Bjorkman K.M. and D.M. Karl. 2003. Bioavailability of dissolved organic phosphorus in

the euphotic zone at station ALOHA, north pacific subtropical gyre. Limnology and

Oceanography 48:1049-1057.

Boye K., V. Noel, S.E. Bone, J.R. Bargar, K. Boye, S. Fendorf, M.M. Tfaily and K.H.

Williams. 2017. Thermodynamically controlled preservation of organic carbon in

floodplains. Nature Geoscience 10:415-419.

Brooker M.R., G. Bohrer and P.J. Mouser. 2014. Variations in potential CH4 flux and

CO2 respiration from freshwater wetland sediments that differ by microsite location,

depth and temperature. Ecological Engineering 72:84-94.

Brooker M.R. 2018. Physical and chemical characterization of self-developing

agricultural floodplains. The Ohio State University, .

Caporaso J.G., C.L. Lauber, W.A. Walters, D. Berg-Lyons, J. Huntley, N. Fierer, S.M.

Owens, J. Betley, L. Fraser, M. Bauer, N. Gormley, J.A. Gilbert, G. Smith and R.

95

Knight. 2012. Ultra-high-throughput microbial community analysis on the illumina

HiSeq and MiSeq platforms. Isme 6:.

Caporaso J.G., J. Kuczynski, J. Stombaugh, K. Bittinger, F.D. Bushman, E.K. Costello,

N. Fierer, A.G. Pena, J.K. Goodrich, J.I. Gordon, G.A. Huttley, S.T. Kelley, D.

Knights, J.E. Koenig, R.E. Ley, C.A. Lozupone, D. McDonald, B.D. Muegge, M.

Pirrung, J. Reeder, J.R. Sevinsky, P.J. Tumbaugh, W.A. Walters, J. Widmann, T.

Yatsunenko, J. Zaneveld and R. Knight. 2010. QIIME allows analysis of high-

throughput community sequencing data. Nature Methods 7:.

Center for Disease Dynamics, Economics & Policy. 2015. State of the world’s

antibiotics. CDDEP, Washington, D.C.

Chee-Sanford J.C., R.I. Aminov, I.J. Krapac, N. Garrigues-Jeanjean and R.I. Mackie.

2001. Occurrence and diversity of tetracycline resistance genes in lagoons and

groundwater underlying two swine production facilities. Appl. Environ. Microbiol.

67:1494-1502.

Chee-Sanford J.C., R.I. Mackie, S. Koike, I.G. Krapac, Y.F. Lin, A.C. Yannarell, S.

Maxwell and R.I. Aminov. 2009. Fate and transport of antibiotic residues and

antibiotic resistance genes following land application of manure waste. J. Environ.

Qual. 38:.

96

Cong J., H. Lu, D. Li, Y. Zhang, J. Cong, X. Liu, H. Xu, Y. Li and Y. Deng. 2015.

Analyses of the influencing factors of soil microbial functional gene diversity in

tropical rainforest based on GeoChip 5.0. Genomics Data 5:397-398.

Conley D.J., H.W. Paerl, R.W. Howarth, D.F. Boesch, S.P. Seitzinger, K.E. Havens, C.

Lancelot and G.E. Likens. 2009. Controlling eutrophication: Nitrogen and

phosphorus. Science 123:1014-1015.

Conroy J.D., L. Boegman, H. Zhang, W.J. Edwards and D.A. Culver. 2011. "Dead zone"

dynamics in lake erie: The importance of weather and sampling intensity for

calculated hypolimnetic oxygen depletion rates. Aquat. Sci. 73:289-304.

Cooper W.T., J.M. Llewelyn, G.L. Bennett and V.J.M. Salters. 2005. Mass spectrometry

of natural organic phosphorus. Talanta 66:348-358.

Dabdoub S.M., M.L. Fellows, A.D. Paropkari, M.R. Mason, S.S. Huja, A.A. Tsigarida

and P.S. Kumar. 2016. PhyloToAST: Bioinformatics tools for species-level analysis

and visualization of complex microbial datasets. Scientific Reports 6:29123.

Davies J. and D. Davies. 2010. Origins and evolution of antibiotic resistance. Microbiol.

Mol. Biol. Rev. 74:417-433.

DeSantis T.Z., P. Hugenholtz, N. Larsen, M. Rojas, E.L. Brodie, K. Keller, T. Huber, D.

Dalevi, P. Hu and G.L. Andersen. 2006. Greengenes, a chimera-checked 16S rRNA

97

gene database and workbench compatible with ARB. Appl. Environ. Microbiol.

72:5069-72.

Doctor D.H., E.C. Alexander, M. Petrič, J. Kogovšek, J. Urbanc, S. Lojen and W.

Stichler. 2006. Quantification of karst aquifer discharge components during storm

events through end-member mixing analysis using natural chemistry and stable

isotopes as tracers. Hydrogeol. J. 14:1171-1191.

Dodd R.J. and A.N. Sharpley. 2015. Recognizing the role of soil organic phosphorus in

soil fertility and water quality. Resources, Conservation and Recycling 105:282-293.

Douglass J.F., M. Radosevich and O.H. Tuovinen. 2015. Molecular analysis of atrazine-

degrading bacteria and catabolic genes in the water column and sediment of a

created wetland in an agricultural/urban watershed. Ecol. Eng. 83:405-412.

Elsbury K.E., A. Paytan, N.E. Ostrom, C. Kendall, M.B. Young, K. McLaughlin, M.E.

Rollog and S. Watson. 2009. Using oxygen isotopes of phosphate to trace

phosphorus sources and cycling in lake erie. Environ. Sci. Technol. 43:3108-14.

Elsenbeer H., D. Lorieri and M. Bonell. 1995. Mixing model approaches to estimate

storm flow sources in an overland flow‐dominated tropical rain forest catchment.

Water Resour. Res. 31:2267-2278.

Feng L., J. Xu, S. Kang, X. Li, Y. Li, B. Jiang and Q. Shi. 2016. Chemical composition

of microbe-derived dissolved organic matter in cryoconite in tibetan plateau glaciers:

98

Insights from fourier transform ion cyclotron resonance mass spectrometry analysis.

Environ. Sci. Technol. 50:13215-13223.

Frey S.K., E. Topp, I.U. Khan, B.R. Ball, M. Edwards, N. Gottschall, M. Sunohara and

D.R. Lapen. 2015. Quantitative campylobacter spp., antibiotic resistance genes, and

veterinary antibiotics in surface and ground water following manure application:

Influence of tile drainage control. Sci. Total Environ. 532:138-153.

Ghosh S. and T.M. LaPara. 2007. The effects of subtherapeutic antibiotic use in farm

animals on the proliferation and persistence of antibiotic resistance among soil

bacteria. ISME Journal 1:191-203.

Giller K.E., E. Witter and S.P. Mcgrath. 1998. Toxicity of heavy metals to

microorganisms and microbial processes in agricultural soils: A review. Soil Biol.

Biochem. 30:1389-1414.

Gimeno-García E., V. Andreu and R. Boluda. 1996. Heavy metals incidence in the

application of inorganic fertilizers and pesticides to rice farming soils.

Environmental Pollution 92:19-25.

Goldstein C., M.D. Lee, S. Sanchez, C. Hudson, B. Phillips, B. Register, M. Grady, C.

Liebert, A.O. Summers, D.G. White and J.J. Maurer. 2001. Incidence of class 1 and

2 integrases in clinical and commensal bacteria from livestock, companion animals,

and exotics. Antimicrob. Agents Chemother. 45:723-726.

99

Gonsior M., M. Zwartjes, W.J. Cooper, W. Song, K.P. Ishida, L.Y. Tseng, M.K. Jeung,

D. Rosso, N. Hertkorn and P. Schmitt-Kopplin. 2011. Molecular characterization of

effluent organic matter identified by ultrahigh resolution mass spectrometry. Water

Res. 45:2943-2953.

Guo J., J.R. Cole, Q. Zhang, C.T. Brown and J.M. Tiedje. 2015. Microbial community

analysis with ribosomal gene fragments from shotgun metagenomes. Appl. Environ.

Microbiol. 82:157-166.

Hansman R.L., T. Dittmar and G.J. Herndl. 2015. Conservation of dissolved organic

matter molecular composition during mixing of the deep water masses of the

northeast atlantic ocean. Mar. Chem. 177:288-297.

Hobman J.L. and L.C. Crossman. 2015. Bacterial antimicrobial metal ion resistance. J.

Med. Microbiol. 64:471-497.

Hu H., J. Wang, J. Li, X. Shi, Y. Ma, D. Chen and J. He. 2016. Long-term nickel

contamination increases the occurrence of antibiotic resistance genes in agricultural

soils. Environ. Sci. Technol. 51:790-800.

Hudson N., A. Baker and D. Reynolds. 2007. Fluorescence analysis of dissolved organic

matter in natural, waste and polluted waters—a review. River Research and

Applications 23:631-649.

100

Jayakaran A.D., D.E. Mecklenburg, J.D. . Witter, A.D. Ward and G.E. and Powell. 2010.

Fluvial processes in agricultural ditches inthe north central region of the united states

and implica-tions for their management. p. 195-222. In M.T. Moore and R. and

Kroger (eds.) Agricultural DrainageDitches: Mitigation wetlands for the 21st

century. Research Signpost, Karala, India.

Johnson T.A., R.D. Stedtfeld, Q. Wang, J.R. Cole, S.A. Hashsham, T. Looft, Y.G. Zhu

and J.M. Tiedje. 2016. Clusters of antibiotic resistance genes enriched together stay

together in swine agriculture. mBio 7:02214-15.

Kamga A.W., F. Behar and P.G. Hatcher. 2014. Quantitative analysis of long chain fatty

acids present in a type I kerogen using electrospray ionization fourier transform ion

cyclotron resonance mass spectrometry: Compared with BF3/MeOH

methylation/GC-FID. Journal of the American Society for Mass Spectrometry

25:880-890.

Karl D.M. 2014. Microbially mediated transformations of phosphorus in the sea: New

views of an old cycle. Annual Review of Marine Science 6:279-337.

Karl, David M.,Tien, Georgia,. 1992. MAGIC: A sensitive and precise method for

measuring dissolved phosphorus in aquatic environments. LNO Limnology and

Oceanography 37:105-116.

101

Kekacs D., B.D. Drollette, M.R. Brooker, D.L. Plata and P.J. Mouser. 2015. Aerobic

biodegradation of organic compounds in hydraulic fracturing fluids. Biodegradation

26:271-87.

Kellerman A.M., D.N. Kothawala, T. Dittmar and L.J. Tranvik. 2015. Persistence of

dissolved organic matter in lakes related to its molecular characteristics. Nature

Geoscience 8:454.

Kerrigan J.F., K.D. Sandberg, D.R. Engstrom, T.M. LaPara and W.A. Arnold. 2017.

Sedimentary record of antibiotic accumulation in minnesota lakes. Science of the

Total Environment.

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

drainage to watershed-scale phosphorus transport. J. Environ. Qual. 44:486-494.

Kruse J., M. Abraham, W. Amelung, C. Baum, R. Bol, O. Kuhn, H. Lewandowski, J.

Niederberger, Y. Oelmann, C. Ruger, J. Santner, M. Siebers, N. Siebers, M. Spohn,

J. Vestergren, A. Vogts and P. Leinweber. 2015. Innovative methods in soil

phosphorus research: A review. Journal of Plant Nutrition and Soil Science 178:43-

88.

Kujawinski E.B. and M.D. Behn. 2006. Automated analysis of electrospray ionization

fourier transform ion cyclotron resonance mass spectra of natural organic matter.

Anal. Chem. 78:4363-4373.

