Sources and Public Perceptions of Contaminants in the Lower Portneuf River Valley: A Case Study for Nitrates and Personal Care Products and Pharmaceuticals

by

Courtney Alecia Ohr

A thesis submitted in partial fulfillment of the requirements for the degree of Master of Science in the Department of Geosciences State University Spring 2016

Use Authorization

In presenting this thesis in partial fulfillment of the requirements for an advanced degree

at Idaho State University, I agree that the Library shall make it freely available for inspection. I further state that permission to download and/or print my thesis for scholarly

purposes may be granted by the Dean of the Graduate School, Dean of my academic division, or by the University Librarian. It is understood that any copying or publication

of this thesis for financial gain shall not be allowed without my written permission.

Signature ______

Date ______

ii

Committee Approval

To the Graduate Faculty:

The members of the committee appointed to examine the thesis of COURTNEY ALECIA OHR find it satisfactory and recommend that it be accepted.

______Sarah E. Godsey, Major Advisor

______John A. Welhan, Committee Member

______Danelle Larson, Committee Member

______Kathleen Lohse, Committee Member

______Dewayne Derryberry, Graduate Faculty Representative

iii

Table of Contents List of Figures ...... viii List of Tables ...... x List of Equations ...... xi List of Acronyms and Abbreviations ...... xii Abstract ...... xiii 1 Chapter 1: Introduction ...... 1 1.1 Motivation ...... 1 1.2 Broader impacts ...... 2 1.3 Study area ...... 2 1.3.1 Urban growth ...... 4 1.3.2 Previous studies of the LPRV ...... 5 1.3.2.1 Private well contamination ...... 5 1.3.2.2 Other known contaminants in the LPRV aquifer ...... 5 1.3.2.3 Recharge ...... 6 1.4 Nitrate contamination and legislation in Idaho ...... 6 1.5 Background of nitrate in groundwater ...... 7 1.5.1 The nitrogen cycle ...... 7 1.5.1.1 Fixation ...... 8 1.5.1.2 Nitrification ...... 8 1.5.1.3 Denitrification ...... 9 1.5.1.4 N leaching in groundwater ...... 9 1.6 The impact of ex-urban development on PPCPs and the nitrogen cycle ...... 9 1.6.1 Septic systems influence on nitrate in groundwater ...... 11 1.7 Geochemical tracers that can determine sources of nitrate-N ...... 11 1.7.1 Isotopes ...... 11

1.7.1.1 Fingerprinting techniques using 15N and  18O from NO3-N ...... 11 1.7.1.2 Using 18O, 2H from well water to interpret groundwater recharge sources ...... 14 1.7.2 Pharmaceuticals and personal care products ...... 15 1.8 Health effects of nitrate and PPCP contamination ...... 18 1.9 Water quality as an ecosystem service ...... 19 1.10 Thesis organization and objectives ...... 20 1.10.1 Chapter 1 ...... 20 1.10.2 Chapter 2 ...... 20 1.10.3 Chapter 3 ...... 20

iv

1.10.4 Chapter 4 ...... 21 References ...... 22 2 Chapter 2: Sources and patterns of drinking water contamination in the LPRV ...... 31 Abstract ...... 31 2.1 Introduction ...... 31 2.2 Study Area ...... 33 2.3 Methods ...... 35 2.3.1 Previous sampling campaigns and data ...... 35 2.3.2 Sampling methods ...... 36 2.3.2.1 Field sampling plan ...... 36 2.3.2.2 Field sampling methods ...... 39 2.3.3 Lab sample processing and analysis ...... 42 2.3.3.1 Sample pre-processing ...... 42 2.3.3.2 Geochemistry for nitrate and groundwater analysis...... 43 2.3.3.3 Emerging Contaminants ...... 43 2.3.4 Statistical and isotopic data analysis ...... 44 2.3.4.1 Temporal trend of nitrate-N concentrations ...... 44 2.3.4.2 Ordinary Kriging ...... 45 2.3.4.3 Getis-Ord Gi* hotspot analysis ...... 46 2.3.4.4 Well capture zone and septic density assignment ...... 47 2.3.4.5 Isotopes O and H as groundwater recharge identifier ...... 50 2.3.4.6 Nitrate isotopic corrections and interpretations ...... 53 2.3.4.7 Major anions as septic tracers ...... 54 2.3.4.8 Presence of PPCPs compared to nitrate-N concentrations ...... 55 2.4 Results ...... 56 2.4.1 Nitrate-N temporal trend ...... 56 2.4.2 Nitrate Prediction and Hotspot Map...... 57 2.4.3 Septic Density vs. Nitrate ...... 58 2.4.4 Spatial Distribution of Nitrate Sources ...... 59 2.4.5 Groundwater Recharge Sources Using Isotopes 18O and 2H ...... 61 2.4.6 Emerging Contaminants ...... 64 2.4.6.1 Major anions ...... 66 2.5 Discussion ...... 70 2.5.1 Nitrate across the U.S...... 70 2.5.2 Nitrate sources and patterns ...... 71 2.5.3 Influences of groundwater recharge on nitrate ...... 73

v

2.5.4 Factors affecting transformation of nitrate during movement through hydrologic systems 76 2.5.5 Emerging contaminants ...... 78 2.5.5.1 Carbamazepine ...... 78 2.5.5.2 DEET ...... 79 2.5.5.3 Diphenhydramine ...... 80 2.5.5.4 Codeine ...... 81 2.5.5.5 Ibuprofen ...... 81 2.5.5.6 Fluoxetine ...... 82 2.5.5.7 Sucralose ...... 83 2.5.5.8 Sulfamethoxazole ...... 83 2.5.5.9 Sulfadimethoxine ...... 84 2.5.5.10 Multiple PPCPs Detected ...... 85 2.5.6 Nitrate and PPCPs ...... 85 2.5.7 Major anions as septic tracers ...... 86 2.5.8 Wells with unknown sources of nitrate-N ...... 87 2.6 Conclusions ...... 88 2.7 Acknowledgements ...... 90 2.8 References ...... 91 3 Chapter 3: Factors Influencing Public Risk Perceptions of Drinking Water Quality in the LPRV and its Effect on Behavior ...... 99 Abstract ...... 99 3.1 Introduction ...... 99 3.1.1 Groundwater and Water Quality in Pocatello, ID ...... 100 3.1.2 Risk perception ...... 101 3.1.3 1.4 Risk Perception of Water Quality ...... 103 3.2 Methods ...... 105 3.3 Results ...... 108 3.3.1 Softening and Filtering ...... 109 3.3.2 Testing and Treating ...... 110 3.3.3 Buying Bottled Water ...... 111 3.3.4 Respondents who reported doing nothing to their water supply ...... 112 3.4 Discussion ...... 113 3.4.1 Risk as analysis...... 113 3.4.2 Risk as feelings ...... 113 3.4.3 Risk as politics ...... 114 3.4.4 Other independent variables ...... 115

vi

3.5 Conclusion ...... 116 3.6 Acknowledgments ...... 116 3.7 References ...... 117 4 Conclusions and Remarks ...... 120 4.1 Summary of thesis work ...... 120 4.1.1 Spatial patterns and sources of Nitrate-N ...... 120 4.1.2 Public risk perceptions that drive action toward water quality improvement ...... 120 4.1.3 Integrating public risk perceptions with measured water quality ...... 121 4.2 Opportunities for future work ...... 122 4.3 References ...... 124 5 Appendices: ...... 125 5.1 Appendix A: PPCPs and Detection Limits ...... 125 5.2 Appendix B: Clean Hand/Dirty Hand Jobs ...... 126 Both CH and DH ...... 126 CH tasks ...... 126 DH tasks ...... 126 5.3 Appendix C: Sampling and Cleaning Procedures ...... 127 5.3.1 Pre-field preparation and cleaning ...... 127 5.3.2 Field Sampling ...... 129 5.3.3 Laboratory Processing of Samples ...... 134 5.3.4 EXTRA Laboratory Preparation and Cleaning ...... 137 5.4 Appendix D: Total dissolved solids, anions, and nitrate-N analyses ...... 139 5.4.1 Total dissolved solids (TDS) and nitrate-N ...... 139 5.4.2 Anions ...... 140 5.5 Appendix E: Cation samples ...... 144 5.6 Appendix F: Septic-impacted wells ...... 144 5.7 Appendix G: Additional information about groundwater, snowpack, and rainfall isotopic signatures ...... 145 5.8 Appendix H: Survey Questions ...... 146 5.9 Appendix I: 2015 Sampling Results ...... 5-150 5.9.1 2015 Results for nitrate-N, DIC, isotopes, anions ...... 5-150 5.9.2 2015 Results for PPCPS ...... 5-155

vii

List of Figures

Figure 1.3.1 Study area and aquifer locations within the LPRV. NAIP imagery distributed by the Land Processes Distributed Active Archive Center (LP DAAC), located at USGS/EROS, Sioux Falls, SD. http://lpdaac.usgs.gov...... 4 Figure 1.7.1 Nitrate δ15N and δ18O values with their corresponding source ranges (adapted from Kendall 1998)...... 12 Figure 1.7.2 Pathways for input and transport of pharmaceuticals into drinking water. This Figure shows human vs. veterinary pharmaceuticals and is a simplified transport diagram adapted from Figure 1.1 in Kummerer (2004)...... 16 Figure 2.2.1 Location and geology of the LPRV. NAIP imagery distributed by the Land Processes Distributed Active Archive Center (LP DAAC), located at USGS/EROS, Sioux Falls, SD. http://lpdaac.usgs.gov...... 35 Figure 2.3.1 2015 well sampling plan highlighting the four targeted strata and the intended locations for resampling (orange circles) and new sampling (blue triangles). Septic density (septic/km2) is displayed in contoured colors with lowest densities uncolored and highest densities in dark orange...... 38 Figure 2.3.2 Wells sampled in 2015...... 39 Figure 2.3.3 Sampling collection system. All parts shown in red and pink were replaced for every sample...... 40 Figure 2.3.4 Sources of nitrate identified using corrected δ18O versus δ15N (adapted from Dejwakh et al. 2011 & Kendall 1998)...... 54 Figure 2.4.1b Nitrate-N concentrations from Caribou Acres well #ID6030005 compared to a 5-year running average of snow water equivalent (SWE) from the Wildhorse Divide SNOTEL site ...... 57 Figure 2.4.2 Predicted nitrate-N based on ordinary kriging, and hotspots (purple) and coldspot (blue) based on a Getis-Ord Gi* analysis. NAIP imagery distributed by the Land Processes Distributed Active Archive Center (LP DAAC), located at USGS/EROS, Sioux Falls, SD. http://lpdaac.usgs.gov...... 58 Figure 2.4.3(a-d) Septic density is a poor predictor of nitrate-N concentrations whether wells are assessed all together, grouped by eastern/western hotspot or if different capture zone methods are used. (a) Septic density vs. nitrate-N using least cost path analysis to define well capture zones. (b) Least cost path analysis capture zones showing only wells located in the east and west hotspots. (c) Septic density vs. nitrate-N using an omnidirectional method to define capture zone and showing only wells located in the east and west hotspots. (d) Septic density vs. nitrate-N using a directional method to define capture zone and showing only wells located in the east and west hotspots...... 59 Figure 2.4.4 δ15NNO3-N (‰) vs. δ18ONO3-N (‰) with δ15NNO3-N corrected for △17O. Point size scaled to nitrate-N (mg/L), with larger points representing greater concentrations ...... 60 Figure 2.4.5 δ15NNO3-N (‰) vs. δ18ONO3-N (‰) with δ15NNO3-N corrected for △17O (adapted from Kendall 1998). Symbol shape indicates likely sources, where M/S is manure or septic and F/P is nitrate derived from fertilizer or precipitation. Blue points indicate wells in which PPCPs were detected and red points indicate wells in which no PPCPs were detected. Larger points represent higher nitrate-N concentrations...... 60 Figure 2.4.6 Wells with nitrate isotope signatures that share boundaries with soil N, fertilizer/precipitation and/or manure/septic (Figure 2.4.5); but were not identified as septic-impacted with other tracing methods (anions/PPCPs)...... 61 Figure 2.4.7 Diagnostic tests for denitrification fractionation versus nitrate mixing sources; (a) nitrate-N exhibits no systematic relationship with δ15N NO3--N. (b) The multiplicative inverse of nitrate-N concentration also exhibits no clear relationship with δ15NNO--N, (c) A spline fit to the natural log of nitrate-N vs. δ15N NO3-N demonstrates a weak curvilinear relationship. Figures b and c indicate there is mixing of nitrate-N sources (i.e. manure/septic sources mixed with soil N sources) rather than fractionation due to

viii

denitrification. If denitrification fractionation had occurred, Figure b would show a curved relationship and Figure c would display a linear relationship...... 61 Figure 2.4.8 D-excess found in groundwater sampled from wells across the LPRV. Lower d-excess indicates less evaporative fractionation or older recharge sources, whereas higher values indicate modern recharge sources or more evaporative fractionation...... 62 Figure 2.4.9 Groundwater sample signatures (blue) with three possible recharge source signatures, rainfall (diamonds), snowpack (crosses), and stream (triangle). Both rainfall and snowpack signatures fall on the local meteoric water line (LMWL) and global meteoric water line (GMWL). Groundwater samples, by contrast, exhibit a trend off the LMWL. .. 63 Figure 2.4.10 Grouping analysis results in four groups of wells in the LPRV based on d-excess and δ18O signatures ...... 63 Figure 2.4.11 Higher nitrate-N concentrations were found in wells where PPCPs were detected...... 66 Figure 2.4.12 Chloride and sulfate concentrations for wells sampled in the 2015 campaign. Different shapes reflect identified nitrate-N sources (Figure 2.4.4) and colors reflect whether PPCPs were detected. Wells with combined manure/septic (M/S) and soil N sources (squares) tend to have higher chloride and sulfate concentrations. Wells with combined sources that include all three potential nitrate-N sources, including ammonium (triangles) have low to middle ranges of sulfate and chloride...... 67 Figure 2.4.13 Chloride and nitrate-N concentrations are not strongly correlated. Identified nitrate-N sources are shown as indicated in Figure 2.4.12. High Cl (>100 ppm), low nitrate (<10ppm) well waters include mixed soil N and manure or septic sources (including squares and triangles) as well as ammonium from fertilizer or precipitation (including crosses and triangles). Possible contributions from manure and septic sources (including diamonds, squares, and triangles) are found across the range of both Cl and nitrate-N concentrations...... 68 Figure 2.4.14 As with Figure 2.4.12, but for the correlation between sulfate and nitrate-N. Similarly, septic and manure sources are found across the range of sulfate and nitrate concentrations...... 68 Figure 2.4.15 Total dissolved solids (TDS) are strongly correlated with specific conductivity (SC) as reported for the 2004 well samples (Meehan 2005). This relationship was used to calculate 2015 TDS from measured SC to assess potential locations of denitrification (see Figure 2.4.16)...... 69 Figure 2.4.16 Spatial distribution of wells sampled in 2015 with high TDS/high nitrate, high TDS/low nitrate, and low TDS/low nitrate well samples. Wells with values of nitrate or TDS that fell within the interquartile range of either distribution are not shown. A single well exhibits high TDS/low nitrate concentrations (blue square)...... 69 Figure 3.1.1 Pathways for input and transport of pharmaceuticals into drinking water. Both human and. veterinary pharmaceuticals can contribute to drinking water contamination (adapted from Figure 1.1 in Kummerer 2004)...... 101 Figure 3.2.1 Independent variables used in maximum-likelihood probit model for testing the relationship between these controls, risk perception and dependent action variables (filtering, softening, treating, testing, buying bottled water, or nothing) ...... 108 Figure 3.3.1 Significant factors that influence risk perception for survey respondents who soften and filter their water. All continuous variables show the sign of the relationship (+/-)...... 109 Figure 3.3.2 Significant factors that influence risk perception for survey respondents who test and treat their water. All continuous variables show the sign of the relationship (+/-)...... 110 Figure 3.3.3 Significant factors influencing risk perception for survey respondents who buy bottled water. All continuous variables show the sign of the relationship (+/-)...... 111 Figure 3.3.4 Significant factors influencing risk perception for survey respondents who do nothing. All continuous variables show the sign of the relationship (+/-)...... 113

ix

List of Tables

Table 2.3.1 Stability requirements for groundwater that were met prior to sampling (adapted from USGS Book 9; Wilde et al. 2006)...... 41 Table 2.4.1 Median nitrate concentrations of paired wells sampled at varying time period ...... 57 Table 2.4.2 Mean (+/- s.e.) isotopic signatures of each group of wells identified in Figure 2.4.9 and mean isotopic signatures of potential recharge sources, snowpack and rainfall ...... 64 Table 2.4.3 Percent contributions to groundwater recharge (and the associated standard error) from snowpack and rainfall end-members based on IsoError analyses of the mean 18O values reported in Table 2.4.2...... 64 Table 2.4.4 Nine pharmaceuticals and personal care products were detected in the 100 private LPRV wells tested in 2015. Some PPCPs were found in only a single well whereas others were detected more frequently. If a substance was detected in 5 or more wells, it was found in wells on both sides of the river. PPCPs that were only infrequently detected were found exclusively west of the river. Concentrations consistently remained <1ppm across all PPCPs...... 65 Table 3.2.1 Demographics that make-up of survey respondents ...... 106 Table 3.3.1 Factors influencing risk perception that lead to softening and filtering. The strongest controls here are demographic. The factor “wastewater system” is positive when the respondent has a residential septic system and negative when the respondent is using city sewer...... 109 Table 3.3.2 Factors influencing risk perception that lead to testing and treating water. The strongest controls are related to education, concern, and whether the respondent’s primary water source was private or municipal...... 110 Table 3.3.3Factors influencing risk perception that lead to buying bottled water ...... 111 Table 3.3.4 Factors influencing risk perception that lead to respondents doing nothing to their water supply ...... 112

x

List of Equations

Equation 2.3.1 Ordinary Kriging Model ...... 45 Equation 2.3.2 Weighting factor from three unknown points ...... 46 Equation 2.3.3 Nitrate concentration at unknown point ...... 46 Equation 2.3.4 Getis-Ord Gi* ...... 47 Equation 2.3.5 Mean nitrate-N concentration ...... 47 Equation 2.3.6 Standard Deviation for all wells ...... 47 Equation 2.3.7 Between group differences ...... 50 Equation 2.3.8 Within group differences ...... 50 Equation 2.3.9 Grouping Analysis ...... 51 Equation 2.3.10 Isotopic mass balance from IsoError ...... 52 Equation 2.3.11 Source proportions of well recharge ...... 52 Equation 2.3.12 Approximation of variances within IsoError ...... 53 Equation 2.3.13 Mann-Whitney ranked medians ...... 55

xi

List of Acronyms and Abbreviations

amsl: above mean sea level Bgal: Billion Gallons EPA: Environmental Protection Agency Gpd: Gallons per day IDEQ: Idaho Department of Environmental Quality ISU: Idaho State University LPRV: Lower Portneuf River Valley MCL: Maximum Contaminant Level mg/L: milligrams per Liter PPCPs: Pharmaceuticals and Personal Care Products TDS: Total Dissolved Solids U.S.: USGS: United State Geological Survey WUI: Wildland Urban Interface

xii

Abstract

Contamination of drinking water from septic systems and leaking sewage pipes may outpace agricultural contributions to nitrate contamination at urban-wildland interfaces.

In this project, I identified septic systems as the most probable sources of nitrate-N contamination in approximately 40% of the 100 wells that I sampled in the Lower

Portneuf River Valley (LPRV), Idaho, and possible sources for an additional 25%. I found source identification is more complex than previously thought likely owing to mixing of multiple sources. I also identified the main factors affecting public risk perceptions and residential modification of water quality; measured quality did not appear to influence public risk perceptions. Instead, three factors were more important: whether or not water was privately supplied, the importance of risk perception associated with unknown dangers, and education level of public survey respondents. These results may help policymakers in making decisions about priorities for extensions in sewer and water services.

xiii

1 Chapter 1: Introduction

1.1 Motivation

Humans depend on clean groundwater for municipal supplies in many cities. The

U.S. Geological Survey (2005) estimated that ~79.6 billion gallons per day (Bgal/day) were withdrawn from groundwater reserves in the United States (Kenny et al. 2009).