102

Kujawinski E.B., K. Longnecker, N.V. Blough, R. Del Vecchio, L. Finlay, J.B. Kitner

and S.J. Giovannoni. 2009. Identification of possible source markers in marine

dissolved organic matter using ultrahigh resolution mass spectrometry. Geochim.

Cosmochim. Acta 73:4384-4399.

Kumar S., M.M. Mukherjee and M.F. Varela. 2013. Modulation of bacterial multidrug

resistance efflux pumps of the major facilitator superfamily. International Journal of

Bacteriology 2013:1-15.

Larsen L., J. Harvey, K. Skalak and M. Goodman. 2015. Fluorescence‐based source

tracking of organic sediment in restored and unrestored urban streams. Limnol.

Oceanogr. 60:1439-1461.

Lee M.S. and E.H. Kerns. 1999. LC/MS applications in drug development. Mass

Spectrom. Rev. 18:187-279.

Li A.D., L.G. Li and T. Zhang. 2015. Exploring antibiotic resistance genes and metal

resistance genes in plasmid metagenomes from wastewater treatment plants. Front.

Microbiol. 6:1025.

Low A., C. Ng and J. He. 2016. Identification of antibiotic resistant bacteria community

and a GeoChip based study of resistome in urban watersheds. Water Res. 106:330-

338.

103

Lu Y.H., X. Li, R. Mesfioui, J.E. Bauer, R.M. Chambers, E.A. Canuel and P.G. Hatcher.

2015. Use of ESI-FTICR-MS to characterize dissolved organic matter in headwater

streams draining forest-dominated and pasture-dominated watersheds. PloS One

10:e0145639.

Lusk M.G. and G.S. Toor. 2016. Dissolved organic nitrogen in urban streams:

Biodegradability and molecular composition studies. Water Res. 96:225-235.

Maizel A.C. and C.K. Remucal. 2017. The effect of advanced secondary municipal

wastewater treatment on the molecular composition of dissolved organic matter.

Water Research 122:42-52.

Mantini D., F. Petrucci, D. Pieragostino, P. Del Boccio, M. Di Nicola, C. Di Ilio, G.

Federici, P. Sacchetta, S. Comani and A. Urbani. 2007. LIMPIC: A computational

method for the separation of protein MALDI-TOF-MS signals from noise. BMC

Bioinformatics 8:1.

Martinez J.L. 2009. Environmental pollution by antibiotics and by antibiotic resistance

determinants. Environmental Pollution 157:2893-2902.

McManus P.S., V.O. Stockwell, G.W. Sundin and A.L. Jones. 2002. Antibiotic use in

plant agriculture. Annu. Rev. Phytopathol. 40:443-465.

Medeiros P.M., M. Seidel, N.D. Ward, E.J. Carpenter, H.R. Gomes, J. Niggemann, A.V.

Krusche, J.E. Richey, P.L. Yager and T. Dittmar. 2015. Fate of the amazon river

104

dissolved organic matter in the tropical atlantic ocean. Global Biogeochem. Cycles

29:677-690.

Mesfioui R., N.G. Love, D.A. Bronk, M.R. Mulholland and P.G. Hatcher. 2012.

Reactivity and chemical characterization of effluent organic nitrogen from

wastewater treatment plants determined by fourier transform ion cyclotron resonance

mass spectrometry. Water Research 46:622-634.

Meyer M.T., J.E. Bumgarner, J.L. Varns, J.V. Daughtridge, E.M. Thurman and K.A.

Hostetler. 2000. Use of radioimmunoassay as a screen for antibiotics in confined

animal feeding operations and confirmation by liquid chromatography/mass

spectrometry. Science of the Total Environment 248:181-187.

Michalak A.M., E.J. Anderson, D. Beletsky, S. Boland, N.S. Bosch, T.B. Bridgeman,

J.D. Chaffin, K. Cho, R. Confesor, I. Daloglu, J.V. Depinto, M.A. Evans, G.L.

Fahnenstiel, L. He, J.C. Ho, L. Jenkins, T.H. Johengen, K.C. Kuo, E. Laporte, X.

Liu, M.R. McWilliams, M.R. Moore, D.J. Posselt, R.P. Richards, D. Scavia, A.L.

Steiner, E. Verhamme, D.M. Wright and M.A. Zagorski. 2013. Record-setting algal

bloom in lake erie caused by agricultural and meteorological trends consistent with

expected future conditions. Proc. Natl. Acad. Sci. U. S. A. 110:6448-52.

Minor E.C., C.J. Steinbring, K. Longnecker and E.B. Kujawinski. 2012. Characterization

of dissolved organic matter in lake superior and its watershed using ultrahigh

resolution mass spectrometry. Org. Geochem. 43:1-11.

105

Monaghan, E. J., Ruttenberg,K.C.,. 1999. Dissolved organic phosphorus in the coastal

ocean: Reassessment of available methods and seasonal phosphorus profiles from

the eel river shelf. LNO Limnology and Oceanography 44:1702-1714.

Munir M., K. Wong and I. Xagoraraki. 2011. Release of antibiotic resistant bacteria and

genes in the effluent and biosolids of five wastewater utilities in michigan. WR

Water Research 45:681-693.

Negreanu Y., Z. Pasternak, E. Jurkevitch and E. Cytryn. 2012. Impact of treated

wastewater irrigation on antibiotic resistance in agricultural soils. Environ. Sci.

Technol. 46:4800-4808.

Niehus R., S. Mitri, A.G. Fletcher and K.R. Foster. 2015. Migration and horizontal gene

transfer divide microbial genomes into multiple niches. Nature Communications 6:.

Ohio Lake Erie Phosphorus Task Force. 2013. Final report II. Ohio EPA, .

Ohio Lake Erie Phosphorus Task Force. 2010. Final report I. Ohio Environmental

Protection Agency, Division of Surface Water, Columbus.

Ohno T. and P.E. Ohno. 2013. Influence of heteroatom pre-selection on the molecular

formula assignment of soil organic matter components determined by ultrahigh

resolution mass spectrometry. Analytical and Bioanalytical Chemistry 405:3299-

3306.

106

Ohno T., R.L. Sleighter and P.G. Hatcher. 2016. Comparative study of organic matter

chemical characterization using negative and positive mode electrospray ionization

ultrahigh-resolution mass spectrometry. Analytical and Bioanalytical Chemistry

408:2497-2504.

Ohno T., J. Chorover, A. Omoike and J. Hunt. 2007. Molecular weight and humification

index as predictors of adsorption for plant- and manure-derived dissolved organic

matter to goethite. European Journal of Soil Science 58:125-132.

Oksanen J., F.G. Blanchet, R. Kindt, P. Legendre, P.R. Minchin, R. O’hara, G.L.

Simpson, P. Solymos, M. Stevens and H. Wagner. 2015. Vegan: Community

ecology package. R package version 2.0-10. 2013. There is no Corresponding

Record for this Reference.

Omernik J.M. 1986. Ecoregions of the conterminous united states.

Pal C., J. Bengtsson-Palme, E. Kristiansson and D.J. Larsson. 2015. Co-occurrence of

resistance genes to antibiotics, biocides and metals reveals novel insights into their

co-selection potential. BMC Genomics 16:964.

Pasek M.A., J.M. Sampson and Z. Atlas. 2014. Redox chemistry in the phosphorus

biogeochemical cycle. Proc. Natl. Acad. Sci. U. S. A. 111:.

107

Powell G.E., A.D. Ward, D.E. Mecklenburg, J. Draper and W. Word. 2007. Special

section: Drainage ditches - two-stage channel systems: Part 2, case studies. Journal

of Soil and Water Conservation. 62:286.

Rabalais N.N., R.E. Turner and W.J. Wiseman. 2002. Gulf of mexico hypoxia, aka "the

dead zone". Annu. Rev. Ecol. Syst. 33:235-263.

Raeke J., O.J. Lechtenfeld, M. Wagner, P. Herzsprung and T. Reemtsma. 2016.

Selectivity of solid phase extraction of freshwater dissolved organic matter and its

effect on ultrahigh resolution mass spectra. Environmental Science: Processes &

Impacts 18:918-927.

Rangasamy K., M. Athiappan, N. Devarajan and J.A. Parray. 2017. Emergence of multi

drug resistance among soil bacteria exposing to insecticides. Microbial Pathogenesis

105:153-165.

Ries III K.G., J.K. Newson, M.J. Smith, J.D. Guthrie, P.A. Steeves, T.L. Haluska, K.R.

Kolb, R.F. Thompson, R.D. Santoro and H.W. Vraga. 2017. StreamStats, Version 4.

Roley S.S., J.L. Tank and M.A. Williams. 2012. Hydrologic connectivity increases

denitrification in the hyporheic zone and restored floodplains of an agricultural

stream. Journal of Geophysical Research. 117:.

Romero J.L., M.J. Grande Burgos, R. Pérez-Pulido, A. Gálvez and R. Lucas. 2017.

Resistance to antibiotics, biocides, preservatives and metals in bacteria isolated from

108

seafoods: Co-selection of strains resistant or tolerant to different classes of

compounds. Frontiers in Microbiology 8:1650.

Rothrock M.J., P.L. Keen, K.L. Cook, L.M. Durso, A.M. Franklin and R.S. Dungan.

2016. How should we be determining background and baseline antibiotic resistance

levels in agroecosystem research? Journal of Environment Quality 45:420-431.

Ruttenberg KC D.S. 2012. Dissolved organic phosphorus production during simulated

phytoplankton blooms in a coastal upwelling system. Frontiers in Microbiology 3:.

Ruttenberg K.C. and S.T. Dyhrman. 2005. Temporal and spatial variability of dissolved

organic and inorganic phosphorus, and metrics of phosphorus bioavailability in an

upwelling-dominated coastal system. Journal of Geophysical Research: Oceans 110:.

Schlich K., T. Klawonn, K. Terytze and K. Hund-Rinke. 2013. Hazard assessment of a

silver nanoparticle in soil applied via sewage sludge. Environmental Sciences

Europe 25:17.

Schmitt H., K. Stoob, G. Hamscher, E. Smit and W. Seinen. 2006. Tetracyclines and

tetracycline resistance in agricultural soils: Microcosm and field studies. Microb.

Ecol. 51:267-276.

Sharma R., R.W. Bella and M.T.F. Wong. 2017. Dissolved reactive phosphorus played a

limited role in phosphorus transport via runoff, throughflow and leaching on

contrasting cropping soils from southwest australia. Sci. Total Environ. 577:33-44.

109

Southam A.D., T.G. Payne, H.J. Cooper, T.N. Arvanitis and M.R. Viant. 2007. Dynamic

range and mass accuracy of wide-scan direct infusion nanoelectrospray fourier

transform ion cyclotron resonance mass spectrometry-based metabolomics increased

by the spectral stitching method. Anal. Chem. 79:4595-4602.

Steffen M.M., T.W. Davis, R.M.L. McKay, G.S. Bullerjahn, L.E. Krausfeldt, J.M.A.

Stough, M.L. Neitzey, N.E. Gilbert, G.L. Boyer, T.H. Johengen, D.C. Gossiaux,

A.M. Burtner, D. Palladino, M.D. Rowe, G.J. Dick, K.A. Meyer, S. Levy, B.E.

Boone, R.P. Stumpf, T.T. Wynne, P.V. Zimba, D. Gutierrez and S.W. Wilhelm.

2017. Ecophysiological examination of the lake erie microcystis bloom in 2014:

Linkages between biology and the water supply shutdown of toledo, OH. Environ.

Sci. Technol.

Storteboom H., M. Arabi, J.G. Davis, B. Crimi and A. Pruden. 2010. Tracking antibiotic

resistance genes in the south platte river basin using molecular signatures of urban,

agricultural, and pristine sources. Environ. Sci. Technol. 44:7397-7404.

Stubbins A., R.G.M. Spencer, H. Chen, P.G. Hatcher, K. Mopper, P.J. Hernes, V.L.

Mwamba, A.M. Mangangu, J.N. Wabakanghanzi and J. Six. 2010. Illuminated

darkness: Molecular signatures of congo river dissolved organic matter and its

photochemical alteration as revealed by ultrahigh precision mass spectrometry.

Limnol. Oceanogr. 55:1467-1477.

110

Stumpf R.P., T.T. Wynne, D.B. Baker and G.L. Fahnenstiel. 2012. Interannual variability

of cyanobacterial blooms in lake erie. PloS One 7:e42444.

Sun Y., Y. Shen, P. Liang, J. Zhou, Y. Yang and X. Huang. 2016. Multiple antibiotic

resistance genes distribution in ten large-scale membrane bioreactors for municipal

wastewater treatment. Bioresour. Technol. 222:100-106.

Udikovic-Kolic N., F. Wichmann, N.A. Broderick and J. Handelsman. 2014. Bloom of

resident antibiotic-resistant bacteria in soil following manure fertilization. Proc. Natl.

Acad. Sci. U. S. A. 111:15202-15207.

Vymazal J. and T. Březinová. 2015. The use of constructed wetlands for removal of

pesticides from agricultural runoff and drainage: A review. Environment

International 75:11-20.

Wilson H.F. and M.A. Xenopoulos. 2009. Effects of agricultural land use on the

composition of fluvial dissolved organic matter. Nature Geoscience 2:37.

Yang L., S. Chang, H. Shin and J. Hur. 2015. Tracking the evolution of stream DOM

source during storm events using end member mixing analysis based on DOM

quality. Journal of Hydrology 523:333-341.

Yang Y., G. Y., D. Xu, L. Wu, J. Zhou, S. Wang, H. Yu, Z. He, Y. Deng, Q. Lin, Y. Hu

and X. Li. 2014. The microbial gene diversity along an elevation gradient of the

tibetan grassland. ISME J.ISME Journal 8:430-440.

111

Yilmaz P., L.W. Parfrey, P. Yarza, J. Gerken, E. Pruesse, C.S. Quast T., J. Peplies, W.

Ludwig and F.O. Glockner. 2013. The SILVA and “all-species living tree project

(LTP)" taxonomic frameworks. Nucleic Acids Res. 42:D643-D648.

Zhou J., Y. Deng, Z. He, L. Wu and J.D. Van Nostrand. 2010. Applying GeoChip

analysis to disparate microbial communities. Microbe Magazine 5:60-65.

112

Appendix A: Sandusky Source Material DOM Analysis

113

Methods

Prior to ESI(-) FT-ICR MS analysis, the protocols used for organic matter

collection were tested to determine their ability to isolate organic phosphorus compounds.

Several organic phosphorus compounds were purchased to be used as reference

compounds: 2-aminoethyl phosphonate (2-AEP); fosfomycin (FOM); n-hexylphosphonic

acid (HexP); glucose-6-phosphate (G6P); phenyl phosphate (PhP); nicotinamide

dinucleotide phosphate (reduced, NADH); monopotassium phosphate (PO4); and sodium

pyrophosphate (P2O7). Each standard was prepared as a stock 1 mg L-1 P solution in DI

water. A sample of primary clarifier water was collected from the Southerly Wastewater

Plant (Columbus, OH) following the methods described in the manuscript. An initial

experiment was designed to determine the carbon, nitrogen, and phosphorus retention

efficiency of four SPE column types (Agilent Bond Elut): functionalized styrene

divinylbenzene (PPL); hydrophobic, bonded silica (C18); polymer anionic exchange

(PAX); strong anionic exchange (SAX). While the manufacturer instructions call for the

adjustment of samples to a pH 2 for the PPL and C18 columns, the PAX and SAX

columns recommend adjusting the sample to a pH 10. The primary clarifier water was

used to determine the retention of phosphorus by all four columns, both at pH 2 and pH

10 with duplicates for each column (n=16). Following the determination of pH

adjustments, a mixture of the reference organic phosphorus compounds was used to

determine the retention of these compounds for each filter at pH 10 using duplicate

columns. However, due to observed desorption, the SAX columns were excluded from

this subsequent analysis (n=6).

114

Primary clarifier water was passed through the SPE columns using the methods

described in the manuscript. The amount of carbon applied to each column type was

determined to meet the maximum sorptive capacity. The effluent from the columns was

collected into combusted glassware. The retention efficiency of NPOC and TDN were

determined by the change in concentration between the influent and effluent samples

measured with the Shimadzu TOC-V/TN. The retention efficiency of TDP was

determined by the change in concentration between the influent and effluent samples

using an Agilent ICP-OES.

A 7.5 mg L-1 P concentration mixture using equal parts (0.9375 mg L-1 P) of the

eight phosphorus reference compounds – six organic, and two inorganic – was prepared

for further analysis of the SPE columns. The mixture was analyzed using ion

chromatography with an AS-11HC column on a Dionex ICS-2100 ion chromatograph

(Dionex Corporation, Sunnyvale, CA). The flow rate was set at 1.5 mL/min for 15 min a

sample, eluted in a 1-60 mM gradient of KOH at 30°C. This method allowed for the

detection of seven out of the eight compounds, with the lone exception being 2-AEP.

These samples were made basic (pH 10) using KOH and gravity filtered through three

SPE column types in duplicate (n=6). The effluent was collected in combusted glassware

and ion chromatography analysis was used to visually detect the presence/absence of the

compounds. TOC/TDN/TDP concentrations were measured on the influent and effluent

samples to determine the retention efficiency.

115

Collection of Mass Spectrometry Data and Peak Detection

The samples were analyzed with electrospray ionization under the negative

ionization mode on a 7T FTICR mass spectrometer (Thermo Fisher Scientific, Waltham,

MA USA). The instrument settings were optimized by tuning on the SRFA standard. The

samples were infused into the ESI interface at 4 μL min-1, and the instrumental and spray

parameters were optimized for each sample. The capillary temperature was set at 250°C,

and the spray voltage was between 3.7 and 4 kV. For each sample, 200 scans were

collected spanning the 200-1000 m/z range. An external calibration mixture (Thermo

Calibration Mix; Thermo Fisher Scientific) was used to calibrate the mass accuracy to

<1.5 ppm. The processed spectra were internally calibrated resulting in a mass accuracy

of <1 ppm (Bhatia et al., 2010). The target average resolving power was 400,000 at m/z

400 (where resolving power is defined as m/Δm 50% where Δm is the width at half-

height of peak m).

Individual transients as well as a combined raw file were collected using xCalibur

2.0 (Thermo Fisher Scientific). Transients were co-added and processed with custom-

written MATLAB code (Southam et al., 2007). Only transients with a total ion current

>20% of the maximum value observed in each sample were added, processed with

Hanning apodization, and zero-filled prior to fast Fourier transformation. All m/z values

with a signal:noise ratio > 10 were retained. Spectra were internally re-calibrated using a

list of m/z values present in the majority of samples. Individual sample peak lists were

then aligned in MATLAB (Mantini et al., 2007). Formula assignments were made

through the custom-built Compound Identification Algorithm at the Wood Hole

116

Oceanographic Institution, as previously described (Kujawinski and Behn, 2006;

Kujawinski et al., 2009).

The nominal oxidation state of carbon (NOSC) for each identified formula was

calculated according to the equation of Koye et al. (2016). The equation is based on the

count of individual atom counts according to equation 1. The distribution of the NOSC

values were considered for each molecular classification, using only unique formula (no

duplicates between 12C and 13C isotopologues).

𝑁𝑂𝑆𝐶 = 4 −4𝐶+𝐻−2𝑂−3𝑁−2𝑆+5𝑃

𝐶 (equation 1)

Results & Discussion

The selection of SPE materials has been principally chosen so that the resulting

sample best reproduces the signature that would be observed in the original sample.

Previous research has used PPL filters for its broad selectivity of carbon (Ohno and

Ohno, 2013). However, phosphorus represents a minor portion of dissolved organic

matter pool. Selective concentration of organic phosphorus compounds enhances their

detectability in the organic matter spectrum (Cooper et al., 2005) Our objective was to

determine which SPE material and methodology would best suit our needs to retain

organic phosphorus compounds. The retention efficiencies of the all four SPE materials

had enhanced P recovery when samples were adjusted to a pH 10 (Figure A.1). Carbon

retention displayed some differences using this method with increased recovery for the

PAX column, but a reduction in the carbon recovery for the other three columns. Most

notably, the SAX column had an increased carbon concentration in the effluent, and

therefore was removed from subsequent analyses. As the majority of phosphorus may

117

have been inorganic in the primary clarifier water, it was important to demonstrate that

these columns were retaining organic phosphorus compounds.

Figure A.1. Retention of carbon, nitrogen, and phosphorus by solid phase extraction

columns. Wastewater primary clarifier water was used to assess the retention of dissolved

organic carbon, total dissolved nitrogen, and total dissolved phosphorus by the solid

phase extraction materials. Samples of the wastewater were adjusted to pH 2 or 10 using

hydrochloric acid or sodium hydroxide, respectively. The change in concentration was

multiplied by the volume which was passed through the filter to estimate the % retention

of these elements.

118

The primary clarifier water was likely to contain minerals that could interfere with

the interpretation of our results. For instance, the presence of magnesium in the water

combined with the pH adjustment could lead to the precipitation of inorganic phosphates

(Karl, David M.,Tien, Georgia, 1992). In fact, precipitates were visually observed in the

samples prior to filtration. Therefore, using the laboratory phosphorus standards allowed

us to detect their retention in the absence of interfering chemicals. Rather than

quantifying the change in concentrations, the ion chromatographs were used to identify

changes to the presence of the standard compounds before and after SPE filtration

(Figure A.2A). The PAX column nearly lacked four of the compounds in its effluent

chromatograph: HexP, PhP, NADH, and P2O7. These represented three organic and one

inorganic compound. Notably, there was a near complete recovery of nitrogen – as

determined by TDN analysis – that could indicate the recovery of the 2-AEP compound

(Figure A.2B). The determined recovery percent of nitrogen and phosphorus matched the

results expected presuming complete recovery of 2-AEP, PhP, NADH, and P2O7. These

results confirmed that the PAX column and methodology was adequate for organic

phosphorus retention, and therefore this solid phase extraction resin was selected for

future analyses.

119

Figure A.2. Recovery of known phosphorus standards. A standard solution consisting of

equal parts phosphorus of: (inorganic) orthophosphate, pyrophosphate,

(organophosphate) D-glucose-6-phosphate, phenylphosphate, NADH,

(organophosphonate) fosfomycin, 2-aminoethyl phosphonate, and n-hexylphosphonate

was prepared. The sample was basified to a pH 10 and passed through the Plexa-PAX,

PPL, and C18 columns. The standard solution was read using ion chromatography before

(influent) the eluent was collect from its respective column. The disappearance of a peak

has been interpreted as the adsorption of that compound to the SPE column. The 2-

aminoethyl phosphonate compound could not be detected using anionic IC. However, the

expected retention % assuming complete recovery of 1-aminoethyl phosphonate,

hexylphosphonate, phenylphosphate, NADH, and pyrophosphate by the Plexa-PAX filter

indicated that this compound also adhered to this filter (e.g., 100% recovery of nitrogen).