Concurrently, groundwater contamination is common and a growing concern in mid- sized cities due to water quality impacts of urban infrastructure, such as high-density septic systems or leaking sewage pipes (O’Driscoll et al. 2010). Reliance on quality groundwater to support mid-sized cities can be observed in southeastern Idaho where the cities of Pocatello and Chubbuck rely on the Lower Portneuf River Valley (LPRV) aquifer as the primary source of potable water. The LPRV aquifer may be vulnerable to contamination due to its geology and increased urban growth in the region.

Groundwater contamination is often controversial because there are multiple

- potential sources of contamination. Common contaminants such as nitrate (NO3-N ) can come from fertilizer, human waste, animal waste, and atmospheric deposition. Less familiar contaminants such as pharmaceuticals and personal care products (PPCPs) often come from human sources rather than agricultural sources (Idaho DEQ 2014, Seiler et al.

1999). Many people think groundwater contamination in the LPRV is due to agricultural influences. However, Meehan (2005) suggested that human waste through septic leakage may be another input, and called for further study to adequately identify nitrate sources.

Unless the sources of groundwater contamination in the LPRV are identified, the community may waste resources on the mitigation of misidentified sources of contamination.

1

1.2 Broader impacts

My thesis has provided information specific to the cities of Pocatello and

Chubbuck in southeastern Idaho, but will also be relevant to mid-sized cities elsewhere.

Cities with similar topography, geology, and growth scenarios can utilize the outputs of this study to make similar inferences about their own contaminant issues or create preemptive policy measures to prevent groundwater contamination. By collaborating with social scientists, my thesis has also assessed the public’s perception of risk levels associated with particular contaminants and the actions encouraged by these perceptions.

Such information can lead to better policy decisions on infrastructure growth and land use. For instance, it may be in the best interest of the city to limit septic systems and extend city services. However, extending city services and other groundwater protection policies can have significant economic consequences for individuals and city taxpayers

(Abdalla 1994), therefore it is important to actually identify nitrate sources and their relative contributions to drinking water contamination.

1.3 Study area

The cities of Chubbuck and Pocatello are located in the northwest-trending LPRV

(-112.415062⁰N, 42.830605⁰W). The valley is bounded by the Bannock Range on the southwest and by the Pocatello Range on the northeast (Figure 1.3.1). The valley floor is nearly flat and is ~1400 meters above mean sea level (amsl) with the Bannock Range reaching >2000 meters amsl. The northern extent of the LPRV merges with the historically volcanic eastern Plain and its southern extent meets the Portneuf

Gap, a narrow bedrock gorge that leads to the Upper Portneuf (Figure 1.3.1). The area has a semi-arid climate with most of its precipitation falling as winter snow (Welhan

2

2006). The primary valley drainage is the Portneuf River, originating in the Gem Valley and Marsh Creek headwaters. The river passes through the Portneuf Gap and flows northwest through the valley into the Snake River.

The geology of the LPRV aquifer contributes to the vulnerability of its water quality. The LPRV aquifer is divided into three sub-systems because of changes in subsurface geology: the northern, eastern, and southern aquifer systems (Figure 1.3.1).

Groundwater in the shallow southern aquifer (~750 m thick) and the eastern Snake River

Plain aquifer recharges the deeper northern aquifer system (~1300 m thick) (Reid 1997).

The northern aquifer is comprised of poorly sorted silty clay-rich gravels and sands. The southern aquifer is a relatively shallow strip aquifer made of permeable Bonneville flood gravels that permit high linear flow velocities (Welhan et al. 1996). Wells in the southern aquifer have high water yields at 30-45 meters below the ground surface. The southern aquifer’s high conductivity allows soluble contaminants to quickly move through it and contaminate the northern aquifer. Because the southern aquifer contributes 5.3 Bgal/yr of recharge to the northern aquifer (Welhan, 2006), it is important to monitor the contaminants entering the southern aquifer. The southern aquifer is separated from the eastern aquifer by a basalt layer from the Portneuf Basalt flow. The sediments underneath the basalt layer consist of low permeable valley fill, most likely of Quaternary or late

Tertiary age. The shallow alluvial/colluvial eastern aquifer comprises poorly sorted clay- rich boulder gravel and silty sand (Welhan et al. 1996). The west side of the southern aquifer is bounded by Proterozoic and early Paleozoic rocks and alluvial fan deposits.

3

Figure 1.3.1 Study area and aquifer locations within the LPRV. NAIP imagery distributed by the Land Processes Distributed Active Archive Center (LP DAAC), located at USGS/EROS, Sioux Falls, SD. http://lpdaac.usgs.gov.

1.3.1 Urban growth

Between 1985 and 2002, Pocatello’s urban area grew by 34% accompanied by a

65% increase in low-density development on the edge or outside of the municipal boundary (McMahan et al. 2002). Parcels of land 90 meters away from an existing sewer line are currently permitted for septic systems because they limit infrastructure expenses for sewage, water, and other various services outside of city limits. Residential areas with

4 private well water supplies are the main focus for this study. As development increases outside of municipal city limits, so does septic density and the potential for contamination of Pocatello’s drinking water. Land use in the southern valley floor is dominantly rural- residential with minor agricultural use. Residential development has spread from the valley to the benches and tributary ranges flanking the Bannock Range. The higher elevations are mostly under the jurisdiction of two federal agencies, the Bureau of Land

Management and the U.S. Forest Service. Within their property boundaries, these agencies have the ability to permit livestock grazing, which can also influence nitrate contamination in groundwater.

1.3.2 Previous studies of the LPRV

1.3.2.1 Private well contamination

Meehan (2005) showed that wells in the LRPV that produced water from near the water table had higher concentrations of nitrate and salts. He also showed that ratios of chloride to nitrate and chloride to sulfate in groundwater fell within a range consistent with septic field effluent, which indicated that nitrate may have originated from human waste and salts from water softener systems (Meehan, 2005). Meehan did not study pharmaceutical or personal care products in the 2005 study.

1.3.2.2 Other known contaminants in the LPRV aquifer

In the 1990s, a trichloroethylene (TCE) plume was detected in the LPRV aquifer, and nitrate levels occasionally exceeded the Environmental Protection Agency’s (EPA) maximum contaminant level of 10 mg/L (CH2M-Hill 1994, Welhan et al. 1996, Meehan

2005). The southern aquifer has been impacted by trichloroethylene (TCE) contamination that extended along the western boundary of the southern aquifer (CH2M-Hill 1994). The

5 high linear flow velocities created an elongated, narrow dispersion pattern that likely originated at the landfill (Meehan 2005). These events led to additional drinking water treatment, remediation of the TCE contamination source, and well abandonment.

One of the remediation tasks was to educate the public on the importance of well testing and the harmful effects of long-term exposure to nitrate (CH2M-Hill 1994). Since the detection of TCE, there have been multiple air strippers installed around the landfill to remove the contaminant.

1.3.2.3 Recharge

Through water balance calculations, Welhan (2006) found that ~74.5% of the southern aquifer’s recharge (or 5.4 ± 0.1 Bgal/year) stems from the Mink Creek/Gibson

Jack/City-Cusick Creek watersheds to the west of the LPRV. The southern aquifer is dependent on this principal recharge source to replenish storage that may be depleted during peak pumping periods.

1.4 Nitrate contamination and legislation in Idaho

Nitrate Priority Areas (NPA) were previously examined by the USGS for wells sampled between the years 1961-2002. The Idaho Department of Environmental Quality

(IDEQ) updated the USGS study in 2008 and again in 2014, when spatiotemporal trends were analyzed and ranked statewide by severity of contamination (Neely 2013). Under current rules, an NPA is delineated if 25% of the wells sampled in the area have nitrate-N concentrations of at least 5 mg/L. Statewide, 25 NPAs were defined in 2002, and this was updated to 32 NPAs in 2008 and 34 NPAs in 2014. Pocatello and the immediate area have three NPAs: Black Cliffs, North Pocatello, and Mink Creek, which have statewide priority rankings of 10, 16, and 20 respectively in 2014’s NPA list. When Idaho NPAs

6 were analyzed for temporal trends, these 3 NPAs were not included because they had few wells that were repeatedly sampled (Neely 2013). Nitrate sources in Idaho are primarily from fertilizers, septic systems, livestock operations, and industries such as food processing (West 2001, Schorzman and Baldwin 2009).

The IDEQ has a groundwater quality improvement plan where a major strategy is public education. The costs of nitrate contamination mitigation are funded by taxpayers, utility customers, and well owners (West 2001). In Idaho, groundwater is the source of

95% of the public water supply systems. Public drinking wells are regulated under the

- Safe Drinking Water Act, which requires nitrate-nitrogen levels below 10 mg/L NO3 -N

(Bowser et al. 1996). The Idaho statute Ground Water Quality Rule (IDAPA 58, Title 01,

Chapter 11) established authority for groundwater protection for all Idaho agencies and programs (West 2001). However, it is the responsibility of owners to monitor private drinking supplies.

1.5 Background of nitrate in groundwater

1.5.1 The nitrogen cycle

Nitrogen (N) is an essential element for life and exists in both inorganic and organic forms. Nitrate levels vary naturally and due to human influences. Microbial nitrification is the oxidation of ammonium through nitrite to nitrate in aerobic conditions by micro-organisms (Kendall & Doctor, 2003), and is a naturally-occurring process that typically generates nitrate levels below 3ppm. Multiple anthropogenic influences can cause nitrate levels above 3ppm in drinking water, including chemical fertilizers, septic effluent, animal manure, and road salts. Reservoirs of natural N include oceans, surface water, groundwater, soils, rock and the atmosphere, which is the largest reservoir. The

7 nitrogen cycle is the movement of N among the atmosphere, biosphere, and geosphere.

The main processes of N cycling are mineralization, nitrification, denitrification, fixation, and uptake (Van Spanning et al. 2005). Microorganisms, mainly bacteria, play a major role in transforming and cycling N. Nitrogen may undergo many transformation processes in the ecosystem before reaching the groundwater.

1.5.1.1 Fixation

N fixation occurs when atmospheric N (N2) is converted into ammonium (NH3), making it available for use by plants and animals. Fixation occurs from free living nitrogen-fixing bacteria in soil and nitrogen-fixing plants, such as legumes. Lightning also contributes to fixation with high pressure and temperature to form NO from N2 and

O2. Oxidization of NO then becomes NO2 and eventually HNO3. Anthropogenic industrial N fixation occurs due to the Haber-Bosch process, which uses pressure and high temperatures to fix N, similar to lightning. Today, about 86% of N is fixed by the

Haber-Bosch process to use as industrial fertilizers (Nieder and Benbi 2008).

1.5.1.2 Nitrification

After fixation, ammonification converts ammonia to ammonium and ammonium is converted to nitrite and then biological oxidation converts nitrite to nitrate (Rivett et al.

2008). Nitrification varies with soil moisture because the reaction requires oxygen.

Throughout the steps of nitrification, the N compounds can be taken up by plants or

+ - - become adsorbed to soil. The remaining of NH4 , NO2 , and NO3 can also be leached into groundwater.

8

1.5.1.3 Denitrification

Denitrification is a biologically stimulated reduction process that converts nitrate to nitrite or nitrate/nitrite to nitrogen gas under anaerobic conditions (Nieder and Benbi

2008). Bacteria are the main actors for nitrate reduction, while abiotic reactions instigate a minor amount. The organisms responsible for denitrification exist in surface water, soils and groundwater. Denitrification is restricted by multiple environmental factors such as the availability of oxygen and nitrogen oxides. Less significant factors include pH, soil water content, porosity, nitrate concentration, and temperature (Van Spanning et al.

2005). Due to the number of factors that are site-specific, predicting denitrification and its extent can be challenging (Rivett et al. 2008).

1.5.1.4 N leaching in groundwater

Leaching of N to groundwater occurs when N is not assimilated into plants or lost to the atmosphere (Di and Cameron 2002, Dinnes et al. 2002). The amount of nitrate leached into groundwater depends on the source flux and environmental factors including soil type, vegetation, rainfall, groundwater recharge, depth to water table and irrigation

(Dinnes et al. 2002). Septic systems are a point source of nitrogen shown to contribute to nitrate leaching into groundwater (MacQuarrie et al. 2001). Once nitrate reaches the groundwater, it migrates in the aquifer through advection and dispersion. These complex processes determine patterns of well water nitrate contamination (Almasri 2006).

1.6 The impact of ex-urban development on PPCPs and the nitrogen cycle

- Nitrate (NO3 ) is an ion that occurs naturally within the nitrogen (N) cycle. The natural N cycle can be altered when humans modify the natural terrain for urbanization or agriculture. These modifications can result in N excess, which in turn creates water

9 quality issues (Groffman & Rosi-Marshall, 2013). The intensification of agriculture at urban fringes can lead to increased point (Gardner & Vogel 2005) and nonpoint source contamination (Frumkin 2002, Barrios 2000). The location and intensity of nitrate contamination reflects natural water flow as well as complex management choices and land use changes. Thus, it is important to understand the relationship between urban growth and the spatial patterns of water quality. For instance, growth at the ex-urban fringe of mid-sized cities typically increases the use of on-site sewage treatment, which in turn increases potential groundwater and stream contamination (Wilhelm et al. 1994).

Some residents of mid-sized cities desire to live at wildland interfaces while maintaining easy access to city resources (Ewing et al. 2014). These desires often drive ex-urban growth, growth that occurs away from city centers, as well as patterns of water contamination (Barrios et al. 2000, Frumkin 2002, Young, 2009). For instance, due to the motivation for living at Wildland Urban Interfaces (WUI), spatial patterns of increased nitrate fluxes reflect ex-urban development patterns (Drake & Bauder 2005, Cadenasso et al. 2008, Claessans et al. 2009, Tu 2011). Often ex-urban development occurs before city planning in those areas, leading to challenges in long-term contaminant management from septic systems (LaGro, 1996).

Septic systems can also introduce other forms of drinking water contamination.

PPCP’s are introduced to ground and surface waters when humans or animals ingest pharmaceuticals and then excrete them into the environment, often via septic or sewer systems. PPCPs are not always removed during wastewater treatment, and can be flushed into ground and surface waters from septic systems, waste treatment plants, or leaking sewage pipes. The effects of PPCP contamination is described further in section 1.8.

10

Water resource management depends on identifying the sources of nitrate and

PPCP contamination (Kendall 1998, Curt et al. 2004, Xue et al. 2009). Source identification allows for targeted water remediation, creating more efficient, effective and less costly improvement efforts. In this study, sources of groundwater contamination are identified with a suite of isotope and chemical markers.

1.6.1 Septic systems influence on nitrate in groundwater

Septic tank systems are used to treat domestic wastewater from individual or community residences. Typically, septic tank systems comprise one or two septic tank chambers where solids settle and undergo anaerobic digestion, and a piped drainfield designed to disperse the non-solid waste water into the soil zone, where abiotic and biotic processes (such as filtration and denitrification) purify the drainfield effluent (Withers et al. 2014). The solids in the chamber should be removed every 1-3 years by a company for disposal (Withers et al. 2014). Septic tank systems can contribute to groundwater contamination through improper construction and siting, spatial density, lack of maintenance, too shallow a water table, or unsuitable site geology and soil (Yates 1985).

In populated areas of Idaho, high septic density and agricultural runoff are the major sources of nitrate contamination (West 2001).

1.7 Geochemical tracers that can determine sources of nitrate-N

1.7.1 Isotopes

15 18 - 1.7.1.1 Fingerprinting techniques using  N and  O from NO3 N

Multiple anthropogenic influences can cause elevated nitrate levels in drinking water, including chemical fertilizers, septic effluent, animal manure, and road salts.

Because there are many possible point and non-point nitrate sources, source identification

11 is difficult (Kendall 1998, Addiscott & Benjamin 2004, Curt et al. 2004, Meehan, 2005,

Powlson et al. 2008, Xue et al. 2009). Most nitrate sources can be differentiated using nitrate stable isotope compositions (Xue et al. 2009) because the physical, chemical, and biological processes of the N cycle fractionate nitrogen unequally depending on their N source (Kendall 1998). Without knowing the exact process that fractionated N, differentiating nitrate sources using only δ15N isotopic composition is too difficult, and is the reason for a dual isotope approach. A dual isotope approach differentiates nitrate

15 - 18 - 15 - sources by using both δ NNO3 and δ ONO3 compositions. The ranges of δ NNO3 and

18 - δ ONO3 values for differing sources are shown below in Figure 1.7.1. However, there are some limitations with the dual isotope approach. For instance, δ15N and δ18O values

15 - 18 - overlap for some nitrate sources. Also, denitrification can alter δ NNO3 and δ ONO3 values, complicating source identification (Michalski et al. 2004)

Figure 1.7.1 Nitrate δ15N and δ18O values with their corresponding source ranges (adapted from Kendall 1998).

12

Normally, δ17O is not used for nitrate differentiation because it has a low natural occurrence and has limited fractionation potential (Michalski et al. 2004a). However, when δ17O is used in conjunction with δ18O and δ15N, it becomes a useful tool. Using

17 - 15 - 18 - 17 - δ ONO3 in a triple isotope approach (δ NNO3 , δ O NO3 , δ ONO3 ) can identify potential microbial denitrification, unmeasured processes (i.e., soil uptake or plant utilization), and atmospheric deposition of nitrate (Michalski et al. 2004b). The triple isotope method has been successfully used to show seasonality in atmospheric nitrate, and to more precisely determine soil and aquatic atmospheric nitrate contributions. By calculating △17O= δ17O -

0.52(δ18O) (Michalski et al. 2004b, Dejwakh et al. 2011), dual isotope plots can be transformed to remove atmospheric nitrate contributions (Michalski et al. 2004b) and estimate the effect of bacterially-mediated denitrification (Dejwakh et al. 2011). The triple isotope approach also provides a better tracer of atmospheric nitrate when more complex nutrient cycling and highly variable seasonal flows are present (Michalski et al.

2004b).

Unfortunately, the triple isotope approach does not allow discrimination between sewage and manure nitrate sources (Xue et al. 2009, Nestler et al. 2011). Differentiation of sewage and manure requires combining other geochemical or other isotopic tracers

15 - 18 - with the δ NNO3 , δ O NO3 isotope approach (Bassett et al. 1995, Karr et al. 2001, Mayer et al. 2002, Min et al 2002, Spruill et al. 2002, Mitchell et al. 2003, Kaushale et al. 2006,

Choi et al. 2007, Xue et al. 2009, Nestler et al. 2011, Fenech et al. 2012). Adding a nitrate

15 - co-migrating isotope tracer, such as boron, with a dual isotope approach (δ NNO3 ,

18 - δ ONO3 ) has been attempted to decipher manure vs. septic sources (Nestler et al. 2011).

However, using boron (δ11B) still produces a manure/septic overlap between +6.39‰ -

13

+12.9‰ (Nestler et al. 2011 and Xue et al. 2009), which can be considerable in some systems. Other geochemical tracers such as pH, conductivity, chloride, nitrate-N concentrations, dissolved organic/inorganic and nitrogen levels have been successfully used with nitrate isotopes to differentiate sewage and manure sources, but they often rely on site-specific information to support source identification (Karr et al.

2001, Mayer et al. 2002, Spruill et al. 2002, Kaushale et al. 2006). For example, Min et al. (2002) and Choi et al. (2007) both identified manure as the dominant nitrate source when there was a strong correlation between increasing groundwater nitrate concentration

15 - and more positive δ NNO3 values. The distinction between manure and sewage was only possible because they were able to rule out sewage as a dominant nitrate source based on land use. Sometimes manure and sewage sources are not co-located, and sources can be distinguished by where nitrate concentrations increase (Mitchell et al. 2003). However, sewage and manure nitrate contamination sources remain difficult to distinguish using only land use types or geochemical tracers (Fenech et al. 2012).