The DOM of our samples were composed of ≤12.8% DOP. Despite our efforts to

enhance organic phosphorus recovery by using the anionic exchange SPE column, the

non-manure samples were composed of less organic phosphorus than samples of Lake

Superior tributaries (Minor et al., 2012). It is noteworthy that we did not discern any

retention of organic phosphorus standards by the C18 column, which had been used in

the Lake Superior study (Minor et al., 2012). Rather than retaining a greater number of

phosphorus compounds, it is possible that our method simply enhanced the recovery

120

amounts rather than isolating new compounds. ESI FT-ICR-MS does not quantify the

concentrations of m/z values so there is no valid way of determining this for our sample

set (Kamga et al., 2014). Additionally, the formula algorithm also has an implicit bias

against organic phosphorus in that it preferentially selects formula with the lowest non-

oxygen (N+S+P) atom counts (Kujawinski and Behn, 2006; Kujawinski et al., 2009). For

every phosphorus atom incorporated in a formula, it becomes less likely for that formula

to be selected. Formula assignments are made within a 1ppm error window, meaning that

more options are available at higher molecular masses. Supporting this notion of an

assignment bias, the organic phosphorus compounds were more often assigned in the

lower molecular masses where there were fewer alternatives (data not shown). Our study

is a rare instance in which organic phosphorus was the intended focal point of ESI FT-

ICR-MS analysis. It would be useful to challenge the existing protocols if this technology

is to be applied for other studies centering around organic phosphorus.

121

Figure A.3. Carbon, nitrogen, and phosphorus concentration of samples in Sandusky

River watershed. (A) The carbon, nitrogen and phosphorus concentrations were measured

as non-purgeable carbon (NPOC), total dissolved nitrogen (TDN); and total dissolved

phosphorus (TDP). The detection limit (DL) for N was 0.01 mg N L-1, while it was 0.03

mg P L-1 leading to a lower limit of quantification (LOQ) of 0.1 mg P L-1. Concentrations

were diluted prior to solid phase extraction.

122

Figure A.4. The distribution of NOSC values by molecular classes.

Table A.1. Adsorption efficiency across samples using the Bond Elut PAX solid phase

extraction resin. Carbon was measured using non-purgeable organic carbon, while

nitrogen and phosphorus were measured as the change in concentrations following

sample dilution and after passing through the solid phase extraction columns. Values

below the limit of quantification (10 µg N L-1, 100 µg P L-1) are reported as estimates.

Where effluent values were above influent, values are reported as <0%.

Sample Replicate C N P

Chicken replicate 1 18% 26% <0%

replicate 2 19% 28% 15%

Dairy replicate 1 20% 13% 6.4%

replicate 2 8% 31% 5.2%

Hog replicate 1 10% 41% <0%

replicate 2 19% 41% <0%

WWTP Effluent replicate 1 44% 32% est. 100%

replicate 2 12% 32% est. 97%

Edge of Field replicate 1 21% 6% est. 17%

replicate 2 16% 7% est. 9.1%

Sandusky River replicate 1 36% 25% est. 59%

replicate 2 28% 33% est. 76%

SRFA - 41% BDL BDL

PLFA - 42% BDL BDL

123

Table A.2. ESI(-) FT-ICR-MS analysis detected a total of 14637 peaks, spread across the samples and replicates. The data was quality

filtered by removing peaks detected in the DI procedural blank; the extraction solvent; singletons (detected in only 1 sample of the

entire dataset); and peaks which had no assigned formula. The reproducibility was determined between sample replicates

(shared#/mean#).

Hog Chicken Dairy Wastewater Edge of field

Sandusky

River NOM

Processing Total DI Solvent 1 2 1 2 1 2 1 2 1 2 1 2 PLFA SRFA

All

Detected

Peaks

14637 3014 534 2497 2352 3452 2493 3519 1626 2388 3602 4449 4702 4169 3154 3412 3707

Remove

Peaks in

Blank

11633 - 377 2094 2053 3219 2312 3215 1232 2245 3341 4338 4483 3983 3023 3262 3415

Remove

Peaks in

Solvent

11246 - - 1983 1931 3098 2200 3070 1128 2171 3270 4260 4396 3895 2939 3146 3325

Removed

Singleton

Peaks

7438 - - 1673 1700 2364 2072 2476 995 2070 3096 3979 4254 3853 2815 2630 3027

Assigned

Formula 7250 - - 1590 1625 2315 2021 2444 964 2046 3071 3974 4220 3846 2798 2622 3021

Reproducibility between sample replicates 88% 88% 68% 81% 90% 85% -

124

Table A.3. ESI(-) FT-ICR-MS analysis provided peaks which were assigned formulas with C/H/O/N/P/S elements. The distribution of

the m/z values detected in each sample were distributed across 8 formula classes. The numbers of formula are printed for each sample

replicate with the number in bold indicating the total number detected in the combined samples. I think the data in this figure are fine,

but it is a little hard to separate the different formula classes without some additional lines or perhaps presenting the data in a figure.

Hog Chicken Dairy WWTP Effluent Edge of field Sandusky River NOM

Total 1 2 1 2 1 2 1 2 1 2 1 2 PLFA SRFA

CHO 3981 772 857 824 801 1139 475 1700 2374 3037 3190 2913 2154

1749 2727 906 908 1144 2413 3356 2924

CHON 2198 502 466 1047 814 923 276 239 479 732 827 769 549

751 119 550 1064 927 488 903 811

CHOP 394 111 99 172 150 203 112 57 126 117 117 109 75

70 78 119 179 207 129 132 111

CHOS 254 62 68 75 69 89 37 7 40 68 35 21 2

38 74 72 82 93 40 75 22

CHONP 147 50 44 81 79 30 23 9 19 7 18 12 6

3 5 53 83 35 19 18 13

CHONS 149 39 46 66 61 37 20 22 19 10 19 15 9

8 12 47 69 40 28 22 16

CHOPS 62 30 24 24 23 18 14 2 5 0 3 2 0

1 2 30 25 18 5 3 2

CHONPS 65 24 21 26 24 5 7 10 9 3 11 5 3

2 4 26 26 9 14 13 6

Total 7250 1590 1625 2315 2021 2444 964 2046 3071 3974 4220 3846 2798

2622 3021 1803 2436 2473 3136 4522 3905

125

Hog Chicken Dairy WWTP Effluent Edge of Field Sandusky River DOM DOP

0 0 0 0 0 1 34 1

0 0 0 0 1 0 317 9

0 0 0 0 1 1 987 12

0 0 0 1 0 0 112 27

0 0 0 1 0 1 26 2

0 0 0 1 1 0 259 26

0 0 0 1 1 1 1506 29

0 0 1 0 0 0 337 54

0 0 1 0 0 1 4 2

0 0 1 0 1 0 17 0

0 0 1 0 1 1 17 1

0 0 1 1 0 0 15 3

0 0 1 1 0 1 1 0

0 0 1 1 1 0 26 1

0 0 1 1 1 1 232 12

0 1 0 0 0 0 715 159

0 1 0 0 0 1 4 0

0 1 0 0 1 0 9 0

0 1 0 0 1 1 59 4

0 1 0 1 0 0 1 1

0 1 0 1 0 1 1 0

0 1 0 1 1 0 6 1

0 1 0 1 1 1 99 6

0 1 1 0 0 0 332 49

0 1 1 0 0 1 4 1

0 1 1 0 1 0 18 0

0 1 1 0 1 1 61 7

0 1 1 1 0 0 5 4

0 1 1 1 0 1 1 0

0 1 1 1 1 0 5 0

0 1 1 1 1 1 227 29

Continued

Table A.4. The Venn counts of Sandusky source material data. The samples columns are

binary (0 not included; 1 included) with the numbers in the DOM and DOP columns

indicating the number of formula for that group of samples.

126

Table A.4 Continued

Hog Chicken Dairy WWTP Effluent Edge of Field Sandusky River DOM

1 0 0 0 0 0 445 100

1 0 0 0 0 1 3 1

1 0 0 0 1 0 8 1

1 0 0 0 1 1 3 0

1 0 0 1 0 0 5 1

1 0 0 1 0 1 0 0

1 0 0 1 1 0 9 3

1 0 0 1 1 1 35 8

1 0 1 0 0 0 326 57

1 0 1 0 0 1 1 0

1 0 1 0 1 0 8 0

1 0 1 0 1 1 1 0

1 0 1 1 0 0 7 1

1 0 1 1 0 1 0 0

1 0 1 1 1 0 11 0

1 0 1 1 1 1 52 4

1 1 0 0 0 0 69 7

1 1 0 0 0 1 1 0

1 1 0 0 1 0 1 1

1 1 0 0 1 1 6 0

1 1 0 1 0 0 1 0

1 1 0 1 0 1 1 0

1 1 0 1 1 0 0 0

1 1 0 1 1 1 45 0

1 1 1 0 0 0 237 30

1 1 1 0 0 1 17 1

1 1 1 0 1 0 14 0

1 1 1 0 1 1 49 4

1 1 1 1 0 0 12 1

1 1 1 1 0 1 1 0

1 1 1 1 1 0 8 0

1 1 1 1 1 1 427 8

127

m/z Formula C13 Hog Chicken Dairy WWTP Effluent Edge of field Sandusky River

432.0675772 C20H20O6NPS - - - 2.14E-04 - - 6.62E-05

464.1477252 C22H28O8NP - - - 3.78E-04 - - 8.25E-05

277.1433072 C10H23O3N4P - - 1.10E-04 4.24E-04 - - 3.39E-05

276.0724817 C11H17O6P 1 2.54E-04 6.89E-04 8.44E-04 - - 4.09E-05

408.2239451 C19H37O7P 1 - - - 7.22E-05 - 5.76E-05

376.0885573 C15H21O9P 1 - - - - 1.13E-04 6.34E-05

420.0783553 C16H21O11P 1 - - - - 9.02E-05 4.77E-05

420.1147941 C17H25O10P 1 - - - - 8.38E-05 5.16E-05

430.0779524 C21H19O8P 1 - - - - 9.68E-05 5.18E-05

434.0940188 C17H23O11P 1 - - - - 8.76E-05 5.08E-05

502.1565882 C22H31O11P 1 - - - - 8.25E-05 5.05E-05

406.135515 C17H27O9P 1 - - 7.44E-05 - 8.54E-05 4.73E-05

332.0623426 C13H17O8P 1 - 2.05E-04 - - 9.30E-05 4.94E-05

392.0834377 C15H21O10P 1 - 2.02E-04 - - 7.88E-05 4.62E-05

302.0881438 C13H19O6P 1 - 2.59E-04 4.41E-04 - 8.02E-05 4.26E-05

304.0674032 C12H17O7P 1 - 5.06E-04 4.50E-04 - 7.36E-05 4.10E-05

318.0830637 C13H19O7P 1 - 5.92E-04 1.91E-04 - 8.09E-05 4.78E-05

330.0830667 C14H19O7P 1 - 2.01E-04 1.47E-04 - 1.36E-04 1.38E-04

Continued

Table A.5. List of potential marker formulas found in source and Sandusky River samples. The mass to charge (m/z) ratios were used

to identify a molecular formula. C13 indicates the presence (1) or absence (0) of a single 13C isotope in the formula. The relative peak

height for the m/z values in the samples is provided, and - signifies that the m/z value was not detected for that sample.

128

Table A.5 Continued

m/z Formula C13 Hog Chicken Dairy WWTP Effluent Edge of field Sandusky River

332.0987091 C14H21O7P 1 - 2.57E-04 1.13E-04 - 9.82E-05 4.42E-05

362.1092918 C15H23O8P 1 - 6.35E-05 7.52E-05 - 9.24E-05 4.80E-05

275.0260021 C9H13O4N2PS - 4.46E-04 1.31E-04 7.34E-04 - 4.33E-05 8.45E-05

294.0619326 C14H15O5P 1 3.39E-04 4.22E-04 6.09E-04 - 9.19E-05 1.27E-04

320.0775812 C16H17O5P 1 3.64E-04 1.78E-04 4.03E-04 - 4.06E-05 4.53E-05

421.236256 C20H39O7P - 3.57E-04 1.97E-04 1.33E-03 - 3.64E-05 7.27E-05

372.1664215 C18H29O6P 1 - - - 3.03E-04 1.33E-04 1.76E-04

386.1820806 C19H31O6P 1 - - - 2.22E-04 9.86E-05 1.10E-04

388.1613368 C18H29O7P 1 - - - 2.70E-04 1.54E-04 1.69E-04

402.1769532 C19H31O7P 1 - - - 2.09E-04 1.05E-04 1.38E-04

403.1165379 C17H25O9P - - - - 1.00E-04 9.28E-05 4.57E-05

413.100923 C18H23O9P - - - - 1.10E-04 8.84E-05 3.93E-05

416.1926161 C20H33O7P 1 - - - 1.74E-04 1.04E-04 1.08E-04

417.095797 C17H23O10P - - - - 9.12E-05 4.37E-05 4.11E-05

418.0990897 C17H23O10P 1 - - - 7.87E-05 1.75E-04 1.74E-04

418.1718666 C19H31O8P 1 - - - 7.17E-05 9.78E-05 5.75E-05

429.1322172 C19H27O9P - - - - 1.18E-04 7.46E-05 3.66E-05

488.1773569 C22H33O10P 1 - - - 5.45E-05 8.69E-05 4.83E-05

344.1351164 C16H25O6P 1 - - 8.28E-05 2.60E-04 1.41E-04 1.60E-04

370.1507808 C18H27O6P 1 - - 9.10E-05 3.58E-04 1.57E-04 2.28E-04

372.1300277 C17H25O7P 1 - - 9.49E-05 3.00E-04 2.15E-04 2.31E-04

Continued

129

Table A.5 Continued

m/z Formula C13 Hog Chicken Dairy WWTP Effluent Edge of field Sandusky River

382.1507627 C19H27O6P 1 - - 1.11E-04 3.36E-04 2.01E-04 2.34E-04

384.166397 C19H29O6P 1 - - 9.51E-05 2.90E-04 1.77E-04 2.08E-04

399.0852169 C17H21O9P - - - 4.55E-04 7.87E-05 8.40E-05 3.89E-05

414.176966 C20H31O7P 1 - - 7.62E-05 3.70E-04 2.13E-04 2.34E-04

458.1668009 C21H31O9P 1 - - 7.86E-05 2.57E-04 1.63E-04 1.92E-04

484.1824486 C23H33O9P 1 - - 7.72E-05 2.19E-04 1.65E-04 2.03E-04

360.130038 C16H25O7P 1 - 7.45E-05 - 2.13E-04 1.59E-04 1.88E-04

456.1147265 C20H25O10P 1 - 7.61E-05 - 2.90E-04 2.52E-04 2.77E-04

342.1194474 C16H23O6P 1 - 1.67E-04 1.17E-04 3.47E-04 1.65E-04 2.11E-04

344.0987134 C15H21O7P 1 - 1.87E-04 1.48E-04 8.20E-05 1.40E-04 1.73E-04

350.0881719 C17H19O6P 1 - 2.26E-04 4.02E-04 2.42E-04 1.37E-04 1.56E-04

354.0830662 C16H19O7P 1 - 3.05E-04 1.23E-04 2.32E-04 2.25E-04 2.50E-04

356.0623315 C15H17O8P 1 - 3.92E-04 1.82E-04 6.87E-05 2.34E-04 2.79E-04

356.1351082 C17H25O6P 1 - 6.45E-05 1.11E-04 2.68E-04 1.47E-04 1.73E-04

358.1143977 C16H23O7P 1 - 6.54E-05 1.13E-04 2.56E-04 2.25E-04 2.31E-04

368.0987153 C17H21O7P 1 - 2.04E-04 1.38E-04 3.00E-04 2.26E-04 2.86E-04

370.1143768 C17H23O7P 1 - 1.06E-04 1.22E-04 2.99E-04 2.25E-04 2.70E-04

380.098741 C18H21O7P 1 - 1.02E-04 3.92E-04 3.54E-04 2.54E-04 3.08E-04

382.1143594 C18H23O7P 1 - 2.00E-04 1.27E-04 3.69E-04 2.99E-04 3.32E-04

396.1300303 C19H25O7P 1 - 2.24E-04 1.46E-04 4.39E-04 3.31E-04 3.48E-04

398.1456726 C19H27O7P 1 - 2.14E-04 1.23E-04 4.31E-04 3.05E-04 3.31E-04

Continued

130

Table A.5 Continued

m/z Formula C13 Hog Chicken Dairy WWTP Effluent Edge of field Sandusky River

410.109275 C19H23O8P 1 - 1.94E-04 1.33E-04 4.50E-04 3.60E-04 4.13E-04

410.145669 C20H27O7P 1 - 2.18E-04 1.17E-04 4.47E-04 2.96E-04 3.49E-04

412.1613237 C20H29O7P 1 - 1.75E-04 1.18E-04 4.35E-04 2.96E-04 3.38E-04

430.1354991 C19H27O9P 1 - 6.31E-05 1.10E-04 2.77E-04 2.47E-04 2.50E-04

440.119842 C20H25O9P 1 - 1.73E-04 9.72E-05 3.63E-04 3.13E-04 3.41E-04

440.1562514 C21H29O8P 1 - 1.85E-04 1.13E-04 3.42E-04 2.81E-04 3.06E-04

426.067739 C18H19O10P 1 3.78E-04 - 1.20E-03 2.03E-04 2.97E-04 3.55E-04

296.0775853 C14H17O5P 1 2.26E-04 5.66E-04 4.98E-04 5.81E-05 1.02E-04 6.58E-05

322.0932496 C16H19O5P 1 5.06E-04 2.23E-04 8.40E-04 7.76E-05 1.00E-04 1.61E-04

336.1088903 C17H21O5P 1 1.91E-04 2.01E-04 1.31E-04 2.15E-04 9.13E-05 1.17E-04

366.0830953 C17H19O7P 1 6.32E-04 2.19E-04 1.42E-03 2.65E-04 1.95E-04 2.55E-04

384.0936035 C17H21O8P 1 2.21E-04 3.47E-04 8.02E-04 2.89E-04 3.47E-04 3.64E-04

131

Figure A.5. Spectra captured from ESI(-) FT-ICR-MS analysis of all sample replicates and blanks.

132

Appendix B: Portage River DOM Mixing Analysis

133

Table B.1. StreamStats data obtained from the four confluence sampling locations. The

contribution of the two tributaries were estimated from their relative 2-year recurrenc

interval flows. The USGS guage station at Woodville was monitored to estimate the

flows during the time of sampling.

Confluence Location

Drainage

Area

(mi2) Latitude Longitude

Slope

(%)

Forest

(%)

2-year

Return

Interval

(cfs)

Contribution

(%)

E

1. Bays Rd

South

Branch

53.9 41.26807 -83.52579 6.7 4.47 1280 61%

2. Bays Rd

East Branch 35.3 41.26807 -83.50652 5.98 5.09 826 39%

3. Rt 281 91.7 41.2826 -83.50993 6.23 4.87 1840

F

1. Bridge St 348 41.40916 -83.45708 4.16 4.27 4940 82%

2. Water St 57.3 41.41011 -83.45848 1.8 4 1120 18%

3. Bierly

Ave 406 41.41254 -83.45453 4.13 4.23 5580

G

1. Toledo St 426 41.4772 -83.29551 3.57 4.44 5620 82%

2. Hessville

Rd 63 41.48674 -83.24105 4.25 8.78 1250 18%

3. Rt 590 493 41.49145 -83.2215 3.74 5 6300

H

1. Chet's

Place 536 41.50562 -83.06928 3.39 5.06 6510 91%

2. Little

Portage 29.9 41.48656 -83.05353 4.82 7.55 649 9%

3. Portage

River

Retreat

579 41.51377 -83.00750 2.84 5.22 6410

Woodville

(USGS

gauge)

421 41.44935 -83.35808 3.79 4.33 5620

134

Table B.2. Nutrient concentrations and solid phase extraction (SPE) efficiencies of the Portage River samples. Phosphorus was

measured using colorimetric methods and via ICP-OES for estimating SPE efficiencies.

Colorimetric

NPOC (mg L-1) TDN (mg L-1) TDP (mg L-1) NPOCSPE (%) TDNSPE (%) TDPSPE (est. %) Sample ID DRP DHP TDP

A. W Township Rd 14 0.099 0.117 0.125 13.73 7.179 0.16 42% 4% 0%

B. Independence Ave 0.102 0.114 0.122 10.31 7.309 0.16 42% 6% 6%

Independence Ave (rep) - - - 10.53 7.417 0.16 27% 6% 13%

C. Tiffin St 0.098 0.109 0.117 12.07 7.165 0.17 37% 4% 6%

D. Fostoria WWTP 0.100 0.112 0.121 46.67 7.309 0.10 34% 5% <0%

E.1. Bays Rd South 0.105 0.119 0.132 8.106 6.078 est. 0.09 51% 1% <0%

E.2. Bays Rd East 0.107 0.116 0.124 12.57 10.24 0.24 40% 10% 8%

E.3. Rt 281 - - - 9.413 8.048 0.17 40% 9% 6%

F.1. Bridge St 0.087 0.098 0.107 8.929 9.876 0.10 34% 3% 20%

Bridge St (rep) - - - 9.807 9.958 0.11 25% 9% 45%

F.2. Water St 0.087 0.096 0.105 8.061 8.773 <0.03 45% 5% BDL

F.3. Bierly Ave 0.086 0.097 0.107 8.808 9.756 est. 0.06 26% 4% 33%

G.1. Toledo St 0.085 0.094 0.105 8.333 9.231 est. 0.06 33% 2% 0%

G.2. Hessville Rd 0.092 0.104 0.113 8.432 7.773 est. 0.03 30% 6% 67%

G.3. Rt 590 0.082 0.094 0.101 8.505 9.05 0.05 41% 7% 20%

H.1. Chet's Place 0.059 0.066 0.073 7.339 5.876 <0.03 24% 9% BDL

H.2. Little Portage 0.064 0.073 0.083 8.441 5.885 est. 0.04 37% 10% 0%

H.3. Portage River Retreat 0.027 0.033 0.038 5.985 8.013 <0.03 41% 7% BDL

SRFA-B (Method) - - - 7.569 0.119 <0.03 32% 15% BDL

DI-B (Blank) - - - BDL BDL <0.03 BDL BDL BDL

135

Table B.3. QA/QC filtering of the data and the number of m/z values remaining in samples at each step.