1.7.1.2 Using 18O, 2H from well water to interpret groundwater recharge sources

Isotope compositions of 18O and 2H in groundwater can further our knowledge of nitrate contamination by identifying recharge areas as well as surface and groundwater interactions affecting contaminant transport and mixing (Fritz, 1980). Understanding groundwater recharge locations and nitrate transport mechanisms can influence the placement of drinking wells or the permitting of possible contamination sources such as agriculture or septic systems.

Groundwater compositions of δD (δ2H) and δ18O are used because water recharged from different sources, elevations, air masses, and seasonal periods carry

14 different isotopic signatures (Kendall 1998). Isotopic fractionation produces different isotopic signatures due to condensation temperature and evaporation, allowing the identification of recharge sources, such as rainfall, river water infiltration, or recharge from melting snowpacks (Fritz, 1980, Krabbenhoft et al. 1990, Wood & Sanford 1995,

Blasch & Bryson 2007). A standard method to determine proportional contributions from multiple sources is to use linear mixing models based on mass balance equations (Fritz

1980).

1.7.2 Pharmaceuticals and personal care products

Because nitrate isotopes cannot differentiate sewage and manure sources (Fenech et al. 2012), emerging contaminants such as pharmaceuticals and personal care products

(PPCPs) may be useful as chemical markers to further identify animal vs. human nitrate sources. Pharmaceuticals used by humans are ingested and passed through the body, and after the waste is excreted, it passes through a sewage or septic system where it is released into downstream waters or the soil. Pharmaceuticals used in veterinary capacities are consumed and then excreted by the animal directly onto the landscape, except when manure is contained in some feed lot settings. Pharmaceuticals used by humans or animals may or may not degrade in sewage, septic, or soil systems. Thus, the substances and/or their by-products may enter surface or groundwater bodies. Personal care products, such as DEET, may not be chemically degraded but can still pass through the same flowpaths when people bathe after the external application of these products.

Personal care products are not normally used in veterinary services (Glassmeyer et al.

2005).

15

Figure 1.7.2 Pathways for input and transport of pharmaceuticals into drinking water. This Figure shows human vs. veterinary pharmaceuticals and is a simplified transport diagram adapted from Figure 1.1 in Kummerer (2004).

Because most PPCPs are synthetic, they have low background levels in the environment

(Benotti & Brownawell 2007). They are designed to persist so that they remain active in the body when used to cure disease (Halling-Sørensen 1998, Jjemba, 2002, Jjemba, 2006,

Anderson et al. 2004). Persistence and low background levels make PPCPs ideal tracers of anthropogenic contamination. However, the biochemical fate and transport rates of

PPCPs in groundwater are still largely unknown, so the utility of different PPCPs are tracers may vary.

PPCPs include a broad array of compounds that are used in distinct settings and break down under different conditions; thus, different PPCPs are useful as indicators of different contamination sources. Some chemical tracers, such as ibuprofen, caffeine, and acetaminophen, readily break down in wastewater treatment plants (WWTPs), and have thus been used to trace raw sewage contamination (Richardson & Bowron 1985,

16

Glassmeyer et al. 2005, Kasprzyk-Hordern et al. 2009). Other compounds, such as carbamazepine, codeine, and diphenhydramine, are not removed by WWTPs, thus making them ideal for indicating the presence of treated sewage contamination in surface waters (Bouraoui et al. 2011). Some pharmaceuticals are used exclusively or primarily in veterinary settings, and some antibiotics such as tetracycline, sulfadimethoxine, and sulfadiazine have been successfully used as manure source identifiers (Batt et al. 2006,

Zhao et al. 2010, Chen et al. 2012). However, there is often an overlap between human and veterinary usage (Kasprzyk-Hordern et al. 2009).

Some PPCP substances have been shown to be ineffective tracers because of short residence times due to degradation and low usage (Fenech 2012). Only a relatively small number of PPCPs have been investigated for their ability to differentiate sewage and manure contamination. The PPCP tracers mentioned above are not necessarily the only suitable tracers; rather they are the ones that have been studied the most, and others compounds may also be useful (Richardson & Bowron 1985, Glassmeyer et al. 2005,

Batt et al. 2006, Kasprzyk-Hordern et al. 2009, Zhao et al. 2010, Chen et al. 2012).

The ability of PPCPs to act as an indicator of sewage or manure is site-specific because of social factors influencing prescription and usage rates among both human and animal populations. In Idaho, common PPCPs that have been detected include caffeine, sulfamethazine, sulfadimethoxine, sitosterol, sulfamethoxazole, carbamazepine, stigmasterol, trimethoprim, and acetaminophen (Batt et al. 2006, Schorzman & Baldwin

2009). Sulfamethazine and sulfadimethoxine are the only chemical substances from the above list that have been shown to be useful to distinguish between manure vs. septic sources (Batt et al. 2006). Other tracers may be suitable as this list is not exhaustive. A

17 more extensive nation-wide survey of PPCPs can be found in Kumar & Xagoraraki

(2010) and Focazio et al. (2008).

1.8 Health effects of nitrate and PPCP contamination

Nitrate is a contaminant of concern because it has been linked to various environmental and health effects. High nitrate in surface waters can lead to eutrophication (Kendall 1998, Monaghan et al. 2007), biodiversity loss, and ecosystem failure (Martin et al. 2012). Human health can also be negatively affected by high nitrate in drinking water (Addiscott & Benjamin 2004, Powlson et al. 2008). Consuming high nitrate levels can cause cyanosis, altered thyroid function, rapid pulse, and a build-up of protein in the liver, spleen, or kidney (Pohl & Colman, 2006). Children under the age of six months can develop nitrate-induced methemoglobinemia, blue-baby syndrome, a condition that produces too much methemoglobin, preventing oxygen uptake in the blood

(Mueller & Helsel, 1997).

Recent research has shown that exposure to low concentrations of PPCPs can also lead to negative environmental and human health effects (Ternes et al. 2004, Dufour-

Rainfray et al. 2011, Seiler et al. 1999, Lucia et al. 2010). Human health risks of environmental PPCP exposure can include the interference of hormone production from endocrine disrupters (Ternes et al. 2004), and prenatal health effects, including gene modification associated with autism (Dufour-Rainfray et al. 2011). A more detailed account of the health and environmental effects from PPCP contamination in the LPRV is covered in section 2.5.3.

18

1.9 Water quality as an ecosystem service

Ecosystem services are benefits given directly or indirectly to the human population from the ecosystems properties or habitat (MEA 2005). These services provide life’s basic needs; changes in these essential services affect human’s livelihood or income, and create political conflict (MEA 2005). Pressures from human population decrease the ability of ecosystems to provide services humans depend on for life (Boulton et al. 2008). At least thirteen different types of goods and services have been identified as supplied by groundwater stored in aquifers (Committee on Valuing Ground Water 1997); some of these goods include potable water, food product processing, as well as irrigation of agriculture, landscape, and turf. Groundwater as a service provides a disposal medium for wastes and other byproducts of human economic activity and allows improved water quality via natural water purification.

The interaction of society and the environment in growing urban infrastructures is complex; therefore, it is important to understand water quality as an urban resource

(Young, 2009; Keeler et al. 2012). In 2005, the USGS estimated that ~79.6 billion gallons per day (Bgal/day) were withdrawn from groundwater reserves in the United

States (Kenny et al. 2009). At the same time, groundwater contamination is common and a growing concern in mid-sized cities due to water quality impacts of urban infrastructure

(O’Driscoll et al. 2010). In a study of 221 United States cities, 55% of the cities showed urbanization in scattered patterns outside of metropolitan city limits (Ewing & Hamidi,

2014). Exurban growth patterns can lead to increased pollution of the natural environment and water resources (Ewing & Hamidi, 2014). Many studies have shown a positive relationship between increased nitrate concentrations at urban-exurban

19 interfaces. These nitrate concentrations could come from septic systems, agriculture, lawn care, and/or animal husbandry (Drake & Bauder 2005, Claessens et al. 2009, Tu

2011). Decreasing habitat quality can diminish soil microbial communities’ ability to metabolize nitrate, thereby allowing more nitrate contamination to enter a groundwater aquifer (Herman et al. 2001).

1.10 Thesis organization and objectives

1.10.1 Chapter 1

The first chapter of this thesis presents project motivations and general background information. This chapter gives a brief overview of previous contamination and aquifer recharge studies conducted in the LPRV as well as a summary of Idaho’s nitrate priority areas. Lastly, the chapter covers an overview of the N cycle, septic systems, and ecosystem service studies focused on water quality and groundwater.

1.10.2 Chapter 2

Chapter 2 covers the present state of spatiotemporal nitrate contamination of the LPRV aquifer and uses multiple identification methods to distinguish possible sources of contamination. I review the established uses and limitations of each identification method, including the first study for the occurrence of pharmaceuticals and personal care products (PPCPs) in the LPRV. This chapter will be developed for publication.

1.10.3 Chapter 3

The third chapter, also to be submitted for publication as a standalone paper, focuses on the public’s risk perceptions of water quality and actions that result from these perceptions. Responses from a MILES social survey of people living in the LPRV were

20 analyzed to gauge their perceptions of water as an ecosystem service, risks to that service, and personal actions they take to improve water quality.

1.10.4 Chapter 4

The final chapter summarizes my key findings and provides recommendations for possible future studies.

21

References

Abdalla, C. W. 1994. “Groundwater Values from Avoidance Cost Studies : Implications for Policy and Future Research.” American Journal of Agricultural Economics 76 (5): 1062–67. doi:10.2307/1243392.

Addiscott, T M, and N Benjamin. 2004. “Nitrate and Human Health.” Soil Use and Management 20 (2): 98–104. doi:10.1111/j.1475-2743.2004.tb00344.x.

Almasri, M. N. 2006. “Nitrate Contamination of Groundwater: A Conceptual Management Framework.” Environmental Impact Assessment Review 27 (3): 220– 42. doi:10.1016/j.eiar.2006.11.002.

Anderson, P. D., V. J. D’Aco, P. Shanzahan, S. C. Chapra, M. E. Buzby, V. L. Cunningham, B. M. Duplessie, et al. 2004. “Screening Analysis of Human Pharmaceutical Compounds in U.S. Surface Waters.” Environmental Science & Technology 38 (3): 838–49. doi:10.1021/es034430b.

Barrios, A. 2000. “Urbanization and Water Quality.” CAE Working Paper Series. WP00- 1. American Farmland Trust’s Center for Agriculture in the Environment, DeKalb, Ill, 1–22.

Bassett, R.L., P.M. Buszka, G.R. Davidson, and D. Chong-Diaz. 1995. “Identification of Groundwater Solute Sources Using Boron Isotopic Composition.” Environmental Science & Technology 29 (12): 2915–22. doi:10.1021/es00012a005.

Batt, A.L., D.D. Snow, and D.S. Aga. 2006. “Occurrence of Sulfonamide Antimicrobials in Private Water Wells in Washington County, Idaho, USA.” Chemosphere 64 (11): 1963–71. doi:10.1016/j.chemosphere.2006.01.029.

Bennett, E. M., G. D. Peterson, and L. J. Gordon. 2009. “Understanding Relationships among Multiple Ecosystem Services.” Ecology Letters 12 (12): 1394–1404. doi:10.1111/j.1461-0248.2009.01387.x.

Benotti, M. J., and B. J. Brownawell. 2007. “Distributions of Pharmaceuticals in an Urban Estuary during Both Dry and Wet Weather Conditions.” Environmental Science & Technology 41 (16). ACS Publications: 5795–5802. doi:10.1021/es0629965.

Blasch, Kyle W., and Jeannie R. Bryson. 2007. “Distinguishing Sources of Ground Water Recharge by Using 2H and 18O.” Ground Water 45 (3): 294–308. doi:10.1111/j.1745-6584.2006.00289.x.

22

Boulton, A. J., G. D. Fenwick, P. J. Hancock, and M. S. Harvey. 2008. “Biodiversity, Functional Roles and Ecosystem Services of Groundwater Invertebrates.” Invertebrate Systematics 22 (2): 103–16. doi:10.1071/IS07024.

Bouraoui, F., B. Grizzetti, and A. Aloe. 2011. “Long Term Nutrient Loads Entering European Seas.” Science of the Total Environment 409 (23): 4899–4916. doi:10.1016/j.scitotenv.2011.08.015.

Bowser, N., S. Goodell, C. Grantham, P. Jehn, G. Winter, L. Holstine, W. Haas, et al. 1996. “Protecting Ground Water Quality in Idaho.” Idaho Ground Water Quality Plan Section 39-123, Idaho Code. doi:10.1017/CBO9781107415324.004.

Brauman, K. A., G. C. Daily, T. K. Duarte, and H. A. Mooney. 2007. “The Nature and Value of Ecosystem Services: An Overview Highlighting Hydrologic Services.” The Annual Review of Environment and Resources. 32 (1): 67–98. doi:10.1146/annurev.energy.32.031306.102758.

Cadenasso, M. L., S. T A Pickett, P. M. Groffman, L. E. Band, G. S. Brush, M. F. Galvin, J. M. Grove, et al. 2008. “Exchanges across Land-Water-Scape Boundaries in Urban Systems: Strategies for Reducing Nitrate Pollution.” Annals of the New York Academy of Sciences 1134: 213–32. doi:10.1196/annals.1439.012.

CH2M-Hill. 1994. “Hydrogeology and Assessment of TCE Contamination in the Southern Portion of the Pocatello Aquifer.” Phase I Aquifer Management Plan Final Report, 1–100.

Chen, Y., H. Zhang, Y. Luo, and J. Song. 2012. “Occurrence and Dissipation of Veterinary Antibiotics in Two Typical Swine Wastewater Treatment Systems in East China.” Environmental Monitoring and Assessment 184 (4): 2205–17. doi:10.1007/s10661-011-2110-y.

Choi, W.J., G.H. Han, S.M. Lee, G.T. Lee, K.S. Yoon, S.M. Choi, and H.M. Ro. 2007. “Impact of Land-Use Types on Nitrate Concentration and δ15N in Unconfined Groundwater in Rural Areas of Korea.” Agriculture, Ecosystems & Environment 120 (2-4): 259–68. doi:10.1016/j.agee.2006.10.002.

Claessens, L., C. L. Tague, L. E. Band, P. M. Groffman, and S. T. Kenworthy. 2009. “Hydro-Ecological Linkages in Urbanizing Watersheds: An Empirical Assessment of in-Stream Nitrate Loss and Evidence of Saturation Kinetics.” Journal of Geophysical Research 114 (G4): G04016. doi:10.1029/2009JG001017.

Committee on Valuing Ground Water. Water Science and Technology Board. Commission on Geosciences, Environment, and Resources and National Research Council. 1997. Valuing Ground Water: Economic Concepts and Approaches. Washington, DC: National Academy Press.

23

Curt, M. D., P. Aguado, G. Sánchez, M. Bigeriego, and J. Fernández. 2004. “Nitrogen Isotope Ratios of Synthetic and Organic Sources of Nitrate Water Contamination in Spain.” Water, Air, and Soil Pollution 151 (1-4). Springer: 135–42.

Daily, G. C., and P. A. Matson. 2008. “Ecosystem Services: From Theory to Implementation.” Proceedings of the National Academy of Sciences 105 (28): 9455– 56. doi:10.1073/pnas.0804960105.

Dejwakh, N. R., T. Meixner, G. Michalski, and J. McIntosh. 2011. “Using 17O to Investigate Nitrate Sources and Sinks in a Semi-Arid Groundwater System.” Environmental Science & Technology 46: 745–51. doi:10.1021.

Di, H. J., and K. C. Cameron. 2002. “Nitrate Leaching in Temperate Agroecosystems: Sources, Factors and Mitigating Strategies.” Nutrient Cycling in Agroecosystems 64 (3): 237–56. doi:10.1023/A:1021471531188.

Dinnes, D. L., D. L. Karlen, D. B. Jaynes, T. C. Kaspar, and J. L. Hatfield. 2002. “Review and Interpretation: Nitrogen Management Strategies to Reduce Nitrate Leaching in Tile-Drained Midwestern Soils.” Agronomy Journal 94 (1): 153–71. doi:10.2134/agronj2002.0153.

Drake, V. M., and J. W. Bauder. 2005. “Ground Water Nitrate-Nitrogen Trends in Relation to Urban Development, Helena, Montana, 1971-2003.” Ground Water Monitoring and Remediation 25 (2): 118–30. doi:10.1111/J.1745- 6592.2005.0017.X.

Dufour-Rainfray, D., P. Vourc’h, S. Tourlet, D. Guilloteau, S. Chalon, and C. R. Andres. 2011. “Fetal Exposure to Teratogens: Evidence of Genes Involved in Autism.” Neuroscience & Biobehavioral Reviews 35 (5): 1254–65. doi:10.1016/j.neubiorev.2010.12.013.

Ewing, R., and S. Hamidi. 2014. “Measuring Sprawl 2014.” Smart Growth America: Making Neighborhoods Great Together 45 (2): 1–45. doi:10.1111/j.1728- 4465.2014.00390.x.

Fenech, C., L. Rock, K. Nolan, J. Tobin, and A. Morrissey. 2012. “The Potential for a Suite of Isotope and Chemical Markers to Differentiate Sources of Nitrate Contamination: A Review.” Water Research 46 (7): 2023–41. doi:10.1016/j.watres.2012.01.044.

Focazio, M. J., D. W. Kolpin, K. K. Barnes, E. T. Furlong, M. T. Meyer, S. D. Zaugg, L. B. Barber, and M. E. Thurman. 2008. “A National Reconnaissance for Pharmaceuticals and Other Organic Wastewater Contaminants in the United States - II) Untreated Drinking Water Sources.” Science of the Total Environment 402 (2-3): 201–16. doi:10.1016/j.scitotenv.2008.02.021.

24

Fritz, P. 1980. Handbook of Environmental Isotope Geochemistry. Edited by P. Fritz and J. C. Fontes. The Terrestrial Environment: 1-40. 1st ed. New York: Elsevier. PMCid: PMC287878.

Frumkin, H. 2002. “Urban Sprawl and Public Health.” Public Health Reports 117 (3): 201–17. doi:10.1016/S0033-3549(04)50155-3.

Gardner, K. K., and R. M. Vogel. 2005. “Predicting Ground Water Nitrate Concentration from Land Use.” Ground Water 43 (3): 343–52. doi:10.1111/j.1745- 6584.2005.0031.x.

Glassmeyer, S., E. Furlong, D. Kolpin, J. Cahill, and S. Zaugg. 2005. “Transport of Chemical and Microbial Compounds from Known Wastewater Discharges : Potential for Use as Indicators of Human Fecal Contamination.” Environmental Science & Technology 39: 5157–69. doi:10.1021/es048120k.

Groffman, P. M., and E. J. Rosi-Marshall. 2013. “The Nitrogen Cycle.” In Fundamentals of Ecosystem Science, 137–58. doi:10.1016/B978-0-08-091680-4.00007-X.

Halling-Sørensen, B., S. N. Nielsen, P. F. Lanzky, F. Ingerslev, H. C. H. Lützhøft, and S. E. Jørgensen. 1998. “Occurrence, Fate and Effects of Pharmaceutical Substances in the Environment-A Review.” Chemosphere 36 (2): 357–93. doi:0.1016/S0045- 6535(97)00354-8.

Herman, J., D. Culver, and J. Salzman. 2001. “Groundwater Ecosystems and the Service of Water Purification.” Stanford Environmental Law Journal 20: 479–95.

Idaho DEQ. 2014. “Idaho-Statistical Guidance for Determining Background Ground Water Quality and Degradation.” Water Quality Division March: 1–78.