DOM DOP

Sample Data -DI Blank -Solvents -Singletons -No Formula Data -Blank (MilliQ) -Solvents -Singletons

Data set 29273 22740 19577 11344 11064 2440 1619 1041 501

DI Blank (MilliQ) 6533 0 0 0 0 821 0 0 0

Solvent (Facility) 871 542 0 0 0 112 71 0 0

Solvent (User) 4376 2633 0 0 0 859 507 0 0

A. W Township Rd 14 5885 5484 5410 5376 5361 145 130 129 125

B. Independence Ave 6049 5641 5558 5525 5508 154 131 128 126

C. Tiffin St 5789 5358 5286 5212 5194 146 125 123 118

D. Fostoria WWTP 8301 7802 7672 7184 7148 241 212 207 178

E.1. Bays Rd South Branch 8186 7691 7574 6308 6239 280 246 242 183

E.2. Bays Rd East Branch 5738 5280 5203 5153 5116 168 138 133 129

E.3. Rt 281 8033 7515 7380 7001 6946 243 204 200 176

F.1. Bridge St 6017 5565 5492 5423 5370 176 145 140 137

F.2. Water St 6027 5579 5473 5385 5353 245 215 207 196

F.3. Bierly Ave 6014 5534 5432 5404 5364 223 179 170 169

G.1. Toledo St 8339 7789 7646 7177 7129 305 259 245 208

G.2. Hessville Rd 6151 5703 5599 5551 5525 176 143 135 131

G.3. Rt 590 6124 5687 5598 5552 5507 186 157 151 144

H.1. Chet's Place 6024 5604 5518 5488 5461 176 146 143 140

H.2. Little Portage 8267 7755 7618 7281 7207 329 298 290 255

H.3 Portage River Retreat 9280 8908 8797 7188 7103 385 356 352 257

SRFA 6291 5990 5911 4627 4606 161 153 151 102

SRFA (Method) 5852 5518 5452 4889 4858 211 204 202 154

136

Figure B.1. Spectra collected by ESI(-) FT-ICR-MS analysis. All samples and replicates (rep) are shown. Replicates were used as a

check, and then removed from the analysis. Solvent blanks were run from the analytical facility as well as from the user. The

methodological SRFA sample was run through the solid phase extraction process while the SRFA sample was prepared directly in the

user solvent.

137

Table B.4. The distribution of atomic composition of formula identified in each Portage

River sample.

Sample ID CHO CHON CHONP CHONPS CHONS CHOP CHOPS CHOS

A. W Township Rd 14 70.66% 25.57% 0.24% 0.43% 0.88% 1.66% 0.00% 0.56%

B. Independence Ave 70.41% 26.00% 0.33% 0.33% 0.82% 1.63% 0.00% 0.49%

C. Tiffin St 73.47% 22.62% 0.33% 0.39% 1.14% 1.50% 0.06% 0.50%

D. Fostoria WWTP 62.66% 32.88% 0.53% 0.41% 1.22% 1.51% 0.04% 0.76%

E.2. Bays Rd East

Branch 72.46% 23.79% 0.53% 0.27% 0.80% 1.68% 0.04% 0.43%

E.1. Bays Rd South

Branch 60.35% 34.46% 0.79% 0.30% 1.17% 1.75% 0.10% 1.09%

E.3. Rt 281 65.52% 30.16% 0.62% 0.40% 1.02% 1.43% 0.09% 0.76%

F.1. Bridge St 65.96% 28.04% 0.49% 0.47% 1.05% 2.67% 0.04% 1.29%

F.2. Water St 71.34% 24.34% 0.32% 0.52% 1.23% 1.60% 0.11% 0.54%

F.3. Bierly Ave 66.52% 28.06% 0.39% 0.54% 1.21% 2.14% 0.07% 1.06%

G.1. Toledo St 63.22% 31.80% 0.69% 0.48% 1.29% 1.68% 0.07% 0.77%

G.2. Hessville Rd 71.19% 25.29% 0.29% 0.34% 0.69% 1.65% 0.09% 0.47%

G.3. Rt 590 70.31% 25.71% 0.42% 0.31% 0.94% 1.73% 0.16% 0.42%

H.1. Chet's Place 61.80% 31.94% 0.92% 0.51% 1.35% 2.08% 0.03% 1.37%

H.2. Little Portage 70.85% 25.07% 0.31% 0.48% 0.81% 1.74% 0.04% 0.71%

H.3. Portage River

Retreat 57.43% 36.21% 0.96% 0.61% 1.34% 2.00% 0.06% 1.41%

138

Table B.5. The molecular class distribution of formula identified in the Portage River

samples.

Sample Carbohydrate

Condensed

hydrocarbon Lignin Lipid Other Protein Tannin

Unsaturated

hydrocarbon

A. W

Township Rd

14

0.8% 0.6% 80.3% 2.5% 1.7% 7.1% 6.1% 0.9%

B.

Independence

Ave

0.8% 0.5% 79.4% 2.9% 1.8% 7.8% 6.0% 0.8%

C. Tiffin St 0.5% 0.5% 79.3% 3.7% 1.6% 8.3% 5.1% 1.0%

D. Fostoria

WWTP 1.5% 1.2% 75.6% 2.8% 2.5% 7.9% 7.4% 1.1%

E.1. Bays Rd

South Branch 2.2% 0.9% 73.0% 3.1% 3.3% 8.8% 7.3% 1.3%

E.2. Bays Rd

East Branch 0.7% 0.5% 80.9% 3.7% 1.1% 8.9% 3.4% 0.9%

E.3. Rt 281 1.6% 0.9% 75.9% 3.5% 2.4% 8.4% 6.2% 1.1%

F.1. Bridge St 1.0% 0.3% 78.0% 2.4% 2.7% 8.4% 6.3% 0.9%

F.2. Water St 0.9% 0.3% 80.8% 3.8% 1.4% 8.1% 3.4% 1.3%

F.3. Bierly

Ave 1.0% 0.3% 77.4% 2.5% 2.7% 8.0% 7.2% 0.9%

G.1. Toledo

St 1.9% 0.7% 77.8% 3.7% 1.9% 8.2% 4.4% 1.3%

G.2. Hessville

Rd 0.6% 0.5% 81.5% 3.1% 1.3% 7.6% 4.7% 0.8%

G.3. Rt 590 0.8% 0.3% 81.0% 2.8% 1.4% 7.7% 5.0% 1.0%

H.1. Chet's

Place 1.8% 1.0% 73.9% 2.6% 3.6% 8.0% 7.8% 1.3%

H.2. Little

Portage 0.9% 0.5% 79.4% 3.0% 1.6% 7.9% 5.7% 0.9%

H.3. Portage

River Retreat 2.3% 0.9% 73.5% 2.5% 4.0% 7.8% 7.6% 1.4%

139

Figure B.2. Correlations between nitrogen and phosphorus concentrations and elemental

compositions. (A) The correlation between total dissolved nitrogen and CHON* formula

was found to not be significant (Pearson correlation, line). (B) The correlation between

total dissolved phosphorus and TDP was significant (Pearson correlation, line).

140

Figure B.3. Hierarchal clustered dendrogram and heatmap based off the Canberra

distance matrix.

141

Figure B.4. Hierarchal clustering of the binary Jaccard distance matrix between samples

collected in both Chapter 1 and Chapter 2. Only m/z values detected in both sets were

considered. The Chapter 1 samples are identified with red bars.

142

Figure B.5. The relative change in peak heights between upstream-downstream samples

in the upper reaches of the Portage River (A through E.2). The distributions were

compiled for all pairs and examined based upon molecular class types assigned based on

their position in the van Krevelen diagrams.

143

Appendix C: Antibiotic Resistance Gene Analysis

144

Table C.1. Yields and purity of DNA extracts of the sediments collected in 2016. Values

were used to calculate gene abundance in sediment g-1 dry weight.

Sample Replicate

DNA concentration

(ng/µl) 260:280 260:230

Sediment

(g dry weight)

CHLP

1 19.3 1.88 1.08 0.60

2 19.7 1.91 1.25 0.33

3 18.8 1.67 1.19 0.41

OO

1 20.6 1.86 1.69 0.30

2 18.9 1.9 1.77 0.27

3 20.2 1.88 1.67 0.33

PP

1 9.5 1.87 1.05 0.35

2 11.4 1.79 1.82 0.35

3 13.7 1.82 1.39 0.39

Table C.2. Methodology used in LC separation of antibiotics. Gradient elution of 0.1%

formic acid in methanol (% B) with respect to time (min) on Waters XSelect CSH C18

column that separated sulfonamides, macrolides, and others via method 1 and

fluoroquinolones and tetracyclines via method 2.

Method 1 Method 2

Time (min) % B Time (min) % B

0.0 0 0.0 0

5.5 100 0.5 0

7.5 100 4.0 40

8.0 0 7.0 100

20.0 0 9.0 100

-- -- 10.0 0

-- -- 20.0 0

145

Analyte Parent Ion Product Ion CE Quantification or

(m/z) (m/z) (V) Confirmation

Sulfonamides

sulfapyridine 250.10 156.00 17 quantification

250.10 108.05 25 confirmation

sulfadiazine 251.05 156.00 15 quantification

251.05 108.05 24 confirmation

sulfamethoxazole 254.05 92.10 29 quantification

254.05 108.00 24 confirmation

sulfamethazine 279.05 186.00 17 quantification

279.05 156.00 20 confirmation

sulfachloropyridazine 285.00 156.06 15 quantification

285.00 92.05 35 confirmation

sulfadimethoxine 311.10 156.06 21 quantification

311.10 92.05 35 confirmation

13C6-sulfamethoxazole 260.05 98.10 32 quantification

(internal standard) 260.05 114.10 27 confirmation

13C6-sulfamethazine 285.05 186.00 22 quantification

(surrogate) 285.05 123.00 20 confirmation

Fluoroquinolones

norfloxacin 320.10 276.10 17 quantification

320.10 302.10 21 confirmation

ciprofloxacin 332.10 231.05 35 quantification

332.10 314.10 21 confirmation

enrofloxacin 360.10 245.10 25 quantification

360.10 316.15 19 confirmation

ofloxacin 362.10 261.10 28 quantification

362.10 318.10 19 confirmation

clinafloxacin 366.10 348.00 20 confirmation

(internal standard) 366.10 305.00 22 quantification

nalidixic acid 233.15 187.00 27 confirmation

(surrogate) 233.15 104.05 40 quantification

Continued

Table C. 3. Methodology used in LC separation of antibiotics. Gradient elution of 0.1%

formic acid in methanol (% B) with respect to time (min) on Waters XSelect CSH C18

column that separated sulfonamides, macrolides, and others via method 1 and

fluoroquinolones and tetracyclines via method 2.