Jjemba, P. K. 2002. “The Potential Impact of Veterinary and Human Therapeutic Agents in Manure and Biosolids on Plants Grown on Arable Land: A Review.” Agriculture, Ecosystems & Environment 93 (1): 267–78. doi:0.1016/S0167-8809(01)00350-4.

Jjemba, P. K. 2006. “Excretion and Ecotoxicity of Pharmaceutical and Personal Care Products in the Environment.” Ecotoxicology and Environmental Safety 63 (1): 113– 30. doi:0.1016/j.ecoenv.2004.11.011.

Karr, Jo. D., W. J. Showers, J. W. Gilliam, and A. S. Andres. 2001. “Tracing Nitrate Transport and Environmental Impact from Intensive Swine Farming Using Delta Nitrogen-15.” Journal of Environment Quality 30 (3): 1163–75. doi:10.2134/jeq2001.3041163x.

Kasprzyk-Hordern, B., R. M. Dinsdale, and A. J. Guwy. 2009. “Illicit Drugs and Pharmaceuticals in the Environment - Forensic Applications of Environmental Data.

25

Part 1: Estimation of the Usage of Drugs in Local Communities.” Environmental Pollution 157 (6): 1773–77. doi:10.1016/j.envpol.2009.03.017.

Kaushal, S. S., W. M. Lewis Jr., and J. H. Mccutchan Jr. 2006. “Land Use Change and Nitrogen Enrichment of a Rocky Mountain Watershed.” Ecological Applications 16 (1): 299–312. doi:10.1890/05-0134.

Keeler, B. L., S. Polasky, K. A. Brauman, K. A. Johnson, J. C. Finlay, A. O’Neill, K. Kovacs, and B. Dalzell. 2012. “Linking Water Quality and Well-Being for Improved Assessment and Valuation of Ecosystem Services.” Proceedings of the National Academy of Sciences of the United States of America 109 (45): 18619–24. doi:10.1073/pnas.1215991109.

Kendall, C., and D. H. Doctor. 2003. “Stable Isotope Applications in Hydrologic Studies.” In Treatise on Geochemistry, Volume 5, 319–64. doi:10.1016/B0-08- 043751-6/05081-7.

Kendall, C. 1998. “Tracing Nitrogen Sources and Cycling in Catchments.” In Isotope Tracers In Catchment Hydrology, edited by C. Kendall and J. J. McDonnell, 1st ed., 519–76. doi:10.1016/B978-0-444-81546-0.50023-9.

Kenny, J. F., N. L. Barber, S. S. Hutson, K. S. Linsey, J. K. Lovelace, and M. A. Maupin. 2009. “Estimated Use of Water in the United States in 2005.” U.S. Geological Survey Circular 1344: 1-44. http://pubs.usgs.gov/circ/1344/pdf/c1344.pdf.

Krabbenhoft, D. P., C. J. Bowser, M. P. Anderson, and J. W. Valley. 1990. “Estimating Groundwater Exchange with Lakes 1. The Stable Isotope Mass Balance Method.” Water Resources Research 26 (10): 2445–53. doi:10.1029/90WR01135.

Kumar, A., and I. Xagoraraki. 2010. “Pharmaceuticals, Personal Care Products and Endocrine-Disrupting Chemicals in U.S. Surface and Finished Drinking Waters: A Proposed Ranking System.” Science of the Total Environment 408 (23): 5972–89. doi:10.1016/j.scitotenv.2010.08.048.

Kummerer, K. 2004. “Pharmaceuticals in the Environment- A Brief Summary.” In Pharmaceuticals in the Environment: Sources, Fate, Effects, and Risks, 3–38. doi: 10.1007/978-3-540-74664-5

LaGro, J. A. 1996. “Designing without Nature: Unsewered Residential Development in Rural Wisconsin.” Landscape and Urban Planning 35 (1): 1–9. doi:10.1016/0169- 2046(96)00314-3.

Landon, M. K., G. N. Delin, S. C. Komor, and C. P. Regan. 2000. “Relation of Pathways and Transit Times of Recharge Water to Nitrate Concentrations Using Stable Isotopes.” Ground Water 38 (3): 381–95. doi:10.1111/j.1745-6584.2000.tb00224.x.

26

Liu, J., T. Dietz, S. R. Carpenter, M. Alberti, C. Folke, E. Moran, A. N. Pell, P. Deadman, T. Kratz, J. Lubchenco, E. Ostrom, Z. Ouyang, W. Provencher, C. L. Redman, S. H. Schneider, W. W. Taylor. 2007. “Complexity of Coupled Human and Natural Systems.” Science 317 (5844): 1513–16. doi:10.1126/science.1144004.

Lúcia, H. M. L. M. S., A. N. Araújo, A. Fachini, A. Pena, C. Delerue-Matos, and M. C. B. S. M. Montenegro. 2010. “Ecotoxicological Aspects Related to the Presence of Pharmaceuticals in the Aquatic Environment.” Journal of Hazardous Materials 175 (1-3): 45–95. doi:10.1016/j.jhazmat.2009.10.100.

MacQuarrie, K. T. B., E. A. Sudicky, and W. D. Robertson. 2001. “Numerical Simulation of a Fine-Grained Denitrification Layer for Removing Septic System Nitrate from Shallow Groundwater.” Journal of Contaminant Hydrology 52 (1): 29–55. doi:0.1016/S0169-7722(01)00152-8.

Martin, J., T. Henrichs, C. Francis, Y. Hoogeveen, P. Kasmierczyk, R. Pignatelli, and S. Speck. 2012. Environmental Indicator Report 2012: Ecosystem Resilience and Resource Efficiency in a Green Economy in Europe. Edited by J. Martin, T. Henrichs, and C. F. Seemore. Copenhagen: European Environment Agency.

Mayer, B., E. W. Boyer, C. Goodale, N. Jaworski, N. Van Breemen, R. W. Howarth, S. Seitzinger, et al. 2002. “Sources of Nitrate in Rivers Draining Sixteen Watersheds in the Northeastern U.S.: Isotopic Constraints.” Biogeochemistry 57-58: 171–97. doi:10.1023/A:1015744002496.

McMahan, J. B., K. T. Weber, and S. J. D. 2002. “Using Remotely Sensed Data in Urban Sprawl and Green Space Analyses.” Intermountain Journal of Sciences 8 (1): 30–37. http://giscenter.isu.edu/research/Techpg/lcc/pdf/greenspace.pdf.

MEA. 2005. “Ecosystems and Human Well-Being. Health Synthesis.” In Millenium Ecosystem Assessment, 1–64. Island, Washington DC: World Health Organization. ISBN: 9241563095.

Meehan, C. W. 2005. “Geochemistry of the Southern Pocatello Aquifer and Its Implications for Aquifer Recharge and Contamination Potential.” Master’s Thesis. Idaho State University, Pocatello. http://geology.isu.edu/thesis/Meehan_Christopher_2005_ISU.pdf.

Michalski, G., J.K. Böhlke, and M. Thiemens. 2004. “Long Term Atmospheric Deposition as the Source of Nitrate and Other Salts in the Atacama Desert, Chile: New Evidence from Mass-Independent Oxygen Isotopic Compositions.” Geochimica et Cosmochimica Acta 68 (20): 4023–38. doi:10.1016/j.gca.2004.04.009.

Michalski, G., T. Meixner, M. Fenn, L. Hernandez, A. Sirulnik, E. Allen, and M. Thiemens. 2004. “Tracing Atmospheric Nitrate Deposition in a Complex Semiarid

27

Ecosystem Using Δ17O.” Environmental Science and Technology 38 (7): 2175–81. doi:10.1021/es034980+.

Min, Joong-Hyuk, Seong-Taek Yun, Kangjoo Kim, Hyoung-Soo Kim, Jeongsang Hahn, and Kwang-Sik Lee. 2002. “Nitrate Contamination of Alluvial Groundwaters in the Nakdong River Basin, Korea.” Geosciences Journal 6 (1): 35–46. doi:10.1007/BF02911334.

Mitchell, R. J., R. S. Babcock, S. Gelinas, L. Nanus, and D. E. Stasney. 2003. “Nitrate Distributions and Source Identification in the Abbotsford-Sumas Aquifer, Northwestern Washington State.” Journal of Environmental Quality 32 (3): 789– 800. doi:10.2134/jeq2003.0789.

Monaghan, R. M., R. J. Wilcock, L. C. Smith, B. Tikkisetty, B. S. Thorrold, and D. Costall. 2007. “Linkages between Land Management Activities and Water Quality in an Intensively Farmed Catchment in Southern New Zealand.” Agriculture, Ecosystems and Environment 118 (1-4): 211–22. doi:10.1016/j.agee.2006.05.016.

Mueller, David S., and Dennis R. Helsel. 1996. “Nutrients in the Nation’s Waters: Too Much of a Good Thing?” U.S. Geological Survey Circular 1136: 24.

Neely, Kenneth W. 2013. “Trend Analyses for Idaho’s Nitrate Priority Areas, 2002- 2011.” Idaho Department of Water Quality: Water Information Bulletin 50 (8): 1– 54.

Nestler, A., Mi. Berglund, F. Accoe, S. Duta, D. Xue, P. Boeckx, and P. Taylor. 2011. “Isotopes for Improved Management of Nitrate Pollution in Aqueous Resources: Review of Surface Water Field Studies.” Environmental Science and Pollution Research 18 (4): 519–33. doi:10.1007/s11356-010-0422-z.

Nieder, R., and D. K. Benbi. 2008. Carbon and Nitrogen in the Terrestrial Environment 45-80. Dordrecht: Springer Netherlands. doi:10.1007/978-1-4020-8433-1.

O’Driscoll, M., S. Clinton, A. Jefferson, A. Manda, and S. McMillan. 2010. “Urbanization Effects on Watershed Hydrology and In-Stream Processes in the Southern United States.” Water 2 (3): 605–48. doi:10.3390/w2030605.

Pohl, H. and J. Colman. 2006. “Interaction Profile for :Atrazine, Deethylatrzine, Diazinon, Nitrate, and Simazine.” U.S. Department of Health and Human Services. Agency for Toxic Substances and Disease Registry, August: 1–68.

Powlson, D. S., T. M. Addiscott, N. Benjamin, K. G. Cassman, T. M. de Kok, H. van Grinsven, J. L’Hirondel, A. Avery, and C. van Kessel. 2008. “When Does Nitrate Become a Risk for Humans?” Journal of Environmental Quality 37 (2): 291–95. doi:10.2134/jeq2007.0177.

28

Reid, T. V. 1997. “Gravity-Derived Subsurface Paleomorphology and Structure of the Lower Portneuf River Valley.” Master’s Thesis. Idaho State University, Pocatello. http://geology.isu.edu/thesis/Reid_Ted_ISU_1997.pdf

Richardson, M. L., and J. M. Bowron. 1985. “The Fate of Pharmaceutical Chemicals in the Aquatic Environment.” The Journal of Pharmacy and Pharmacology 37 (1): 1– 12. doi:10.1111/j.2042-7158.1985.tb04922.x.

Rivett, M. O., S. R. Buss, P. Morgan, J. W. N. Smith, and C. D. Bemment. 2008. “Nitrate Attenuation in Groundwater: A Review of Biogeochemical Controlling Processes.” Water Research 42 (16): 4215–32. doi:10.1016/j.watres.2008.07.020.

Schorzman, K., and J. Baldwin. 2009. “Nitrate and Emerging Contaminants Evaluation of Springdale, Idaho: Cassia County Nitrate Priority Area.” Ground Water Quality Technical Report, 39: 1–50.

Seiler, R. L., S. D. Zaugg, J. M. Thomas, and D. L. Howcrof. 1999. “Caffeine and Pharmaceuticals as Indicators of Waste Water Contamination in Wells.” Groundwater 37 (3): 405–10. doi:10.1111/j.1745-6584.1999.tb01118.x.

Spruill, T. B., W. J. Showers, and S. S. Howe. 2002. “Application of Classification-Tree Methods to Identify Nitrate Sources in Ground Water.” Journal of Environment Quality 31: 1538–49. doi:10.2134/jeq2002.1538.

Ternes, T. A., A. Joss, and H. Siegrist. 2004. “Scrutinizing Pharmaceuticals and Personal Care Products in Wastewater Treatment.” Environmental Science and Technology 38 (20): 392A – 399A. doi:10.1021/es040639t.

Tu, J. 2011. “Spatially Varying Relationships between Land Use and Water Quality across an Urbanization Gradient Explored by Geographically Weighted Regression.” Applied Geography 31 (1): 376–92. doi:10.1016/j.apgeog.2010.08.001. van Spanning, R. J. M., M. J. Delgado, and D. J. Richardson. 2005. “The Nitrogen Cycle: Denitrification and Its Relationship to N2 Fixation.” In Nitrogen Fixation Origins, Applications, and Research Progress: Nitrogen Fixation in Agriculture, Forestry, Ecology, and the Environment, edited by D. Werner and W. E. Newton, 4th ed., 277–327. Dordrecht: Springer.

Welhan, J. A. 2006. “Water Balance and Pumping Capacity of the Lower Portneuf River Valley Aquifer, Bannock County, Idaho.” Idaho Geological Survey Report 06-5, no. July. Moscow: 1–36. http://www.idwr.idaho.gov/browse/waterinfo/ESPAM/model_boundary/IGS_Welha n_06-5.pdf.

Welhan, J. A., C. Meehan, and T. Reid. 1996. “The Lower Portneuf River Valley Aquifer: A Geologic/hydrologic Model and Its Implications for Wellhead Protection

29

Strategies.” In Final Report for the EPA Wellhead Protection Demonstration Project and the City of Pocatello Aquifer Geologic Characterization Project, 131.

West, D. 2001. “Nitrates in Ground Water: A Continuing Issue for Idaho Citizens Nitrates in Ground Water.” In Ground Water Quality Information Series, 1–26. Boise: Idaho Department of Environmental Quality.

Wilhelm, S. R., S. L. Schiff, and J. A Cherry. 1994. “Biogeochemical Evolution of Domestic Waste Water in Septic Systems: 1. Conceptual Model.” Ground Water 32 (6): 905–16. doi:10.1111/j.1745-6584.1994.tb00930.x.

Withers, P. J. A., P. Jordan, L. May, H. P. Jarvie, and N. E. Deal. 2014. “Do Septic Tank Systems Pose a Hidden Threat to Water Quality?” Frontiers in Ecology and the Environment 12 (2): 123–30. doi:10.1890/130131.

Wood, W. W., and W. E. Sanford. 1995. “Chemical and Isotopic Methods for Quantifying Groundwater Recharge in a Regional, Semiarid Environment.” Ground Water 33 (3): 458–68.

Xue, D., J. Botte, B. De Baets, F. Accoe, A. Nestler, P. Taylor, O. Van Cleemput, M. Berglund, and P. Boeckx. 2009. “Present Limitations and Future Prospects of Stable Isotope Methods for Nitrate Source Identification in Surface and Groundwater.” Water Research 43 (5): 1159–70. doi:10.1016/j.watres.2008.12.048.

Yates, M. V. 1985. “Septic Tank Density and Ground-Water Contamination.” Ground Water 23 (5): 586–91. doi:10.1111/j.1745-6584.1985.tb01506.x.

Young, R. F. 2009. “Interdisciplinary Foundations of Urban Ecology.” Urban Ecosystems 12 (3): 311–31. doi:10.1007/s11252-009-0095-x.

Zhao, L., Y. H. Dong, and H. Wang. 2010. “Residues of Veterinary Antibiotics in Manures from Feedlot Livestock in Eight Provinces of China.” Science of The Total Environment 408 (5): 1069–75. doi:10.1016/j.scitotenv.2009.11.014.

30

2 Chapter 2: Sources and patterns of drinking water contamination in the LPRV

Abstract

The Lower Portneuf River Valley (LPRV) aquifer is the primary source of potable water for the population of Pocatello, ID, USA. The aquifer is potentially vulnerable to contamination due to geologic factors and ex-urban growth. Here hotspots of nitrate-N groundwater contamination in the LPRV were identified and possible sources of nitrate contamination by sampling private drinking wells across the LPRV were assessed. Triple

17 , 18 15 isotopes of dissolved nitrate-N (δ ONO3 δ ONO3, and δ NNO3) were used to determine the main source of nitrate contamination and evaluate whether denitrification occurred.

Personal care products and pharmaceuticals (PPCPs) were used as chemical markers to differentiate septic and manure sources of nitrate contamination. Wells sampled in 2015 revealed two nitrate hotspots located in Mink Creek and towards lower South 5th Ave. To determine the relationship between recharge and surface contamination, probable sources

18 2 of groundwater recharge from isotopes of water, δ OH2O and δ HH2O were assessed.

Snowpack was identified as the principal source of recharge. There is the possibility of strong evaporative fractionation that would suggest the possibility of contamination at the surface as waters remain sufficiently near the surface to evaporate prior to recharge. The results of all analyses, including triple nitrate isotope signatures, PPCP tracers, nitrate concentrations, and major anions indicated septic sources of nitrate were likely or possible for about 2/3 of the wells sampled in 2015.

2.1 Introduction

- Nitrate (NO3 ) is an ion that occurs naturally within the nitrogen (N) cycle. When humans modify the natural terrain for urbanization or agriculture, the natural N cycle is 31 also altered. These modifications can result in N excess, which in turn creates water quality issues (Groffman & Rosi-Marshall, 2013). The motivation for people to live at

Wildland Urban Interfaces (WUI) results in cities growing beyond the urban fringe with spatial patterns of increased nitrate fluxes that reflect ex-urban development patterns

(Drake & Bauder 2005, Cadenasso et al. 2008, Claessans et al. 2009, Tu 2011). Often ex- urban development precedes city planning, leading to challenges in long-term contaminant management from septic systems (LaGro 1996).

Septic systems can also introduce nitrate and other forms of drinking water contamination, such as pharmaceuticals and personal care products (PPCPs). Nitrate is a contaminant of concern because it has been linked to various negative health effects.

Consuming high nitrate levels can cause blue baby syndrome, cyanosis, altered thyroid function, rapid pulse, and a build-up of protein in the liver, spleen, or kidney (Pohl &

Colman, 2006). Recent research has shown that exposure to low concentrations of PPCPs can also lead to negative health effects such as hormone interference (Ternes et al. 2004) and gene modification associated with autism (Dufour-Rainfray et al. 2011).

Multiple anthropogenic influences can cause elevated nitrate levels in drinking water, including chemical fertilizers, septic effluent, animal manure, and road salts.

Because there are many possible point and non-point nitrate sources, source identification is difficult (Kendall 1998, Addiscott & Benjamin 2004, Curt et al. 2004, Powlson et al.

15 - 18 - 17 - 2008, Xue et al. 2009). Using a triple isotope approach (δ NNO3 , δ O NO3 , δ ONO3 ), potential sources of nitrate-N can be identified and corrected for microbial denitrification, unmeasured processes (i.e., soil nitrification), and atmospheric deposition of nitrate

(Michalski et al. 2004b). Unfortunately, the triple isotope approach does not distinguish

32 between sewage vs. manure nitrate sources (Xue et al. 2009, Nestler et al. 2011).

Therefore, emerging contaminants such as pharmaceuticals and personal care products

(PPCPs) may be useful to identify nitrate sources by acting as chemical markers. PPCPs include a broad array of compounds that are used in distinct settings and break down under different conditions; thus, different PPCPs are useful as indicators of different contamination sources. Common PPCPs that have been detected include caffeine, sulfamethazine, sulfadimethoxine, sitosterol, sulfamethoxazole, carbamazepine, stigmasterol, trimethoprim, and acetaminophen (Batt et al. 2006, Schorzman & Baldwin

2009). Sulfamethazine and sulfadimethoxine have been shown to be useful manure vs. septic identifiers (Batt et al. 2006).