146

Table C.3 Continued

Analyte Parent Ion Product Ion CE Quantification or

(m/z) (m/z) (V) Confirmation

Tetracyclines Tetracycline 445.10 410.10 19 quantification

445.10 427.05 11 confirmation

doxycycline 445.10 321.05 31 quantification

445.10 428.15 18 confirmation

oxytetracycline 461.10 426.10 17 quantification

461.10 443.10 12 Confirmation

chlortetracycline 479.05 462.10 20 quantification

& degradation products 479.05 444.10 17 Confirmation

481.05 464.10 20 quantification

481.05 446.10 30 Confirmation

demeclocycline 465.10 448.05 20 quantification

(surrogate) 465.10 430.05 17 Confirmation

Macrolides erythromycin 734.4 158.15 35 quantification

734.4 576.35 15 Confirmation

erythromycin-H2O 716.45 158.15 35 quantification

716.45 558.35 15 Confirmation

roxithromycin 837.45 158.10 35 quantification

837.45 679.45 20 confirmation

Tylosin 916.45 174.10 40 quantification

916.45 772.45 30 confirmation 13C2-erythromycin 736.40 160.15 35 quantification

736.40 578.35 20 confirmation

13C2-erythromycin-H2O

718.40 160.15 35 quantification

718.40 560.35 20 confirmation

Non-categorized Carbadox 263.10 130.05 22 quantification

263.10 231.05 13 confirmation

trimethoprim 291.10 230.10 23 quantification

291.10 123.05 24 confirmation

Lincomycin 407.30 126.10 35 quantification

407.30 359.20 18 confirmation

Simeton 198.20 68.10 33 quantification

(internal standard) 198.20 100.10 27 confirmation

147

Genbank ID Gene Gene category

2015 2016

CHLP OO PP CHLP OO PP

302407215 pel Cdeg Carbon Cycling 168.6 91.4 122.8 24.8 31.3 20.6

284165967 phytoene synthase Secondary metabolism 141.9 88.1 121.5 26.2 32.2 19.9

327480404 MFS antibiotic Antibiotic resistance 128.8 88.5 108.4 25.6 32.3 21.3

89069932 ompR Stress 129.7 78.0 104.1 18.5 25.3 18.6

196259193 fnr Stress 120.3 79.3 104.6 24.8 32.4 23.0

91802036 ompR Stress 127.4 75.6 95.6 22.9 29.9 21.3

50954857 gdh Nitrogen 132.2 67.1 99.0 18.7 29.1 16.3

329118461 nikA Metal Homeostasis 126.4 73.5 98.2 21.2 31.1 18.2

390569537 MFS antibiotic Antibiotic resistance 104.6 77.5 113.9 24.6 36.7 23.3

153962580 sqr Sulfur 123.0 74.8 97.0 20.2 29.8 17.5

315603069 nikA Metal Homeostasis 109.3 81.0 103.2 25.4 30.3 21.6

325265934 nitroreductase b Organic Remediation 128.8 67.0 97.4 18.2 26.3 15.8

148273451 nhaA Metal Homeostasis 106.6 80.1 105.1 19.7 26.4 19.3

335036555 znuC Metal Homeostasis 117.0 72.9 101.2 0.0 0.0 0.0

120605943 merT Metal Homeostasis 105.6 80.1 104.3 0.0 0.0 0.0

127514333 MFS antibiotic Antibiotic resistance 116.0 76.8 96.8 23.9 31.2 21.8

170738862 arsB Metal Homeostasis 118.1 71.8 97.8 19.6 28.7 16.7

345010049 corA Metal Homeostasis 100.1 94.4 93.0 30.2 36.9 24.9

178465870 MFS antibiotic Antibiotic resistance 118.9 71.3 96.9 18.1 26.7 17.2

323276022 MFS antibiotic Antibiotic resistance 116.4 67.6 101.0 18.1 28.0 16.6

334103082 nikA Metal Homeostasis 115.2 71.6 97.9 21.7 31.4 22.0

283786071 CsoS1 CcmK Carbon Cycling 114.3 71.5 96.8 17.2 22.8 16.7

345013283 MFS antibiotic Antibiotic resistance 117.9 67.2 95.6 18.5 26.4 16.9

311112862 MFS antibiotic Antibiotic resistance 94.0 79.0 103.5 21.8 33.0 21.5

83838314 pstB Stress 108.6 72.7 93.0 18.0 25.3 16.4

171472824 sod FeMn secondary metabolism 110.3 68.3 94.8 19.6 28.0 17.4

254388596 pstA Stress 95.5 80.6 96.7 26.2 26.1 19.8

42627732 catechol Organic Remediation 109.4 69.0 94.4 19.1 28.7 16.7

238059362 arsB Metal Homeostasis 92.0 79.0 100.9 21.8 23.2 19.7

238757841 spiC virulence 96.8 70.9 98.6 19.4 30.7 16.8

Continued

Table C.4. Gene probe normalized signals for 99.9th percentile of detected values in the

GeoChip analysis on the sediments collected in 2016.

148

Table C.4 Continued

Genbank

ID Gene Gene category

2015 2016

CHLP OO PP CHLP OO PP

83950262 ompR Stress 95.4 73.7 97.1 17.6 23.3 17.3

292675888 Mex

Antibiotic

resistance 96.8 68.5 99.7 18.4 27.4 15.7

339611386 MATE antibiotic

Antibiotic

resistance 91.6 85.8 87.4 26.2 32.6 22.5

221723831 catechol b

Organic

Remediation 92.9 74.5 97.2 27.3 34.2 21.6

91978835 ompR Stress 95.5 71.7 93.9 22.2 29.9 20.4

367038109 Ara Carbon Cycling 93.8 72.1 92.9 0.0 0.0 0.0

285019189 kup Metal Homeostasis 91.8 81.5 85.4 0.0 0.0 0.0

167561665 mdla

Organic

Remediation 89.7 77.4 91.1 24.0 31.6 20.7

297193687 phytase Phosphorus 84.6 77.6 94.7 19.6 28.9 17.3

299134974 TIM Carbon Cycling 82.4 84.9 85.6 28.0 32.8 23.3

84388610 tktA Carbon Cycling 89.9 74.9 88.1 22.2 30.0 19.1

377807676 vana Carbon Cycling 90.3 74.3 86.2 0.0 0.0 0.0

29832895 TerD Metal Homeostasis 82.7 71.1 95.8 17.1 25.1 16.2

205363968

lycopene beta

cyclase

Secondary

metabolism 86.0 69.1 93.0 17.5 25.6 16.4

356639591 one ring 23diox

Organic

Remediation 91.9 69.8 85.5 20.1 29.0 17.7

294629873 MFS antibiotic

Antibiotic

resistance 87.7 66.8 91.0 17.6 23.0 16.4

336321746 MFS antibiotic

Antibiotic

resistance 88.8 74.5 81.0 22.2 29.9 19.1

148257018 Fnr Stress 87.1 75.1 79.0 29.2 35.8 22.4

292815360 mntH Nramp Metal Homeostasis 86.4 67.9 86.3 24.1 27.5 22.6

356882203 Mex

Antibiotic

resistance 84.8 72.9 82.7 21.7 29.5 18.6

74318372 ompR Stress 86.2 74.0 79.8 0.0 0.0 0.0

70728315 degP Stress 78.4 75.3 84.5 22.0 27.2 20.5

315500429 Mex

Antibiotic

resistance 81.5 67.6 87.7 18.1 27.3 16.5

380766219 mcra Carbon Cycling 82.5 67.9 85.9 17.5 25.7 16.7

209959496 b lactamase

Antibiotic

resistance 87.3 66.0 82.7 17.5 25.8 16.7

357408389 amyA Carbon Cycling 85.5 68.0 79.2 18.5 25.0 17.9

340360549 amyA Carbon Cycling 76.8 68.5 83.6 19.0 26.8 17.3

182411857 mannanase Carbon Cycling 78.5 66.7 77.6 17.9 25.1 16.5

331021446 sodA Stress 80.1 63.6 78.6 19.0 6.5 16.0

116104071 Cas6e Other 78.5 65.5 78.1 26.2 32.2 19.9

326315542 ChrA Metal Homeostasis 78.9 64.7 75.9 21.6 30.6 18.1

295690051 Mex

Antibiotic

resistance 67.6 68.9 81.5 20.2 22.8 19.7

Continued

149

Table C.4 Continued

Genbank

ID Gene Gene category

2015 2016

CHLP OO PP CHLP OO PP

209958899 Fnr Stress 76.6 64.8 74.7 26.2 25.8 19.1

91802028 soxY Sulfur 78.7 64.0 70.6 20.7 29.2 20.4

153887635 urec Nitrogen 75.6 60.6 77.1 18.3 17.8 17.4

302554252 MFS antibiotic

Antibiotic

resistance 70.6 65.5 76.5 30.9 23.2 16.1

365864160 MFS antibiotic

Antibiotic

resistance 73.5 66.6 72.1 19.8 26.1 19.3

219949839 Fnr Stress 73.8 62.7 72.5 24.5 34.3 22.2

116608759 MFS antibiotic

Antibiotic

resistance 72.8 68.9 64.5 26.3 39.3 25.4

224825661 Iro virulence 65.7 69.4 68.7 17.8 26.2 17.0

255920442 AceB Carbon Cycling 68.9 64.6 69.5 24.8 29.8 20.3

429731628 pspA Stress 69.0 69.5 62.5 29.9 37.0 27.3

169637599 hdrB Carbon Cycling 67.8 64.9 67.7 22.2 31.7 22.0

259023078 nikA Metal Homeostasis 66.8 64.8 68.8 17.3 21.3 16.9

384568098

ABC antibiotic

transporter

Antibiotic

resistance 63.3 68.6 68.6 22.6 26.7 22.0

227820954 MFS antibiotic

Antibiotic

resistance 62.3 67.4 68.9 21.7 29.2 21.2

214028433 Tet

Antibiotic

resistance 75.7 62.3 60.0 26.0 24.7 21.5

115259366 nikA Metal Homeostasis 74.8 55.9 67.3 20.9 16.9 13.9

67524479 tannase Cdeg Carbon Cycling 73.0 55.4 69.4 24.7 31.4 21.5

170144304 sigma 24 Stress 64.1 62.9 69.2 17.0 22.2 16.3

170143040 Mex

Antibiotic

resistance 70.8 60.9 63.8 19.1 28.7 16.7

150

Genbank

ID Gene Gene category

2016 2015

CHLP OO PP CHLP OO PP

429731628 pspA Stress 29.9 37.0 27.3 69.0 69.5 62.5

345010049 corA Metal Homeostasis 30.2 36.9 24.9 100.1 94.4 93.0

116608759 MFS antibiotic Antibiotic resistance 26.3 39.3 25.4 72.8 68.9 64.5

148257018 fnr Stress 29.2 35.8 22.4 87.1 75.1 79.0

270731414 cap virulence 24.9 36.6 24.0 60.5 53.8 61.9

390569537 MFS antibiotic Antibiotic resistance 24.6 36.7 23.3 104.6 77.5 113.9

299134974 TIM Carbon Cycling 28.0 32.8 23.3 82.4 84.9 85.6

221723831 catechol b Organic Remediation 27.3 34.2 21.6 92.9 74.5 97.2

170741894 phytoene synthase

Secondary

metabolism 27.0 34.2 20.5 61.7 59.8 63.1

339611386 MATE antibiotic Antibiotic resistance 26.2 32.6 22.5 91.6 85.8 87.4

219949839 fnr Stress 24.5 34.3 22.2 73.8 62.7 72.5

418294838 pspA Stress 26.2 31.6 22.5 66.0 59.6 63.7

196259193 fnr Stress 24.8 32.4 23.0 120.3 79.3 104.6

327480404 MFS antibiotic Antibiotic resistance 25.6 32.3 21.3 128.8 88.5 108.4

312113990 Cas7 Other 23.9 32.5 22.6 58.4 56.1 61.5

284165967 phytoene synthase

Secondary

metabolism 26.2 32.2 19.9 141.9 88.1 121.5

116104071 Cas6e Other 26.2 32.2 19.9 78.5 65.5 78.1

254386677 corA Metal Homeostasis 25.9 30.4 21.9 50.6 59.8 59.1

217978651 ompR Stress 23.8 32.2 21.6 34.6 35.2 34.2

67524479 tannase Cdeg Carbon Cycling 24.7 31.4 21.5 73.0 55.4 69.4

315603069 nikA Metal Homeostasis 25.4 30.3 21.6 109.3 81.0 103.2

127514333 MFS antibiotic Antibiotic resistance 23.9 31.2 21.8 116.0 76.8 96.8

302407215 pel Cdeg Carbon Cycling 24.8 31.3 20.6 168.6 91.4 122.8

311112862 MFS antibiotic Antibiotic resistance 21.8 33.0 21.5 94.0 79.0 103.5

167561665 mdla Organic Remediation 24.0 31.6 20.7 89.7 77.4 91.1

169637599 hdrB Carbon Cycling 22.2 31.7 22.0 67.8 64.9 67.7

334103082 nikA Metal Homeostasis 21.7 31.4 22.0 115.2 71.6 97.9

255920442 AceB Carbon Cycling 24.8 29.8 20.3 68.9 64.6 69.5

292815360 mntH Nramp Metal Homeostasis 24.1 27.5 22.6 86.4 67.9 86.3

91802036 ompR Stress 22.9 29.9 21.3 127.4 75.6 95.6

Continued

Table C.5. Gene probe normalized signals for 99.9th percentile of detected values in the

GeoChip analysis on the sediments collected in 2015.