Water resource management depends on identifying the sources of nitrate and

PPCP contamination (Kendall 1998, Curt et al. 2004, Xue et al. 2009) and source identification allows for targeted water remediation, creating more efficient, effective, and less costly improvement efforts. In this study, the main sources of nitrate contamination were identified by comparing δ18O and δ15N compositions of dissolved nitrate, nitrate-N concentration, geochemical tracers of major anions, and chemical markers. If septic effluent is the main source of increased nitrate-N concentrations, then clustered wells with significantly higher nitrate concentrations (hotspots) will have

PPCPs detected more frequently than those without elevated nitrate concentrations.

2.2 Study Area

The cities of Chubbuck and Pocatello (-112.415062⁰, 42.830605⁰) have approximately 5,000 and 22,400 households, respectively, that rely on clean groundwater to meet their drinking needs (U.S. Census Bureau 2010). The LPRV is a northwest-

33 trending valley located in within a watershed comprising an upper basin of ~2970 km2 and a lower basin of ~ 370 km2 (Welhan et al. 1996). The primary drainage is the

Portneuf River, which originates in the headwaters of Gem Valley and Marsh Creek. The river enters the LPRV through the Portneuf Gap and flows northwest through the valley into the Snake River. The area has a semi-arid climate with most of its precipitation falling as winter snow and 75% of aquifer recharge derived from snowpack (Welhan

2006).

The geology of the LPRV aquifer contributes to the vulnerability of its water quality. The LPRV aquifer is divided into three sub-systems that differ in their geology: the northern, eastern, and southern aquifer systems (Figure 2.2.1). The groundwater in the shallow southern aquifer (~750 m thick) recharges the deeper northern aquifer system

(~1300 m thick) (Reid 1997). The northern aquifer has considerable storage potential in a thick sequence of poorly sorted silty clay-rich gravels and sands. The southern aquifer is a shallow strip aquifer made of permeable Bonneville flood gravels, which permit high linear flow velocities (Welhan et al. 1996). The southern aquifer’s high hydraulic conductivity allows contaminants to quickly move through the southern aquifer into the northern aquifer. The southern aquifer is separated from the eastern aquifer by a basalt layer from the Portneuf Basalt flow. The sediments underneath the basalt layer consist of low permeability valley fill that predates the Bonneville flood. The shallow alluvial/colluvial eastern aquitard has poorly sorted clay- rich boulder gravel and silty sand (Welhan et al. 1996). The west side of the southern aquifer is contained by

Proterozoic rocks and alluvial fan deposits.

34

Figure 2.2.1 Location and geology of the LPRV. NAIP imagery distributed by the Land Processes Distributed Active Archive Center (LP DAAC), located at USGS/EROS, Sioux Falls, SD. http://lpdaac.usgs.gov. 2.3 Methods

2.3.1 Previous sampling campaigns and data

Current drinking water conditions were assessed in private drinking wells around the Lower Portneuf River Valley in southeastern Idaho. These conditions were compared with previous sampling results from an Idaho State University (ISU) graduate thesis

(Meehan 2005), as well as data aggregated from other agencies, including the Idaho

Department of Environmental Quality (IDEQ), the City of Pocatello, and the Idaho

35

Department of Water Resources (IDWR). Within the aggregated data set, a 25-year study period (1990-2015) was chosen because laboratory analytical errors were not always reported prior to 1990, and those that were reported were significantly larger than errors reported after that date. The 25-year dataset contains 1724 records; of which, 82 have inadequate location information leaving 1642 records. One well location had nitrate-N levels of 48-111 mg/L in 2004. The four records from this well site were excluded because these extreme nitrate concentrations were likely caused by a nearby leaking sewer main (Meehan 2005).

2.3.2 Sampling methods

2.3.2.1 Field sampling plan

For the 2015 LPRV sampling campaign, targeted wells were randomly selected from four strata (Figure 2.3.1). The four strata were chosen to fill gaps in previous sampling campaigns and to sample high septic density zones and IDEQ nitrate priority areas. At least five wells in each stratum had been sampled in previous campaigns and at least five wells had not been previously sampled. A total of fifty wells were targeted in the four strata. An additional twenty-five wells were randomly selected if they had been previously sampled and had nitrate-N concentrations above 3 mg/L. The remaining twenty-five wells were randomly selected from wells not previously sampled by Meehan

(2005). Thus, 50 wells were targeted that had never been sampled in previous campaigns, as well as 50 wells for which some prior sampling information was available.

To define the “high septic density” priority strata, septic densities were calculated based on the same method used by Meehan (2005). Septic tank locations were determined by overlaying the city sewer information on the Bannock County parcel layer

36 and address database, such that if a county parcel had an assigned address, then it was assumed a residence with a potential waste source was located on the property. If this parcel was outside a 60-meter buffer of the city sewer line, the property was assumed to have a septic system, and a septic location was assigned to the parcel block center. The septic density raster was calculated using the ESRI (Environmental Systems Research

Institute) point-density tool, which calculates density by taking the number of points within a specified neighborhood radius and divides by the area of the neighborhood

(Chang 2010). Septic density was determined for each sampling location based on a 1-km radius neighborhood size. The method’s accuracy for defining septic density has not been assessed nor has the method been used to determine change in septic density over time.

Modifications to the targeted sampling plan occurred when well access was denied or a well water supply had been converted to city utilities. Whenever possible, nearby wells were sampled if initial targets were unavailable for any reason. The resulting 2015 sample locations (Figure 2.3.2) included 37% resampled wells and 63% newly sampled wells. Water supplies from each of the 100 wells were analyzed for nitrate, dissolved inorganic carbon (DIC), major anions, three stable isotopes of NO3-N, two stable isotopes of H2O, and 26 pharmaceuticals and personal care products

(Appendix A). All analytical techniques are described in section 2.3.3.

37

Figure 2.3.1 2015 well sampling plan highlighting the four targeted strata and the intended locations for resampling (orange circles) and new sampling (blue triangles). Septic density (septic/km2) is displayed in contoured colors with lowest densities uncolored and highest densities in dark orange.

38

Figure 2.3.2 Wells sampled in 2015. 2.3.2.2 Field sampling methods

The sampling method used for this research was adapted from the USGS National

Field Manual for the Collection of Water Quality Data (Wilde et al. 2006). The collection system (Figure 2.3.3) and protocols were designed to be suitable for both nitrate and

PPCP sampling, and to minimize exposure to the atmosphere while collecting samples.

All sampling materials in contact with a water sample were replaced after sampling each well to reduce cross-contamination. The tubing was replaced after each sampling site, and

Tygon tubing was selected to be chemically inert, to limit corrosion from exposure to sunlight and from chemical reactions. The bottles used for PPCP samples were 1L wide-

39 mouthed high-density polyethylene (HDPE) bottles that were EPA-certified level 1 cleaned (US-EPA 1992). The nitrate samples were collected in 500 mL bottles that were

EPA-certified level 1 cleaned, or in 500 mL wide mouthed HDPE bottles that had been leached for over 24 hours with deionized water (DIW).

Figure 2.3.3 Sampling collection system. All parts shown in red and pink were replaced for every sample. Modified after Welhan, written communication (2015).

The clean-hand (CH)/dirty-hand (DH) sampling technique adapted from the

USGS requires two or more people to efficiently complete the field sampling campaign.

Each person performs a set of designated tasks to decrease the probability of cross- contamination. Tasks completed by DH included contact with potential sources of contamination, such as calibrating the YSI unit, while the CH collected the field-filtered pharmaceutical samples, for example. The complete list of these tasks can be found in

Appendix B within the attached sampling protocols.

40

After permission had been received from the homeowner, the address of the well and well information known to the homeowners was recorded. When permission was denied, the sampling crew sampled at the closest house, if available, or the next planned sampling location. When permission was granted, the sampling crew used CH/DH to set- up instrumentation (Figure 2.3.3) on an outside water spigot not connected to water- storage tanks, holding or pressurization tanks or to any in-home filtration or water treatment device located upstream of the sampling point. If it was the first sample of the day, the YSI was calibrated by placing the calibration solution in well water for approximately 5 minutes to equilibrate the solution to groundwater temperatures. After the equipment was in place and sampling bottles were labeled, the well was pumped for approximately 10 minutes to purge the well pump, plumbing, and the well casing. The

YSI was then used to test for water stability and ensure the water column had been purged. When pH, specific electrical conductivity, and temperature met USGS criteria

(Table 2.3.1), the sampling bottles were dumped of the DIW from pre-leaching, then triple rinsed with sample water and filled, leaving no headspace at the top of the bottle.

The pumping rate was also recorded based off the time it took to fill a five-gallon bucket from the garden hose discharge.

Field Measurement Stability Criteria pH ±0.1 standard units

Temperature °C ±0.2°C

Specific electrical conductance (SC) ±3% , for SC > 100 µS/cm

(in microSiemens per centimeter at 25 °C) ±5% for SC ≤ 100µS/cm

Table 2.3.1 Stability requirements for groundwater that were met prior to sampling (adapted from USGS Book 9; Wilde et al. 2006).

41

Quality control was maintained by collecting two blanks samples, one for nitrates and one for PPCPs, at every tenth sample site. After 50 sampling sites had been completed, five PPCP duplicates were gathered at every 10th sampling site. All samples were stored on ice in coolers in the field until returning to the laboratory each day. The samples were then processed as described in the sample processing section 2.3.3.

2.3.3 Lab sample processing and analysis

2.3.3.1 Sample pre-processing

Nitrate samples were processed in the lab after returning and within six hours of collecting the first sample in the field. One sample was left in the cooler while the rest were placed in the fridge until ready to be processed. The workspace was initially covered with aluminum foil to decrease contamination risks. While wearing nitrile gloves, five 60mL narrow-mouthed HDPE bottles were labeled with a sample ID for nitrate (PVNnit.sample number), nitrate isotope (PVNiso.sample number), cation

(PVNcat.sample number), anion (PVNan.sample number), and DIC (PVNdic.sample number). These bottles were previously filled with DI water for leaching as described above, and the DI water was dumped before filtering began. The 60mL bottles were filled with filtered sample water. The sample water was filtered using a syringe with a tipped nonsterile 0.45µm nylon Whatman Puradisc filter. Before the bottles were filled, the filter was pre-conditioned by running about 1mL of sample through the unit. The 60mL bottles were also triple rinsed by placing ~5mL of filtered water in the bottle and shaking rapidly for thirty seconds; this was repeated two more times. The bottles for nitrate, isotopes, anions, and DIC were then filled with filtered sample, leaving a minimum amount of headspace for expansion, and placed in the freezer until ready for analysis.

42

2.3.3.2 Geochemistry for nitrate and groundwater analysis

Nitrate-N, DIC, and anion samples were analyzed in Soil and Watershed

Biogeochemistry Laboratory at Idaho State University (ISU). Once thawed, nitrate-N samples were run on a SmartChem 200 Discrete Analyzer Auto-Spectrophotometer

(Westco Scientific, Brookfield), against the Fisher SPEX Certiprep 1000 mg/L nitrate standard. Groundwater anions were also run on a Dionex ICS 5000 (ThermoScientific,

Waltham) with the ThermoScientific Dionex Seven Anion Standard II. Dissolved organic carbon (DOC) samples were run on a Shimadzu instrument, which also allowed DIC to be measured. Nitrate-N was analyzed for δ15N, δ18O, and △17O at Purdue University’s

Purdue Stable Isotope (PSI) lab, using a dual inlet continual flow Delta V Advantage

IRMS. Groundwater isotopes 18O and 2H were analyzed at ISU’s Stable Isotope

Laboratory in the Interdisciplinary Laboratory of Elemental and Isotopic Analysis

(ILEIA) with a TC/EA-IRMS. The groundwater isotopes were run from sample remaining in the DIC 60mL HDPE bottles, meaning the sample had been thawed, re- frozen, and thawed again before analysis. During all processing events, samples were uncapped for less than 40 minutes.

2.3.3.3 Emerging Contaminants

PPCP samples were filtered in-situ with clean nonsterile 0.7 µm Puradisc nylon filters and placed into level 1 certified bottles that had been triple rinsed with filtered sample water. After filtering, the samples were placed into 1 gallon freezer bags and stored at 4⁰C during transport to the laboratory where they were frozen. Upon completion of the sampling campaign, all PPCP samples were sent for analysis to the Indiana State

Department of Health Chemistry lab in Indianapolis, Indiana. Upon arrival, the samples

43 were thawed, weighed and then 2 ml of EDTA solution, 5 ml of methanol, and an internal standard solution were added to all samples. The samples were then drawn through a

Water HLB 6 ml, 500 mg cartridge using a vacuum manifold. The volume of sample solution extracted was calculated and eluted with 12 ml of methanol followed by 3ml of acetone. Sample volumes were reduced by evaporation under nitrogen at 40⁰C to approximately 200-300 µL. These extracts were then brought to a final volume of 1 mL with 0.1% formic acid in reagent water. The extracts were then frozen until ready for analysis. Analysis was completed on an API4000 LC/MS system using positive and negative modes. Twenty labeled internal standards were added prior to extraction and then were used in the quantitation of compounds, excluding sucralose. Sucralose was quantitated without internal standards because the standards were unavailable. A suite of

26 PPCPs were analyzed for each sample; the full results can be found in Appendix H.

2.3.4 Statistical and isotopic data analysis

2.3.4.1 Temporal trend of nitrate-N concentrations

To test for changes in nitrate concentration over time, all wells that were sampled over multiple decades were identified. Nitrate concentrations for each of these wells were compiled, and the probability distributions for samples collected in the 1990’s, 2000’s, and 2010’s were compared to one another. To test for changes in nitrate concentrations over time, two-sample Wilcoxon signed ranks test were completed; first, the distribution of nitrate concentrations from the 1990's was tested against the distribution of the 2000's and 2010's, and then the distribution from the 2000's was tested against that from the

2010's. This test assumes that the differences between time periods at each well are

44 independent of one another and that they are from the same continuous, symmetric distribution (Hollander et al. 2014).

Any trend in nitrate concentrations over time may be due to changes in nitrate inputs or changes in water inputs. For example, if flows of water increased, concentrations might decrease even if nitrate inputs are unchanged. Thus, time series of nitrate concentrations from the Caribou Acres well #ID6030005 were correlated with recorded snow water equivalent (SWE) from the Wildhorse Divide SNOTEL site.

2.3.4.2 Ordinary Kriging

Nitrate-N was interpolated from all measurements in the database to the entire

LPRV using ordinary kriging. Ordinary Kriging (OK) is a linear prediction method used to estimate unknown values at a location from other known values close to the unknown location (Chang 2010). Weighting of known points depends in part on the distance between each sampled well and the unknown location. Equation 2.3.1 represents the OK model used, where 푍 is the observed nitrate-N concentration at each si, or known sampling point used for estimation, 푍̂(푠0) is the unknown nitrate-N concentration at location 푠0, and W(푠푖) is the weight assigned to the measured value at each well location

푠푖.

Equation 2.3.1 Ordinary Kriging Model

The weighting factor is determined by solving a set of simultaneous equations such as Equation 2.3.2, where the unknown value at location 0 is estimated from, in this example, three known points (locations 1, 2, 3). The weights are solved using γ(hij), a

45 measure of the degree of spatial dependence or the semivariance between the known

th points i and j, γ(hi0) the semivariance between the i known point and the unknown point, and λ is a Lagrange multiplier that ensures a minimum estimation error (Equation 2.3.3,

Chang 2010). Once the weights are known and they sum to 1 (Equation 2.3.3), Equation

2.3.1 is used to estimate the unknown nitrate concentration at location 0.

Equation 2.3.2 Weighting factor from three unknown points

Equation 2.3.3 Nitrate concentration at unknown point

Esri's geostatistical tool package allows users to choose between multiple semivariogram functions in order to find the best fit. Because it provided the lowest RMSE among a suite of semivariograms considered in this study, a stable anisotropic semivariogram with a nugget (distance from the origin) of 0.2 was selected.

2.3.4.3 Getis-Ord Gi* hotspot analysis

The data used for hotspot analysis consisted of all samples between 1990 and

2015. A significant hot- or coldspot was distinguished by spatial clustering of nitrate concentrations that occupy the tails of the standard normal distribution, or z-score. The spatial analysis tool, Getis-Ord Gi* is a local statistics tool that was used to identify nitrate spatial patterns or hotspots at individual drinking wells in the LPRV. Getis-Ord

Gi* (Equation 2.3.4) depends on the weight (wi,j) between locations i and j and the total number of features (n) (Getis & Ord, 1992).

46

Equation 2.3.4 Getis-Ord Gi*

Equation 2.3.5 Mean nitrate-N concentration

Equation 2.3.6 Standard Deviation for all wells

Each well in the dataset received a z-score and p-value indicating a clustered or non-clustered high or low nitrate values. A well was only considered a significant hotspot when it had a high nitrate value and was also surrounded by neighboring wells with high nitrate values. The mean nitrate concentration (푋̅ in Equation 2.3.5) for all locations was

∗ compared to the standard deviation of all wells (S in Equation 2.3.6), and when a high 퐺푖 is calculated, a statistically significant z-score was assigned to the well. The neighborhood to define a hot- or coldspot was defined by a fixed distance, as suggested by Mitchell (2005), and this fixed distance was automatically generated to ensure all wells had at least one neighbor (excluding spatial outliers found on the East Bench and S.

5th Ave. where neighbors were separated by more than 718 m). Although the Wilcoxon test showed a significant increase of 0.86 NO3-N mg/L over the 25-year span, this temporal trend can be ignored in the hotspot analysis because this slight increase in nitrate concentration was far less than the spatial variation in the nitrate concentration and the concentrations of hot- and coldspots across the study area.

2.3.4.4 Well capture zone and septic density assignment

Omnidirectional and Directional Methods

47

Three methods were used to define a well’s capture zone, which was assumed to be the area where septic density would have an influence on water quality for a given well. The first method was a simple omnidirectional capture zone method in which created 1000 m buffers around each well and calculated septic density for that well based on the number of septic systems within that buffer area.

The second method assumed that flows from uphill were more likely to contribute to a well than all areas around the well. For this method, a directional capture zone was created by identifying the highest elevation point within a 250 m, 500 m, and 1000 m buffer. Once three elevations were determined for each well, a wedge-shaped capture zone was defined for each well that encompassed all three elevations points.

Least cost path analysis

The third method used is known as least cost path analysis. Similar to the directional wedge capture zone method described above, this path analysis technique assumed that surface topography was a good surrogate for defining each well’s capture zone. Path analysis relies on three rasters: a source raster, a cost raster, and a backlink cost distance raster. The source raster is the destination of a recharge path, and for this study, the recharge path destinations were represented by well locations within the

LPRV. A cost raster describes the impedance that recharge encounters while moving across the topographic surface. For this study, the impedance was represented by a reclassified slope raster. The original slope raster was created from a 1-meter bare-earth elevation model created from a LiDAR (light detection and ranging) dataset that had been filled to eliminate depressions. Cells with a higher slope were assigned a lower impedance, while cells with a lower slope were assigned a higher impedance. This cost

48 raster ensured that flows would move along the steepest slope to any given well. The third and final required raster, the backlink cost distance raster, represents the effort required for recharge to travel in a particular direction and distance. To create this raster, the initial cost for recharge to travel between nodes was calculated by averaging the reclassified slope values between the starting and ending nodes to assess path steepness.

This initial cost raster was then scaled based on distance traveled; lateral connections were scaled by a factor of 1, while diagonal paths were scaled by 1.414 to represent the longer flowpath (Chang 2008).

All 3 rasters were calculated for all possible cell connections within a 1-km radius of each well. Based on these rasters, the cumulative cost was calculated by summing the costs associated with each link that connected two cells. The path with the lowest cost was then assumed to be the most likely contributing path. To efficiently determine the least-cost path, the Dijkstra's algorithm was used to iterate through paths (Chang 2008).

Once the cost of each link was determined, the cell with the lowest cost distance was identified as the source of well recharge and that cost value was written to the output raster. This process continued until the least-cost path had reached the source or well.