151

Genbank

ID Gene Gene category

2016 2015

CHLP OO PP CHLP OO PP

91978835 ompR Stress 22.2 29.9 20.4 95.5 71.7 93.9

214028433 Tet Antibiotic resistance 26.0 24.7 21.5 75.7 62.3 60.0

254388596 pstA Stress 26.2 26.1 19.8 95.5 80.6 96.7

227820954 MFS antibiotic Antibiotic resistance 21.7 29.2 21.2 62.3 67.4 68.9

330828074 Mex Antibiotic resistance 23.9 30.9 17.0 31.7 30.6 29.8

384568098

ABC antibiotic

transporter Antibiotic resistance 22.6 26.7 22.0 63.3 68.6 68.6

84388610 tktA Carbon Cycling 22.2 30.0 19.1 89.9 74.9 88.1

336321746 MFS antibiotic Antibiotic resistance 22.2 29.9 19.1 88.8 74.5 81.0

326772217 nikA Metal Homeostasis 23.0 26.4 21.8 49.8 46.8 56.7

209958899 fnr Stress 26.2 25.8 19.1 76.6 64.8 74.7

399074437 MFS antibiotic Antibiotic resistance 22.2 27.3 21.5 56.9 59.9 61.4

329118461 nikA Metal Homeostasis 21.2 31.1 18.2 126.4 73.5 98.2

297151115 MFS antibiotic Antibiotic resistance 28.5 21.9 20.0 60.3 61.4 69.6

91802028 soxY Sulfur 20.7 29.2 20.4 78.7 64.0 70.6

326315542 ChrA Metal Homeostasis 21.6 30.6 18.1 78.9 64.7 75.9

302554252 MFS antibiotic Antibiotic resistance 30.9 23.2 16.1 70.6 65.5 76.5

356882203 Mex Antibiotic resistance 21.7 29.5 18.6 84.8 72.9 82.7

70728315 degP Stress 22.0 27.2 20.5 78.4 75.3 84.5

114776318 sigma 38 Stress 22.5 24.5 22.1 36.3 39.1 34.9

170782088 SMR antibiotics Antibiotic resistance 24.7 22.6 21.5 33.6 28.8 28.9

61678850 pectinase (pectate lyase) Carbon Cycling 26.6 22.2 19.9 30.0 26.2 31.0

217969666 ompR Stress 26.2 21.5 20.9 54.5 44.1 43.5

309813148 cadBD Metal Homeostasis 22.4 22.4 23.4 23.3 27.9 26.3

260576032 nikA Metal Homeostasis 20.9 30.2 17.0 54.4 49.4 52.1

88807717 LPOR

Secondary

metabolism 21.8 26.2 19.9 55.9 48.8 57.3

153962580 sqr Sulfur 20.2 29.8 17.5 123.0 74.8 97.0

258520125 nirk Nitrogen 31.1 17.8 18.5 28.9 33.4 27.4

156972415 oxyR Stress 22.0 25.0 20.4 54.2 46.3 69.1

398070662 hrcA Stress 24.9 21.0 21.3 63.0 60.8 63.1

372476560 natB Metal Homeostasis 21.2 27.5 18.2 50.9 46.4 46.0

238757841 spiC virulence 19.4 30.7 16.8 96.8 70.9 98.6

356639591 one ring 23diox

Organic

Remediation 20.1 29.0 17.7 91.9 69.8 85.5

Continued

152

Table C.5 Continued

Genbank

ID Gene Gene category

2016 2015

CHLP OO PP CHLP OO PP

193223170 nikA Metal Homeostasis 26.1 19.6 20.1 50.9 45.4 49.9

297193687 phytase Phosphorus 19.6 28.9 17.3 84.6 77.6 94.7

148273451 nhaA Metal Homeostasis 19.7 26.4 19.3 106.6 80.1 105.1

241320734 cysJ Sulfur 19.6 28.1 17.8 65.2 58.8 59.8

219945981 ompR Stress 20.2 28.4 16.8 47.2 45.6 48.0

332187265 phytoene synthase

Secondary

metabolism 18.8 28.7 18.0 40.4 35.0 41.2

365864160 MFS antibiotic Antibiotic resistance 19.8 26.1 19.3 73.5 66.6 72.1

170738862 arsB Metal Homeostasis 19.6 28.7 16.7 118.1 71.8 97.8

171472824 sod FeMn

secondary

metabolism 19.6 28.0 17.4 110.3 68.3 94.8

170745160 NiCoT Metal Homeostasis 19.1 28.0 17.9 56.7 51.1 55.0

119716849 corA Metal Homeostasis 20.3 28.2 16.5 43.3 37.2 39.3

398193010 MFS antibiotic Antibiotic resistance 19.3 28.1 17.4 63.8 55.1 59.5

21929222 mnp Carbon Cycling 20.4 21.7 22.8 29.1 29.8 25.8

238059362 arsB Metal Homeostasis 21.8 23.2 19.7 92.0 79.0 100.9

42627732 catechol Organic Remediation 19.1 28.7 16.7 109.4 69.0 94.4

170143040 Mex Antibiotic resistance 19.1 28.7 16.7 70.8 60.9 63.8

255923242 nikA Metal Homeostasis 18.9 28.8 16.9 56.0 46.6 54.1

269097992 clpP Stress 20.4 27.8 16.0 35.5 35.0 36.2

50954857 gdh Nitrogen 18.7 29.1 16.3 132.2 67.1 99.0

153

Table C.6. Probe counts for the metal homeostasis gene probes. Probes are summarized

by their associated metal foe the samples collected across the two years.

CHLP OO PP

2015 2016 2015 2016 2015 2016

Iron 2011 2228 2165 2283 1768 2237

Nickel 1406 1461 1493 1503 1262 1476

Potassium 1290 1431 1409 1453 1094 1410

Sodium 1105 1169 1181 1235 958 1190

Magnesium 994 1102 1075 1130 869 1116

Zinc 948 1004 999 1021 835 989

Arsenic 900 970 945 972 795 960

Copper 856 912 898 910 735 904

Tellurium 844 869 881 892 743 897

Chromium 686 685 731 720 625 706

Mercury 489 484 492 496 426 494

Silver 461 464 477 461 397 474

Manganese 440 480 467 474 377 477

Cadmium 399 406 411 412 346 402

Calcium 71 71 73 77 63 74

Aluminum 59 69 67 70 51 73

Cobalt 58 68 59 64 50 67

Lead 21 17 20 19 17 21

Silicon 9 12 9 13 7 10

Boron 1 5 2 2 0 4

Selenium 1 1 2 1 1 1

Magnesium Cobalt 654 692 675 671 578 673

Zinc Cadmium Cobalt 477 496 496 508 413 509

Multiple metals 450 477 458 501 384 478

Nickel Cobalt 25 22 26 28 20 26

154

Table C.7. The functionality of the metal genes detected across both GeoChip datasets.

Any positive probe detected across the samples was counted once.

Metal Transport Detoxification Sequestration Storage Biosynthesis Total

Iron 1906 0 0 360 0 2266

Nickel 1539 0 0 0 0 1539

Potassium 1471 0 0 0 0 1471

Sodium 1237 0 0 0 0 1237

Magnesium 1127 0 0 0 0 1127

Zinc 1034 0 0 0 0 1034

Arsenic 405 580 0 0 0 985

Copper 917 21 0 0 0 938

Tellurium 424 495 0 0 0 919

Chromium 736 15 0 0 0 751

Cobalt Magnesium 705 0 0 0 0 705

Cadmium Cobalt

Zinc 517 0 0 0 0 517

Mercury 131 385 0 0 0 516

Silver 497 0 0 0 0 497

Manganese 496 0 0 0 0 496

Multiple metals 478 0 15 0 0 493

Cadmium 430 0 0 0 0 430

Calcium 76 0 0 0 0 76

Aluminum 69 0 0 0 0 69

Cobalt 64 0 0 0 0 64

Cobalt Nickel 27 0 0 0 0 27

Lead 22 0 0 0 0 22

Silicon 8 0 0 0 1 9

Selenium 0 2 0 0 0 2

Boron 2 0 0 0 0 2

155

Figure C.1.Probe counts in the organic remediation category of the GeoChip analysis.

Each count is further described by the subcategory, and, for the aromatics subgroup, to

the secondary subcategory. Probe counts were averaged across the three samples

analyzed in each year, and error bars represent the standard deviation for those counts.

156

Table C.8. Fluidigm readings across samples and replicates. A’-‘ indicates that the gene

was not detected or was outside the range of detection for that replicate.

CHLP OO PP

Gene rep 1 rep 2 rep 3 rep 1 rep 2 rep 3 rep 1 rep 2 rep 3

aacA 4748 12531 7143 3131 5664 1695 91999 22555 26726

aadA5 - - - - - - - - -

aadD - - - - - - - - -

acrD - - - - - - - - -

ampC - - 781 1011 - 898 - - -

arr2 - - - - - - - - -

blaKPC - - - - - - - - -

blaNDM1 - - - - - - - - -

blaNPS - - - - - - - - -

blaOXA 2536 - 1654 1552 3821 5325 7497 - -

blaSHV 97124 138334 203061 35451 60575 94571 180270 141975 94078

blaVIM - - - - - - - - -

cadA 1806 1058 2171 1536 2671 2655 6829 6110 4618

catB8 - - - - - - - - -

chrA 494 - - 1106 1303 - - - -

cmlB - - 978 - 1093 - - - -

copA - - 791 - - - - - -

ctxm32 693 - 1652 - 1727 1206 - - -

dfr13 - - - - - - - - -

ereB - - - - - - - - -

ermB - - - - - - - - -

ermF - - - - - - - - -

floR - - - - - - - - -

imp13 - - - - - - - - -

intI1 87262 109009 176937 256811 320315 510983 521073 407451 385500

intI2 603 - 1596 1236 - - - - -

intI3 - - - - - - - - -

mefE - - - - - - - - -

merA 1734 1142 2298 99717 61077 63225 4606 4321 35304

mexB 18684 12285 12362 14657 19711 21337 40898 49303 28259

qacF - - - - - - - - -

qnrA - - - - - - - - -

qnrB 499 - 1487 1914 1314 1523 2027 620 -

rcnA - - - - - - - - -

strB 665 64 490 2029 854 1685 - 2042 946

sul1 1428 - 780 5102 3173 1530 4868 8085 84305

sul2 1566 - 2526 5019 4897 5632 2614 9240 18977

sul3 671 - 1193 1125 1308 1114 1089 662 875

tetA 4762 2581 5205 3959 - 9236 22800 20428 12379

tetL 450 493 269 184 674 772 1103 1018 2536

tetM - - - - - - - - -

tetS - - - - - - - - -

tetW 642 - 734 1109 926 846 - - -

tetX - - - - - - - - -

vanA - - - - - - - - -

vanB - - 971 - - 1396 - - -

157

Table C.9. Sequence reads from Illumina sequencing

Sample Sequences OTU97 Berger

Parker Shannon

PD Whole

Tree

2015

CHLP 6,695 1687 0.003 7.36 88.5

OO 2,348 1531 0.005 7.19 83

PP 1,990 1578 0.007 7.26 85.6

2016

CHLP 16,953 890 0.021 6.31 50.8

OO 34,333 971 0.015 6.45 56.5

PP 16,951 955 0.023 6.43 53.4

Figure C..2 The number of gene probes related to metal homeostasis and concentrations

measured for those metals. The number of probes were determined for the entire number

measured across the 2016 GeoChip set, while the concentration of the metal was the

average value detected in the three sediments collected that same year. Error bars

represent the standard deviation for the three sediments, following the averaging of

replicates between sites. The abundance of gene probes for multiple metals were not

accounted for in their respective metal groups.