After determining the least-cost path, the backlink cost distance raster was then used to create a recharge contributing area boundary using a watershed analysis. When multiple upslope least-cost paths led to one well location, these paths were merged to create a larger contributing area to that well. If there were few upslope recharge paths, then the contributing area was smaller (Chang 2010). A 100 m buffer was defined around the least cost contributing path(s) to define the capture zone assumed to contribute possible contaminants to a drinking well.

49

Septic density for each well was defined as the total number of septic systems within that well’s capture zone. Each well’s nitrate-N concentration was plotted against septic density to determine if nitrate concentration increased with increasing septic density.

2.3.4.5 Isotopes O and H as groundwater recharge identifier

δ18O and δ2H in groundwater sampled from all wells were used in the EPA’s

IsoError mixing model, and the spatial distributions of these water isotopes and nitrate concentration were analyzed together. Grouping analysis was used to assess any spatial patterns in deuterium excess and 18O. Deuterium excess (d-excess or d) was calculated as d=δD-8*δ18O, and was used to indicate the relative proportions of 18O and 2H contained in water (Lewis et al. 2013) or as an index of deviations from the global meteoric water line (GMWL). These analyses were used to assess surface recharge and possible contamination locations.

Grouping analysis (Equation 2.3.7 and Equation 2.3.8) finds the optimal number of groups by looking at the between group differences (BGD), and the within-group similarity (WGS), then selecting the grouping with the best R2, which depends on both

BGD and WGS (equation 2.3.9).

Equation 2.3.7 Between group differences

Equation 2.3.8 Within group differences

50

Equation 2.3.9 Grouping Analysis

In Equation 2.3.7 & Equation 2.3.8, ni is the number of features in group i, nc is the

푘 number of groups, nv is the number of variables used to group the features, 푉푖푗 is the value of kth variable of the jth feature in the ith group, 푉̅̅̅푘̅ is the mean value of the kth

̅̅̅푘̅ th variable, and 푉푖 is the mean value of the k variable in group i. ESRI implements these equations in the grouping analysis tool by creating a minimum spanning tree which summarizes the spatial relationships of the wells and their associated d-excess and 18O values (Jain 2010). The minimum spanning tree connects each well and the relationship between each well is given a weight that is proportional to the similarity of the connected wells. Groups are then selected that maximize both within-group similarity and between- group differences.

The grouping results were used in the EPA IsoError tool to determine proportional contribution of snowpack and rainfall to well recharge within each group based on the group mean δ18O and δ2H. Endmember snowpack signatures were averaged from Scout Mountain and Upper Gibson Jack samples collected 3/2013-3/2015 by individuals from the Finney Lab at ISU. Rainfall signatures were averaged from samples collected continuously from 3/2013 to 4/2015 at a Gibson Jack sampling location.

Rainfall samples that were indistinguishable from snowfall were excluded from the analysis. Linear mixing models such as IsoError use basic isotopic mass balance equations (Equation 2.3.10) to identify the proportions of two end-members such as

1 1 2 18 2 snowpack and rainfall (푓푎, 푓푏) for source isotope signatures δ O (δ 푎, δ 푏) and δ H (δ 푎,

51

2 1 2 δ 푏) which contribute to the observed mixtures of groundwater signatures (δ 푚, δ 푚). The two mixture equations are iteratively solved for all possible combinations of the two source proportions, where all sources (푓푎, 푓푏) must sum to 1 (Equation 2.3.11). The predicted mixtures were then compared to the observed mixture signature sampled from the well. When a predicted mixture was equal to the observed mixture signature or fell within the mass balance tolerance of ±0.1‰, the mixture combination was considered a feasible solution and kept, while unfeasible solutions were discarded (Phillips and Gregg

2003).

Equation 2.3.10 Isotopic mass balance from IsoError, assuming end member, a and b contribute to mixture m.

Equation 2.3.11 Source proportions of well recharge, assuming two end members, a and b. Once the feasible mixture solutions were chosen, the final solution was

2 determined from the solution with the lowest variability in fractional contribution, σ (푓푎) from the approximation of variances (equation 2.3.12). The first-order Taylor series

̅ approximation of variances for 푓푎 at the mean isotopic signatures of mixture M (훿 푚),

̅ ̅ ̅ ̅ ̅ source A (훿 퐴), and source B (훿 퐵) were used because 훿 푀, 훿 퐴, 훿 퐵 are assumed to be independently measured (Phillips and Gregg 2001). The square of the standard errors for

휎2 , 휎2 , the isotopic signatures from mixture M, and sources A and B are represented as 훿푀 훿퐴

휎2 and 훿퐵, respectively.

which reduces to:

52

Equation 2.3.12 Approximation of variances within IsoError, assuming two end members, a and b. 2.3.4.6 Nitrate isotopic corrections and interpretations

Groundwater nitrate isotope compositions depend on source mixing and source enrichment or depletion (Wood & Sanford 1995). Therefore, multi-tracer and multi- isotope approaches are more appropriate than dual-isotope approaches for identifying nitrate sources (Kendall et al. 2008). Dual isotope results can be misinterpreted if microbial denitrification processes affecting both oxygen and nitrogen stable isotopes are modified simultaneously, but not accounted for in the interpretation (Dejwakh et al.

2011). Triple isotopes of dissolved nitrate (δ18O, δ17O, and δ15N) can permit correction of initial dual isotope results to identify isotopically distinct signatures of septic/animal waste, applied fertilizers, and atmospherically-derived nitrate (Dejwakh et al. 2011,

18 17 Doctor 2003). Changes in δ ONO3 composition were corrected using Δ O as described in Dejwakh et al. (2011). The mass dependent fractionation of nitrate δ18O varies directly with δ17O with a factor of 0.52 (Kendall et al. 2008). The excess δ17O represented as

Δ17O (Δ17O= δ17O - 0.52 X δ18O) was used to correct for denitrification, allowing for the accurate delineation of atmospheric and other nitrate sources (Dejwakh et al. 2011). Out of the 100 wells sampled, only 52 samples from the 2015 campaign were processed for

δ18O, δ17O, and δ15N by the Purdue Stable Isotope lab (PSI). Δ17O corrections were applied and δ18O and δ15N were used in a mixing model (Figure 2.3.4) to identify the sources of nitrate. Wells with nitrate concentrations below 1 mg/L were excluded from the mixing model.

53

Figure 2.3.4 Sources of nitrate identified using corrected δ18O versus δ15N (adapted from Dejwakh et al. 2011 & Kendall 1998). 2.3.4.7 Major anions as septic tracers

- 2- Chloride (Cl ), sulfate (SO4 ), and TDS were used as conservative tracers of septic influence. Half of the wells sampled in Meehan (2005) were analyzed on a HACH

DR/2000 spectrophotometer while the other half were analyzed by Intermountain

Analytical Services, Inc. in Pocatello, ID. Both sets of analyses showed similar precision, and therefore were used to check the precision of the 2015 anion results. The anion ratios,

- 2- Cl / SO4 , were compared from wells sampled in both 1993/2004 and 2015 through a

Wilcoxon Signed Rank test to determine if the ratios had changed over time. Because

- 2- there was not a temporal change of Cl / SO4 ratio, it was assumed that a dilution correction need not be applied to the 2015 anion concentrations in order to compare the

2005 and 2015 data. To further constrain nitrate source identification, the 2015 Cl- and

2- SO4 concentrations were plotted against 2015 nitrate-N concentrations.

The 2015 sampling campaign did not include TDS; therefore, specific conductivity (SC) was used to estimate TDS. The correlation between SC and TDS based on the 2005 data was evaluated and then applied to the 2015 SC data to estimate TDS.

54

The 25th and 75th percentiles were used to define low and high TDS and nitrate-N thresholds. These thresholds bounded three key categories:

1. high TDS/high nitrate-N (>461.18 mg/L TDS and >2.85 mg/L nitrate-N)

2. high TDS/low nitrate-N (>461.18 mg/L TDS and <0.73 mg/L nitrate-N), and

3. low TDS/low nitrate-N (<337.57 mg/L TDS and <0.73 mg/L nitrate-N).

2.3.4.8 Presence of PPCPs compared to nitrate-N concentrations

Wells sampled in 2015 were grouped into two categories: wells in which PPCPs were detected and wells in which they were not. Differences in the median nitrate-N concentrations of these two groups were compared using a non-parametric Mann-

Whitney Rank Sum test. This test was used because sample population sizes differed between the two groups and neither group had a normal distribution (Neely 2013). The test assumes that both groups were randomly sampled and the sample sites were mutually independent (Drake & Bauder 2005). Rather than comparing nitrate concentrations directly, the values for all samples are ranked, and then rankings are compared (equation

2.3.13), where n1 the number of sampled wells in which PPCPs were not detected (based on the initial observation that the median nitrate value was lower for this group), n2 is number of sampled wells in which PPCPs were detected, and Ri represents sum of ranks for each sample group (Hollander et al. 2014).

Equation 2.3.13 Mann-Whitney ranked medians

55

2.4 Results

2.4.1 Nitrate-N temporal trend

Nitrate-N concentrations averaged 3.36 mg/L and ranged from 0 to 22.98 mg/L.

Eleven out of the 100 wells were higher than 10 mg N/L maximum contaminant level

(MCL) specified by the EPA for regulated public water systems. Three nonparametric

Wilcoxon test showed a significant change in median nitrate-N concentrations between paired wells tested in the 1990s, 2000s and 2015. Paired nitrate-N concentrations increased by 0.86 mg/L from the 1990’s to 2015 (sig. level = 0.036). However, median nitrate-N concentrations did not significantly differ when paired wells were compared between the 1990s and 2000s (sig. level=0.627) or when wells were compared between the 2000s and 2015 (sig. level=0.150). Table 2.4.1 displays the median nitrate-N concentrations for paired wells tested in the 1990s-2015, 1990s-2000s, and 2000s-2015.

Although Meehan (2005) observed an inverse relationship between SWE and nitrate-N

(Figure 2.4.1a) at a well in the eastern aquifer that had been sampled from the early 1980s through mid-2000s, this relationship was weak in the extended record from another well that continued to be sampled through 2015 (Figure 2.4.1b).

Paired wells tested in the 1990’s, 2000s, and 2010-2015 Year Median nitrate-N concentration (mg/L) 1990’s 1.17 2000’s 1.93 2010-2015 2.03 n 26 Paired wells tested in the 1990’s and 2000’s Year Median nitrate-N concentration 1990’s 2.51 2000’s 2.54 n 24 Paired wells tested in the 2000’s and 2010-2015 Year Median nitrate-N concentration 2000’s 2.94 2010-2015 2.86 n 20

56

Table 2.4.1 Median nitrate concentrations of paired wells sampled at varying time period

Figure 2.4.1a Nitrate-N concentrations from well DEQ-82 compared to SWE, Figure 5-2 of Meehan (2005)

Figure 2.4.1b Nitrate-N concentrations from Caribou Acres well #ID6030005 compared to a 5-year running average of snow water equivalent (SWE) from the Wildhorse Divide SNOTEL site

2.4.2 Nitrate Prediction and Hotspot Map

The Getis-Ord Gi* hotspot analysis detected two hotspots (purple hatching,

Figure 2.4.2) and one coldspot (blue hatching), all below the 5% significance level. The z-scores for the hotspots fell between +3.75 to +7.08 standard deviations from the mean, and the coldspot fell -2.66 to -4.06 standard deviations from the mean.

57

Figure 2.4.2 Predicted nitrate-N based on ordinary kriging, and hotspots (purple) and coldspot (blue) based on a Getis-Ord Gi* analysis. NAIP imagery distributed by the Land Processes Distributed Active Archive Center (LP DAAC), located at USGS/EROS, Sioux Falls, SD. http://lpdaac.usgs.gov

2.4.3 Septic Density vs. Nitrate

Septic density was weakly correlated with nitrate-N concentrations for all wells sampled in 2015 (Figure 2.4.3a). This was the case regardless of the method of determining a well’s capture zone, or whether all data were assessed or the relationship was evaluated only within the two hotspots (Figure 2.4.3b-d).

58

Figure 2.4.3(a-d) Septic density is a poor predictor of nitrate-N concentrations whether wells are assessed all together, grouped by eastern/western hotspot or if different capture zone methods are used. (a) Septic density vs. nitrate-N using least cost path analysis to define well capture zones. (b) Least cost path analysis capture zones showing only wells located in the east and west hotspots. (c) Septic density vs. nitrate-N using an omnidirectional method to define capture zone and showing only wells located in the east and west hotspots. (d) Septic density vs. nitrate-N using a directional method to define capture zone and showing only wells located in the east and west hotspots.

2.4.4 Spatial Distribution of Nitrate Sources

The results of the triple isotopes of dissolved nitrate (δ18O, δ17O, and δ15N) are

15 17 shown in Figure 2.4.4 with δ NNO3-N corrected for △ O. Using the standard mixing

15 model developed by Kendall (1998), the δ NNO3-N signatures of the 2015 sampled wells showed overlap with multiple nitrate signatures, including ammonium from fertilizer and precipitation, soil nitrogen, and manure/septic (Figure 2.4.5). After using all available methods (see sections 2.4.4-2.4.6) to identify nitrate sources, the sources for some wells remained unclear. The nitrate isotope signatures and concentrations for these wells are shown in Figure 2.4.6, demonstrating that all of their nitrate concentrations are ≤ 2.51 mg/L, below the threshold for natural nitrate variability. Diagnostic tests to assess

59 whether denitrification or mixing and/or shared isotopic signatures caused the overlapping source signatures revealed that denitrification was not the main driver of the

15 - complex δ NNO3-N source values because there was not a linear fit of ln(NO3 ) and a

- curved fit for 1/NO3 (Figure 2.4.7b and Figure 2.4.7c) (Kendall et al. 2008).

15 18 15 17 Figure 2.4.4 δ NNO3-N (‰) vs. δ ONO3-N (‰) with δ NNO3-N corrected for △ O. Point size scaled to nitrate-N concentration (mg/L), with larger points representing greater concentrations

15 18 15 17 Figure 2.4.5 δ NNO3-N (‰) vs. δ ONO3-N (‰) with δ NNO3-N corrected for △ O (adapted from Kendall 1998). Symbol shape indicates likely sources, where M/S is manure or septic and F/P is nitrate derived from fertilizer or precipitation. Blue points indicate wells in which PPCPs were detected and red points indicate wells in which no PPCPs were detected. Larger points represent higher nitrate-N concentrations.

60

Figure 2.4.6 Nitrate isotope signatures and concentrations (shown as numbers adjacent to each point) for all wells that were not identified as septic-impacted through the methods described in sections 2.4.4-2.4.6. All concentrations are below the 3 mg/L natural nitrate concentration threshold (Madison and Brunett 1985).

Figure 2.4.7 Diagnostic tests for denitrification fractionation versus nitrate mixing sources; (a) nitrate-N exhibits no 15 - systematic relationship with δ N NO3 -N. (b) The multiplicative inverse of nitrate-N concentration also exhibits no clear 15 - 15 - relationship with δ NNO -N, (c) A spline fit to the natural log of nitrate-N vs. δ N NO3 N demonstrates a weak curvilinear relationship. Figures b and c indicate there is mixing of nitrate-N sources (i.e. manure/septic sources mixed with soil N sources) rather than fractionation due to denitrification. If denitrification fractionation had occurred, Figure b would show a curved relationship and Figure c would display a linear relationship.

2.4.5 Groundwater Recharge Sources Using Isotopes 18O and 2H

Possible sources of groundwater recharge include both rain and snowmelt, which plot along the global and local meteoric water lines (GMWL and LMWL) whereas groundwater samples fall off both lines (Figure 2.4.9). Before assessing mixing ratios, groundwater samples were grouped based on d-excess values, and four groups were identified (Figure 2.4.10), which were somewhat spatially clustered in a Mink

Creek/Gibson-Jack group, a southern group, a northern group, and a landfill group

61

(Figure 2.4.10). Each group’s mean δ18O value was used in IsoError to calculate the contributions of rainfall and snowpack to groundwater recharge. Snowpack is the dominant recharge source for all groups. Groundwater wells in the Mink Creek/Gibson

Jack area as well as near the landfill have greater snowpack contributions than wells in the northern or southern aquifer groups (Table 2.4.3). The average snowpack and rainfall contributions for all wells located in the LPRV was roughly 80% and 20%, respectively.

This end-member mixing analyses using IsoError was repeated for δ 2H and the results were statistically indistinguishable from those reported here for δ18O.

Figure 2.4.8 D-excess found in groundwater sampled from wells across the LPRV. Lower d-excess indicates less evaporative fractionation or older recharge sources, whereas higher values indicate modern recharge sources or more evaporative fractionation.

62

Figure 2.4.9 Groundwater sample signatures (blue) with three possible recharge source signatures, rainfall (diamonds), snowpack (crosses), and stream (triangle). Both rainfall and snowpack signatures fall on the local meteoric water line (LMWL) and global meteoric water line (GMWL). Groundwater samples, by contrast, exhibit a trend off the LMWL.

Figure 2.4.10 Grouping analysis results in four groups of wells in the LPRV based on d-excess and δ18O signatures

63

Group δ2H δ180 Source δ2H δ18O (‰) (S.E.) (‰) (S.E.) (‰) (‰)

-129.12 (0.71) -17.65 (0.20) Landfill Snowpack -134.26 (2.83) -17.98 (0.37) Mink Creek/Gibson- -131.57 (0.73) -17.47 (0.10) Jack

Northern -128.54 (0.81) -16.87 (0.11) Rainfall -99.13 (4.4) -12.72 (0.74)

-127.39 (0.74) -16.34 (0.11) Southern

-129.26 (0.46) -16.92 (0.08) Average

Table 2.4.2 Mean (+/- s.e.) isotopic signatures of each group of wells identified in Figure 2.4.9 and mean isotopic signatures of potential recharge sources, snowpack and rainfall

Group Rainfall (%) (SE) Snowpack (%) (SE)

6 (0.077) 94 (0.077) Landfill

Mink Creek/Gibson- 8 (0.08) 92 (1.30) Jack 16 (0.95) 84 (3.9) Northern

20 (0.41) 80 (1.7) Southern

20 (0.075) 80 (3.7) Average

Table 2.4.3 Percent contributions to groundwater recharge (and the associated standard error) from snowpack and rainfall end-members based on IsoError analyses of the mean 18O values reported in Table 2.4.2.

2.4.6 Emerging Contaminants

The location of wells in which PPCPs were detected will be reported as east or

west of the Portneuf River to protect homeowner anonymity. Of 100 personal drinking

wells tested for PPCPs, 31 wells had at least 1 of the 9 different PPCPs shown in Table

2.4.4. An additional 17 PPCPs were also tested for (see full list in Appendix A), but they

64 were not detected in any wells. Wells in which PPCPs were detected also had significantly higher nitrate concentrations (Figure 2.4.11). The Mann-Whitney test showed that the median nitrate-N concentration in wells in which PPCPs were detected exceeded the median nitrate-N concentrations in wells in which no PPCPs were detected

(p-value =0.029).

PPCP Detected # of Concentration Type of Substance Found E/W Detections Range (ng/L) of the river Carbamazepine 2 13-20 Antiepileptic and W mood stabilizer N,N-Diethyl-m- 3 9.3-17 Insect repellent W toluamide (DEET) Diphenhydramine 1 155 Antihistamine W Codeine 1 10 Analgesic and cough W reliever Ibuprofen 5 6.1-16 Nonsteroidal anti- W, E inflammatory Fluoxetine 1 152 Selective serotonin W reuptake inhibitor (SSRI) Sucralose 18 29-656 Non-nutritive W, E sweetener Sulfadimethoxine 2 2.3-27 Antibiotic used by W domesticated animals Sulfamethoxazole 14 14-256 Antibiotic used by W, E humans Table 2.4.4 Nine pharmaceuticals and personal care products were detected in the 100 private LPRV wells tested in 2015. Some PPCPs were found in only a single well whereas others were detected more frequently. If a substance was detected in 5 or more wells, it was found in wells on both sides of the river. PPCPs that were only infrequently detected were found exclusively west of the river. Concentrations consistently remained <1ppm across all PPCPs.

65

Figure 2.4.11 Higher nitrate-N concentrations were found in wells where PPCPs were detected. 2.4.6.1 Major anions

Chloride/sulfate ratios from wells sampled in 2015 did not differ from Cl/SO4 ratios from those same wells that had been sampled during previous campaigns

(Wilcoxon Signed Rank test, significance level=0.064). The Mann-Whitney test discerned a difference between anion concentrations in all wells sampled in 2015 and all wells sampled in previous years. Spatial variability in any given year exceeded the difference in anion concentrations detected between years.

Anion concentrations increased with increasing nitrate-N, with variability partially explained by sources from the 2015 nitrate isotopic analysis (Figure 2.4.12).

Wells in which combined manure/septic (M/S) and soil N sources were identified (square symbols) had higher chloride and sulfate concentrations (Figure 2.4.12) and nitrate concentrations (Figure 2.4.13 and Figure 2.4.14). High chloride, coupled with sulfate and

TDS were assumed to indicate water softener influence. When high nitrate and water softener signatures coincided, it was assumed the well was septic-impacted, since the combined signatures represent a residential impact. There were two wells with high

66 chloride/anion (mg/L) and high nitrate that were not previously identified as septic- impacted. Because TDS and specific conductivity (SC) are strongly correlated (Figure

2.4.15), TDS for wells in the 2015 campaign could be calculated. An additional 8 wells were classified as septic-impacted because they had a combination of high TDS (mg/L) and high nitrate-N (mg/L).

Figure 2.4.12 Chloride and sulfate concentrations for wells sampled in the 2015 campaign. Different shapes reflect identified nitrate-N sources (Figure 2.4.4) and colors reflect whether PPCPs were detected. Wells with combined manure/septic (M/S) and soil N sources (squares) tend to have higher chloride and sulfate concentrations. Wells with combined sources that include all three potential nitrate-N sources, including ammonium (triangles) have low to middle ranges of sulfate and chloride.

67

Figure 2.4.13 Chloride and nitrate-N concentrations are not strongly correlated. Identified nitrate-N sources are shown as indicated in Figure 2.4.12. High Cl (>100 ppm), low nitrate (<10ppm) well waters include mixed soil N and manure or septic sources (including squares and triangles) as well as ammonium from fertilizer or precipitation (including crosses and triangles). Possible contributions from manure and septic sources (including diamonds, squares, and triangles) are found across the range of both Cl and nitrate-N concentrations.

Figure 2.4.14 As with Figure 2.4.12, but for the correlation between sulfate and nitrate-N. Similarly, septic and manure sources are found across the range of sulfate and nitrate concentrations.

68

TDS = 43.52 + 0.5644*SC R2: 0.899

Figure 2.4.15 Total dissolved solids (TDS) are strongly correlated with specific conductivity (SC) as reported for the 2004 well samples (Meehan 2005). This relationship was used to calculate 2015 TDS from measured SC to assess potential locations of denitrification (see Figure 2.4.16).

High TDS/High NO3

High TDS/Low NO3

Low TDS/Low NO3

Figure 2.4.16 Spatial distribution of wells sampled in 2015 with high TDS/high nitrate, high TDS/low nitrate, and low TDS/low nitrate well samples. Wells with values of nitrate or TDS that fell within the interquartile range of either distribution are not shown. A single well exhibits high TDS/low nitrate concentrations (blue square).

69

2.5 Discussion

2.5.1 Nitrate across the U.S.

In the LPRV, nitrate was observed at levels below the detection limit of 0.005 mg/L and up to 26 mg/L, which reflects both natural and human-caused sources.

Naturally-occurring nitrate levels of less than 3 mg/L are often due to microbial nitrification, or the oxidation of ammonium through nitrite to nitrate under aerobic conditions by micro-organisms (Kendall & Doctor, 2003). Multiple anthropogenic influences can cause nitrate levels above 3 mg/L in drinking water, including contributions from chemical fertilizers, septic effluent, animal manure, and road salts.

Nitrate is the most common chemical contaminant in aquifers across the globe and the levels of contamination are increasing (Spalding and Exner 1993). In a 2002 study of

1242 private drinking wells across the contiguous U.S., 136 wells or 11% exceeded the nitrate maximum contaminant level (MCL) of 10 mg/L set by the EPA (Squillace et al.

2002). This is similar to this 2015 study in which 11 of 100 sampled wells exceed the

MCL. Across the U.S., average nitrate-N concentrations vary between <0.3 mg/L to 3 mg/L for rural domestic wells and community water systems (Spalding and Exner 1993).

Shallow groundwater beneath agriculture areas across the U.S. had median concentrations of 3.1 mg/L, while concentrations beneath urban lands had median concentrations 1.4 mg/L (Burow et al. 2010).

When drinking water contains nitrate concentrations above the MCL, there is the potential for methemoglobinemia (Spalding and Exner 1993) and spontaneous abortions

(CDCP 1996). Nitrate concentrations above 4 mg/L have been linked to non-Hodgkin’s lymphoma (Ward et al. 1996) and studies have shown a positive trend between municipal

70 water supplies with nitrate and a risk for bladder, ovarian, and thyroid cancers (Weyer et al. 2001, Ward et al. 2010). From an ecological standpoint, nitrate is a concern because it is often a limiting nutrient in surface waters, which are sustained by groundwater, especially during low flow periods. Excess nitrate can cause structural changes in vegetation, loss of plant diversity (Verhoeven et al. 2006), and nitrate toxicity in aquatic animals. Nitrate toxicity transforms oxygen-carrying pigments into forms incapable of carrying oxygen and these effects increase with increasing nitrate concentration and exposure time (Camargo et al. 2005).

2.5.2 Nitrate sources and patterns

Nitrate in the LPRV aquifer derives from three sources according to the triple isotope approach, including: manure/septic, ammonium nitrification from fertilizer or precipitation, and soil nitrogen. However, these nitrate sources have overlapping isotopic signature ranges (Figure 2.4.5), which complicates source identification. For instance, wells with 15N values between +2 to 9‰ may derive from either a soil or manure/septic source or a combination of the two. Many of the samples indicate a potential mixture of two or even three sources of nitrate. Nitrate-N from two wells has a definitive manure/septic sources, and another two wells have a definitive fertilizer/precipitation source. Another 48 wells have mixed nitrate signatures (Figure 2.4.5) and the remaining

48 wells were not run for nitrate isotopes.

Fractionation from denitrification is not expected to impede source identification

18 15 in this study because a 1:2 relationship was not observed between δ ONO3 and δ NNO3

17 (Figure 2.4.4 and Kendall et al. 2008) and because δ ONO3 was used to correct enrichment caused by denitrification. The process of denitrification would cause

71 fractionation of both O and N isotopes of dissolved nitrate so that groundwater samples

18 15 would plot along a positively sloped fractionation line when δ ONO3 and δ NNO3were compared (Brown et al. 2001). By excluding denitrification, we focus on discerning relative contributions of soil nitrogen and manure/septic.

Land use and nitrate-N concentrations could help decipher more probable sources of the 48 mixed source wells. In areas of increased nitrate concentrations (hotspots), present and past land uses can give further clues to the dominant nitrate sources. There

15 are two possible legacy sources that could explain nitrate concentrations and δ NNO3 variability. The Mink Creek area hosted a Civilian Conservation Corps (CCC) side camp starting in December of 1935 (Pocatello Chieftain 1936) and Frazier’s poultry farm; both are possible sources of nitrate in the groundwater. The CCC camp did not have disposal facilities, instead they used latrines (Hanson 1973). If the latrine waste has not yet flushed through the area, then there could a pool of nitrate still contributing to the groundwater system. Frazier’s poultry farm could also be a potential past or present nitrate source from poultry manure. The farm was closed in 1998 and the conditions of waste are unknown. In addition, septic systems are used in the present-day Mink Creek watershed and some areas of the east bench and South 5th Ave. Current and past land use types suggest manure/septic may be an important nitrate source.

In the residential areas lining the western edge of the main aquifer from the

15 landfill area to the northwest, the δ NNO3 values are consistent with manure/septic, soil

15 N, and fertilizer sources. Two wells in the area have very light δ NNO3 values, most likely corresponding with fertilizers used by nearby agriculture and a golf course. Wells on the eastern edge of the main aquifer have heavier N signatures consistent with soil or

72 manure/septic systems. The area historically has had a high septic density, and is believed to have low rates of water recharge (Welhan 2006), suggesting the nitrate hotspot in the

South 5th area may be a mix of historical and current contamination.

There was no correlation between septic density and nitrate regardless of whether or not wells were assessed as one large group, or were analyzed only for hotspot locations, and irrespective of which of the three methods were used to define well capture zones (Figure 2.4.2 and Figure 2.4.3a-d). This lack of correlation could be because septic is not the only source of nitrate-N in drinking wells or could be because we did not did not accurately estimate each well’s capture zone, or the true septic influence in this area.

To improve this method, it would be useful to review actual septic permitting rather than inferring septic locations from zoning and distance from the city sewer line. A groundwater model that does not necessarily rely on inferring groundwater flow direction from surface topography would also help refine well capture zones.

2.5.3 Influences of groundwater recharge on nitrate

The small but significant increase in nitrate concentrations may be due to decreased recharge rather than an increase in nitrate loading. Past work showed that groundwater nitrate-N concentrations may be diluted by increased groundwater recharge during wet periods, as shown by an inverse relationship between snow water equivalent

(SWE) and nitrate concentration (Figure 2.4.1a). The well used in Meehan’s (2005) analysis was located in the eastern aquifer and was not resampled in 2015. Instead nitrate concentration from an intensively sampled Caribou Acres well in Mink Creek with known high groundwater flow velocities were compared to SWE records (Figure 2.4.1b).

A dilution effect may still be suggest by the weak increase relationship between nitrate-N

73 and SWE measured at the Wildhorse Divide SNOTEL site above Mink Creek, but the effect is weaker than observed by Meehan (2005) in the slow moving groundwater of the eastern aquifer. Changes in groundwater recharge would be expected to affect wells sourced by modern waters the most, and so groundwater isotopes were examined to assess which wells might be most susceptible to recharge variations.

Both snowpack and rainfall contribute to groundwater in the LPRV, and the

18 2 distribution of δ OH2O and δ HH2O reflects the distribution of water recharge sources, which may influence the spatial pattern of nitrate-N concentrations. The composition of

18 2 δ OH2O and δ HH2O in groundwater depends on recharge sources and evaporative processes (Gonfiantini et al. 1998). Groundwater sources based on d-excess show some spatial clustering across the LPRV. Weak grouping is most likely based on elevation

(Figure 2.4.8): groundwater samples in the main aquifer and other low-elevation sites generally have lower values of d-excess (green and blue points in Figure 2.4.8), suggesting older recharge, a greater influence due to evaporation, or the input of upper- basin ground water. One well higher in elevation on the east bench shares a similar signature to the main aquifer. Some wells clustered in Mink Creek and Gibson-Jack have larger d-excess values (red and orange points in Figure 2.4.8), indicating modern recharge and/or less evaporative fractionation. Two lower-elevation areas (boxes in

Figure 2.4.8) have wells in close proximity to one another with signatures consistent with both modern and old recharge and/or varying degrees of evaporative fractionation. Well samples with intermediate evaporative signatures between 5.35-7.92 ‰ (shown in yellow in Figure 2.4.8) appear throughout the LPRV.

74

Wells that reflect lower snowpack contributions or more evaporative fractionation may be more vulnerable to contamination because those waters likely spend more time near the surface, near possible contamination sources. Welhan (2006) assumed that snowpack was the principal recharge source to the LPRV and that rainfall contributed negligible amounts to recharge. Thus, based on his study, groundwater δ18O values closer to -18 ‰ would be expected.

The 2015 sampling results show that snowpack is the main recharge source, which is consistent with Welhan (2006). However, they also show that rainfall has the ability to contribute more recharge than previously believed, as suggested by the results for the northern and the southern group (Table 2.4.3). Groundwater with more snowpack or more rainfall would be expected to vary along the global and local meteoric water lines (GMWL and LMWL; Figure 2.4.9). The average isotopic signature of rainfall falls toward the upper right-hand corner of the graph, due to recharge that occurred during warmer months of the year, and the average snowpack signature falls toward the lower left-hand corner of the plot due to recharge that was primarily from snow that accumulated during colder months of the year. The fact that most groundwater isotope signatures do not fall on the LMWL and GMWL (Figure 2.4.9), could be interpreted in multiple ways: 1) evaporation occurred prior to recharge; either before rainfall reached the surface, during run-off and subsequent infiltration, and/or during soil-zone infiltration, 2) the presence of another recharge source that has not been recognized, and/or 3) isotopic fractionation occurred during processing (thawing and refreezing) of samples. We can rule out the latter because there were no signs of defective containers after thawing, and samples were uncapped for <40 minutes in all cases. An additional

75 recharge source such as a geothermal source could be responsible for the groundwater isotope trend. In the Boise, Idaho area, there are geothermal groundwater sources with groundwater ages between modern to 30 ka (Schlegel et al. 2009) and similar sources may be present in the LPRV (Welhan and Dombrowski 2014). Finally, groundwater with a more evaporative signature would imply that there is more time for near-surface interactions prior to recharge than previously recognized, which would increase the risk of contamination from human and animal sources. This may also partially explain why wells in South 5th Ave. hotspots (Figure 2.4.2) have significantly higher nitrate-N concentrations than elsewhere in the LPRV, because such a dramatic shift due to evaporation of local precipitation and surface ponded water might be more likely to occur in this low-elevation area of the valley. If the groundwater trend truly reflects an evaporitic shift, this could imply that anthropogenic influences are changing the groundwater system on a larger scale than previously believed, such as via domestic irrigation which pumps groundwater to the surface where it partially evaporates before re-entering the local groundwater system.

2.5.4 Factors affecting transformation of nitrate during movement through hydrologic

systems

Nitrate concentrations depend on input from potential nitrate sources and contaminant movement and reactions within a groundwater system. When little vertical dispersion is present within an aquifer, nitrate concentrations can be expected to decrease with groundwater age and with depth below the ground surface, because of less exposure to contaminant sources in deeper flow systems and through processes such as dilution and denitrification in shallow flow systems (Bohlke and Denver 1995).

76

Denitrification can affect nitrate concentrations and inferred sources in multiple ways. For instance, denitrification alters nitrate by the reduction of N, which enriches both δ18O and δ 15N values of the residual dissolved nitrate as it lowers the nitrate concentration. In systems with abundant biologically mobile nitrogen, however, denitrification can have little measurable effect on the 15N/14N ratio of residual nitrate depending on the size of the nitrate pool even while δ18O of dissolved nitrate changes. In such a situation, a sample’s 15N/14N ratio may not appear to have experienced N- fractionation when in fact nitrate alteration did occur (Curtis et al. 2011). The broader implication is that a N-fractionation model may misidentify the presence of denitrification. In the future, soil samples could establish the magnitudes of the available

N pools in the LPRV. Even without rapid cycling, denitrification alters nitrate by the reduction of N, resulting in the removal of nitrate from a system, which enriches both

15 18 δ N and δ O values of dissolved nitrate and changes the original nitrate concentration.

To calculate the degree of denitrification, and then correct the nitrate concentration to reflect what it was before denitrification, one can sample “excess air”, atmospheric gas that has been trapped in soil pores, and “excess N2”, N2 originating from denitrification and remains in the recharge solution (Green et al. 2005).

Denitrification can vary spatially across an aquifer. After accounting for groundwater age, denitrification was observed to occur uniformly in the aquifer of the

San Joaquin River watershed. By contrast, aquifer denitrification varied spatially within the Elkhorn River watershed, Nebraska (Green et al. 2005). The LPRV aquifer would be expected to follow spatial denitrification trends similar to the Yakima River watershed in

Washington, because similar to the LPRV, denitrification in the Yakima River watershed

77 is limited by aerobic conditions. This watershed has places with active denitrification that were disconnected and irregular (Green et al. 2005). The LPRV would also most likely exhibit spatially irregular zones of active denitrification, with much of the denitrification possibly occurring in the eastern aquifer due to its relatively low hydraulic conductivity.

2.5.5 Emerging contaminants

Nitrate sources of manure/septic cannot be differentiated through analysis of isotopes, however emerging contaminants such as PPCPs can act as chemical markers to identify these sources and help understand nitrate patterns in the LPRV. Because pharmaceuticals used by humans are released from septic systems into soil and groundwater, when they are detected in groundwater, the pharmaceuticals act as tracers for contamination sources. PPCPs include many compounds that are used in distinct settings and degrade under different conditions; thus, when human versus husbandry pharmaceuticals are detected in groundwater they give a better understanding of human impacts on the LPRV. The following describes the characteristics of the PPCPs that were detected from drinking wells sampled in the 2015 sampling campaign.

2.5.5.1 Carbamazepine

Carbamazepine is an antiepileptic that has been found to survive intact after 8-10 years of travel through either an anoxic saturated or aerobic unsaturated subsurface

(Drewes et al. 2003, Sui et al. 2015). According to Sui et al. (2015), carbamazepine is commonly detected in groundwater, most likely due to a lack of degradation or adsorption. Previous studies have shown carbamazepine concentrations found in monitoring and private wells typically vary between 2-75 ng/L (Drewes 2003, Osenbruck et al. 2007, Musolff et al. 2009, & Sui et al. 2015), with one concentration of 420 ng/L

78 found in Helena, Montana (Miller and Meek 2006). The well concentrations detected in the LPRV are 13 and 20 ng/L, or below the detection limit of 1.0 ng/L. These concentrations are below the guidance limit of 40 µg/L for drinking water suggested by the Minnesota Department of Health (MDH) (Minnesota Division of Environmental

Health 2013). The MDS set this limit due to laboratory studies showing health effects of decreased white blood cell counts (Basta-Kaim et al. 2008), sex hormone disturbances

(Isojarvi et al. 2004), decreased bone density in children (Chou et al. 2006), and thyroid hormone disturbances (Benedetti et al. 2005). The health effects of long-term low concentration exposure of carbamazepine have not been studied enough (Stackelberg et al. 2004, Minnesota Division of Environmental Health 2013) to set federal maximum exposure regulations.

In ecological systems studies, carbamazepine has been shown to accumulate in plant bodies (Bell et al. 2011). For instance, large concentrations of carbamazepine have been found accumulated in root tissues and can travel into aboveground tissues, such as beans (Bell et al. 2011). Pharmaceuticals that accumulate in plant tissues may later be consumed by animals and then transferred through food chains, impacting the whole food web (Bell et al. 2011). Carbamazepine has been shown to survive common wastewater treatment systems (chlorine oxidation) and therefore may have a long lasting imprint on ecosystems (Fenech et al. 2012).

2.5.5.2 DEET

DEET is the active ingredient in insect repellents. DEET was detected in 3 LPRV wells with concentrations that ranged between 9.3-17 ng/L. Typical concentrations of

DEET in water samples range between 2.2-1010 ng/L (Stackelberg et al. 2004 & Sui et

79 al. 2015). Results from clinical reports determined a daily intake of 0.082 µg/kg/day was determined safe for humans (Blanset et al. 2007). The residence times for DEET have not been established and therefore need further study (Sui et al. 2015). However, both carbamazepine and DEET survived removal methods that included carbon filtering, chlorine oxidation, ozone and ozone/hydrogen peroxide oxidation, membranes, biological processes, and magnetic ion exchange (Snyder et al. 2007).

Health effects in animals have been well-studied, and have shown that oral exposure ranging above 100 mg/kg-day can result in tremors, convulsions, abnormal movement, decreases in weight, decreased activity, and a delayed response to heat

(Blanset et al. 2007). Humans in clinical case reports have shown seizures, coma, encephalopathy, motor disturbances, and even death when 50mL of DEET was ingested

(Blanset et al. 2007).

2.5.5.3 Diphenhydramine

Diphenhydramine is an antihistamine that is not often found in groundwater, but that may be because it is not yet commonly tested for. The detection of diphenhydramine has been found at concentrations ranging between 23 ng/L and 148 ng/L (Focazio et al.

2008, Collier 2007). Within the LPRV, one instance of diphenhydramine was detected at a concentration level of 155 ng/L, above previous studies (Collier 2007).

Diphenhydramine is a suggested septic tracer because it is only used by humans and because it is not removed by wastewater treatment (Fenech et al. 2012). Due to its low prevalence in groundwater studies, health effects from diphenhydramine at low continuous concentrations have yet to be addressed. However, concentrations of diphenhydramine from 10 to 100 ng/L have been found in aquatic ecosystems (Ferrer et

80 al. 2004, Focazio et al. 2008, Bell et al. 2011), and have caused observable behavioral effects on aquatic populations in experimental settings when concentrations were at 560 ng/L (Berninger et al. 2011). Rosi-Marshall et al. (2013) concluded that diphenhydramine exposure altered bacterial community composition, specifically increasing Pseudomonas and decreasing Flavobacterium. These results suggest a need for further monitoring of diphenhydramine in the natural environment, especially since the substance is water soluble and survives waste water treatment (Berninger et al. 2011, Rosi-Marshall et al.

2013).

2.5.5.4 Codeine

Codeine is an opiate that was detected once in the LPRV at a concentration of 10 ng/L. In previous studies, codeine has been detected in wastewater treatment influent/effluent and surface waters at concentrations from 5.7 to 120 ng/L (Boleda et al.

2009, Rosi-Marshall et al. 2014). Codeine may convert to its derivative, norcodeine, decreasing detectability (Boleda et al. 2009). However, codeine can survive intact after waste water treatment (Kasprzyk-Hordern et al. 2009). Opiates and illicit drugs are an understudied group of PPCPs, and their ecological and human health effects at low concentration in water supplies are unknown (Rosi-Marshall et al. 2014).

2.5.5.5 Ibuprofen

Ibuprofen is a common anti-inflammatory detected in groundwater. It is commonly detected due to its wide use and non-prescriptive nature (Sui et al. 2015). In the LPRV, there were 5 drinking wells with detectable ibuprofen concentrations that ranged between 6.1-16 ng/L. Concentrations of ibuprofen in past studies extend from 120 to 270 ng/L (Collier et al. 2007, Boleda et al. 2009, Sui et al. 2015). Anti-inflammatories

81 including ibuprofen were found to be readily removed during groundwater recharge with retention times shorter than six months (Drewes et al. 2003). It may be possible that certain areas of the LPRV have extremely fast and direct well recharge routes, while other routes of well recharge allow ibuprofen to be removed.

When ingested by pregnant women in the third trimester, ibuprofen dosages of 2 mg/L have been linked to closure of the blood vessels and inability to pump blood through the heart in fetuses (Collier 2007).

2.5.5.6 Fluoxetine

Fluoxetine is a selective serotonin reuptake inhibitor (SSRI) and is one of the most widely used synthetic antidepressants (Santos et al. 2010). Within the samples of the LPRV, there was only one well in which fluoxetine was detected at a concentration of

152 ng/L. The negative effects of fluoxetine in ecological systems are relatively well researched in experimental settings. The presence of fluoxetine at concentrations between

100-500 ng/L induced spawning and oocyte maturation of invasive zebra mussels, reduced the reproduction of freshwater mudsnails, and increased developmental abnormalities in Japanese medaka without causing changes in reproduction success rates

(Fong 1998, Brooks et al. 2003, Foran et al. 2004, Nentwing 2007). Safe concentrations have not been established because studies rarely have detected contamination above detection limits (Buszka et al. 2009, Bell et al. 2011, de Jongh et al. 2012), and the low detection limit (8.3 ng/L) in this study revealed only one well with detectable levels of the SSRI.

There is some research that suggests little concern for human health risks from fluoxetine, however this may be due to a lack of prevalence in groundwater and therefore

82 decreased risk of exposure (Bell et al. 2011). De Moel et al. (2006) found that when advanced treatment of drinking water includes oxidative and active carbon treatment, fluoxetine was removed.

2.5.5.7 Sucralose

Sucralose is a common sweetener sold as SPLENDA by McNeil Nutritionals and is the most prevalent PPCP detected in the LPRV. It was found in 18 wells with concentrations ranging from 29-656 ng/L. Residents would have to consume 192,200 liters of drinking water at the highest concentration of sucralose found in the LPRV, to consume a 12 oz. can of diet soda. This artificial sweetener is commonly found in septic system plumes. When compared to other artificial sweeteners, sucralose is more readily broken down (Sempvoort et al. 2011 & Robertson et al. 2013). Sucralose hydrolysis can occur at acidic pH levels and elevated temperatures; however, when pH is near neutral, as it is in the LPRV, hydrolysis is not detectable (Soh et al. 2011). Microbial degradation can degrade 45% of sucralose in a four-week time span (Labare et al. 1994). Thus, concentrations of sucralose across the LPRV may vary based microbial degradation rather than the original input concentrations. The health effects of long-term low dose sucralose exposure are unknown (Soh et al. 2011).

2.5.5.8 Sulfamethoxazole

Sulfamethoxazole is an antibiotic frequently prescribed to humans and the second most prevalent PPCP detected in the LPRV. Out of the 14 wells with detectable levels of sulfamethoxazole, the maximum concentration in the LPRV was 256 ng/L, which exceeds the previously reported maximum concentration of 131 ng/L (Zimmerman 2005,

Benotti et al. 2009, Focazio et al. 2008, Illinois EPA 2008, Tabe 2010, Schaider et al.

83

2014). Residents would have to consume 314,000 liters of drinking water at the highest concentration of sulfamethoxazole found in the LPRV, to consume a normal dose of 80 mg. Microbial degradation of sulfamethoxazole is low (Barber et al. 2009) and it has been found to alter microbial communities (Haack et al. 2012). The health concerns associated with sulfamethoxazole are not well established, but exposure could create problems for people who are allergic to it (Schaider et al. 2014).

2.5.5.9 Sulfadimethoxine

Sulfadimethoxine is an antibiotic commonly prescribed to domesticated animals, including farm animals and household pets. The detection of this drug indicates contamination by animal sources because it is not approved for human usage (Batt et al.

2006). In the LPRV, this pharmaceutical was detected at concentrations of 2.3 and 27 ng/L in two wells on the west side of the Portneuf River. Residents would have to consume 37,000,000 liters of drinking water at the highest concentration of sulfadimethoxine found in the LPRV, to consume one dose of sulfadimethoxine at 100 mg. The poultry industry commonly uses sulfadimethoxine to maintain healthy animals

(Srivastava et al. 2009). However, the wells with detectable sulfadimethoxine are not located near Frazier’s poultry farm. Sulfadimethoxine will ionize and increase in mobility if pH values are near neutral as they are in the LPRV (6.9-7.5) (Batt et al. 2006).

Increased mobility may explain why wells near Frazier’s farm did not have detectable levels of this pharmaceutical. It is also possible that sulfadimethoxine was not an antibiotic used at Frazier’s poultry farm.

Antibacterial resistance is the major concern for environmental sulfadimethoxine exposure. Antibiotics used in agriculture could lead to bacterial resistance to antibiotics

84 used in human medicine (Lathers 2002). Human exposure to sulfadimethoxine has not been studied and the drug is not approved by the FDA for human consumption.

2.5.5.10 Multiple PPCPs Detected

Some of the well sites have a mixture of detected PPCPs, creating separate health risk considerations. For instance, sulfamethoxazole has been shown to inhibit metabolism of other drugs such as anticonvulsant medications. This could pose a problem for users of three of the wells that contain both sulfamethoxazole and carbamazepine (Schaider et al.

2014). Although caffeine is one of the most frequent PPCPs found in wastewater or groundwater, it was not detected in the LPRV groundwater system (Schaider et al. 2014).

Some of the wells without any detectable PPCPs could reflect low loading into the residential septic effluent or environmental removal of the PPCP like caffeine, rather than a complete absence in the LPRV (Schaider et al. 2014, Reddy et al. 2005, Celiz et al.

2009). Environmental treatment of septic effluent includes soil absorption and aerobic sand filters which can result in fewer contaminants reaching the groundwater. For instance, >99.9% of caffeine is shown to be removed by soil absorption (Godfrey et al.

2007).

2.5.6 Nitrate and PPCPs

The wells with detectable PPCPs were more heavily concentrated in the study area’s southwest quadrant. However, wells with detectable PPCPs were not spatially clustered, nor did many of the pharmaceuticals maintain similar concentrations across wells. For instance, wells with sucralose had concentrations that ranged from 29-656 ng/L in wells separated by an average distance between each well of 6,586 m. This variability may suggest spatial dissimilarity in hydraulic conductivity and

85 biogeochemical conditions that affect solute transport of locally introduced contaminants, or may reflect location-dependent, differential contaminant loading or removal.

Excluding sulfadimethoxine, the PPCPs that were detected serve as tracers for septic-impacted drinking wells. The wells in which PPCPs were detected also had significantly higher nitrate concentrations, meaning that wells with nitrate above 4 mg/L could reflect septic-impacted drinking water, which includes eight wells identified as potentially having mixed nitrate sources (soil N, manure/septic, and/or fertilizer). In addition to these eight wells, there a total of 31 wells were identified as septic-impacted because of the presence of one or more human-utilized PPCPs.

It is important to note that nitrate concentrations in the LPRV could be directly affected by exposure to sulfamethoxazole, which can alter microbial community structure. Underwood et al. (2011) showed that such a shift altered biogeochemical transformations of nitrogen, effectively decreasing the potential for nitrate reduction by

47%. Although sulfamethoxazole was detected in only 14 wells, nitrate hotspots overlapped with half of those wells. The population of sulfamethoxazole detection and other detected PPCPs are not large enough for this to be a significant pattern.

2.5.7 Major anions as septic tracers

There was not a difference between the Cl/SO4 ratios sampled at multiple time periods. By contrast, the Mann-Whitney test suggested that anion concentrations have changed over time. Because 29 wells were sampled in both periods and there was not a discernable temporal change, it can be inferred that the apparent temporal difference of the Mann-Whitney test statistics are driven more by the difference in sampling locations of the anion concentrations from the 71 wells that were only sampled in the 2015 campaign. Future studies could further analyze the variation of anion concentrations 86 spatially and temporally to test whether groundwater dilution during periods of increased recharge is responsible.

Septic-influenced wells are expected to have a relationship of high TDS and high

- 2- nitrate and have increasing nitrate concentrations with increasing Cl and SO4 . There are

18 wells with high TDS/high nitrate, and of these wells, eight have not been previously identified as septic-impacted. Only one well had high TDS/low nitrate, possibly suggesting that denitrification occurred. If so, then this was the only indication that

2- denitrification altered nitrate concentrations. Chloride and SO4 are used as septic tracers because high concentrations are believed to come from water softeners. There were 16 wells that had sources of nitrate including ammonium from fertilizer/precipitation, soil N, and manure septic, and also had nitrate that increased with increase Cl-. Of the 16 wells, five were not previously identified as septic-impacted. Wells that have high nitrate without high Cl- could still be septic-impacted if wells with nitrate contamination lack water softeners. Wells that have high Cl-, but low nitrate could reflect contamination from road salts or denitrification (Meehan 2005).

2.5.8 Wells with unknown sources of nitrate-N

Twenty-two of the 52 wells that were originally analyzed for nitrate isotopes have isotopic signatures that overlap with three potential nitrate sources, including ammonium from fertilizer and precipitation, soil nitrogen, and manure/septic (Figure 2.4.5). These wells could have nitrate sourced from one or any combination of these three N-sources, and it is possible that soil N is present in all of the LPRV wells that were sampled.

Seventeen of these wells with unclassified N-sources have nitrate-N below 2.5 mg/L, a concentration that is found in many natural soil settings. These low concentrations may

87 signify that these wells are not impacted by contamination, but rather by the normal nitrogen cycle. However, low nitrate-N concentrations alone do not preclude the possibility anthropogenic nitrate contamination because other wells that were identified as septic-impacted had similar concentrations.

2.6 Conclusions

The LPRV wells sampled in 2015 revealed two nitrate hotspots located in Mink

Creek and around South 5th Ave. It was thought that the main cause of nitrate contamination in the LPRV drinking wells was septic systems. The results of the triple isotope signatures, PPCP tracers, nitrate concentrations, and major anions indicated that septic sources of nitrate were likely or possible for about two-thirds of the wells sampled in 2015. This multi-tracer approach conclusively identified two wells as exclusively fertilizer-impacted. Forty-one wells were definitively confirmed as septic-impacted and,

11 of these had nitrate concentrations over the EPA’s MCL. "Forty-one wells were definitively confirmed as septic-impacted, and 11 of these had nitrate concentrations over the EPA’s MCL. There was insufficient information to identify a nitrate source for the remaining 35 wells, but septic sources could not be excluded.

18 2 Evidence from δ OH2O and δ HH2O suggests snow and rain both contribute to groundwater recharge and the presence of an evaporation-like isotopic signature suggests the possibility that localized infiltration of standing water perhaps due to domestic irrigation may have much more of an impact on shallow ground water that previously recognized, with an attendant increased possibility of locally introduced contamination.

The spatial variability of nitrate concentrations could be evaluated further if wells were tested seasonally for nitrate-N. Many of the analyses, such as chloride and sulfate,

88 showed spatial heterogeneity suggesting the existence of complex flow paths and multiple nitrate sources. To determine if even more private drinking wells are impacted by septic systems, additional tracers should be used to investigate contaminant transport pathways. For instance, ibuprofen may be useful for tracing local-scale contaminant transport pathways because of its short residence time. Furthermore, a better understanding is required of groundwater flow system in this hydrogeologically complex terrain. Finally, the patterns and temporal growth of septic density in the LPRV should be validated through permit records, so as to test the assumptions made in this study and to track potential changes in septic contributions over time.

The PPCPs proved to be useful forensics tools for identifying anthropogenic contamination in wells. Isotopes of nitrate were helpful in differentiating sources of fertilizer-derived nitrogen. In future analyses, a mixing model similar to the one used for recharge sources should be used to better describe wells which are more susceptible to contamination as well as the dominant contaminant source of wells with mixed nitrate sources. This would require a controlled laboratory setting to model the LPRV soil subsurface and determine the degradation rate of PPCPs and the change in nitrate-N isotope composition as they travel through the LPRV’s variable substrate (e.g., beneath the eastern hotspot versus the western hotspot).

Both the isotope and PPCPs analyses are expensive and require scientific expertise

(Fenech et al. 2014). Therefore, future work could substitute microbiological identification of fecal indicator bacteria instead of isotopes and PPCPs. This approach would give the same identification of manure/septic that isotopes provided for this study at a lower cost (Fenech et al. 2014). However, fecal indicators would not be able to

89 differentiate sewage and manure contamination, and this study has already suggested that septic is the dominant nitrate source. Future studies should use the decision-support tool created by Fenech et al. (2014), along with the results presented here as guidance to identify the best nitrate sourcing method for individual projects.

2.7 Acknowledgements

This work has been supported by the NSF award number IIA-1301792 from the NSF

Idaho EPSCoR Program and by the National Science Foundation. Its contents are solely the responsibility of the authors and do not necessarily represent the official views of

NSF. In addition, I would like to extend thanks to Bruce Finney, Clarissa Enslin, David

Rogers, and Dave Huber for their insights. For the hard work with sampling efforts, I would also like to thank Martin Ventura, Michael Martin, Bailee Nye, and Oscar Ebanja.

90

2.8 References

Addiscott, T. M., and N. Benjamin. 2004. “Nitrate and Human Health.” Soil Use and Management 20 (2): 98–104. doi:10.1079/SUM2004256.

Basta-Kaim, Agnieszka, Bogusława Budziszewska, and Władysław Lasoń. 2008. “Effects of Antiepileptic Drugs on Immune System.” Przegla̧ d Lekarski 65 (11): 799–802. PMID: 19205364.

Batt, A. L., D. D. Snow, and D. S. Aga. 2006. “Occurrence of Sulfonamide Antimicrobials in Private Water Wells in Washington County, Idaho, USA.” Chemosphere 64 (11): 1963–71. doi:10.1016/j.chemosphere.2006.01.029.

Bell, K. Y., M. J. M. Wells, K. A. Traexler, M. Pellegrin, A. Morse, and J. Bandy. 2011. “Emerging Pollutants.” Water Environment Research 83 (10): 1906–84. doi:10.2175/106143011X13075599870298.

Benedetti, M. S., R. Whomsley, E. Baltes, and F. Tonner. 2005. “Alteration of Thyroid Hormone Homeostasis by Antiepileptic Drugs in Humans: Involvement of Glucuronosyltransferase Induction.” European Journal of Clinical Pharmacology 61 (12): 863–72. doi:10.1007/s00228-005-0056-0.

Benotti, M. J., and B. J. Brownawell. 2007. “Distributions of Pharmaceuticals in an Urban Estuary during Both Dry and Wet Weather Conditions.” Environmental Science & Technology 41 (16): 5795–5802. doi:10.1021/es0629965.

Berninger, J. P., B. Du, K. A. Connors, S. A. Eytcheson, M. A. Kolkmeier, K. N. Prosser, T. W. Valenti, C. K. Chambliss, and B. W. Brooks. 2011. “Effects of the Antihistamine Diphenhydramine on Selected Aquatic Organisms.” Environmental Toxicology and Chemistry 30 (9): 2065–72. doi:10.1002/etc.590.

Blanset, D. L., J. Zhang, and M. G. Robson. 2007. “Probabilistic Estimates of Lifetime Daily Doses from Consumption of Drinking Water Containing Trace Levels of N,N -Diethyl-Meta-Toluamide (DEET), Triclosan, or Acetaminophen and the Associated Risk to Human Health.” Human and Ecological Risk Assessment: An International Journal 13 (3): 615–31. doi:10.1080/10807030701341209.

Bohlke, J. K., and J. M. Denver. 1995. “Combined Use of Groundwater Dating, Chemical, and Isotopic Analyses to Resolve the History and Fate of Nitrate Contamination in Two Agricultural Watersheds, Atlantic Coastal Plain, Maryland.” Water Resources Research 31 (9): 2319–39. doi:10.1029/95WR01584.

Boleda, M. R., M. T. Galceran, and F. Ventura. 2009. “Monitoring of Opiates, Cannabinoids and Their Metabolites in Wastewater, Surface Water and Finished Water in Catalonia, Spain.” Water Research 43 (4): 1126–36. doi:10.1016/j.watres.2008.11.056.

91

Brooks, B. W., C. M. Foran, S. M. Richards, J. Weston, P. K. Turner, J. K. Stanley, K. R. Solomon, M. Slattery, and T. W. La Point. 2003. “Aquatic Ecotoxicology of Fluoxetine.” Toxicology Letters 142 (3): 169–83. doi:10.1016/S0378- 4274(03)00066-3.

Burow, K. R., B. T. Nolan, M. G. Rupert, and N. M. Dubrovsky. 2010. “Nitrate in Groundwater of the United States, 1991 - 2003.” Environmental Science & Technology 44 (13): 4988–97. doi:10.1021/es100546y.

Buszka, P. M., D. J. Yeskis, D. W. Kolpin, E. T. Furlong, S. D. Zaugg, and M. T. Meyer. 2009. “Waste-Indicator and Pharmaceutical Compounds in Landfill-Leachate- Affected Ground Water near Elkhart, Indiana, 2000-2002.” Bulletin of Environmental Contamination and Toxicology 82 (6): 653–59. doi:10.1007/s00128- 009-9702-z.

Cadenasso, M. L., S. T. A. Pickett, P. M. Groffman, L. E. Band, G. S. Brush, M. F. Galvin, J. M. Grove, et al. 2008. “Exchanges across Land-Water-Scape Boundaries in Urban Systems: Strategies for Reducing Nitrate Pollution.” Annals of the New York Academy of Sciences 1134: 213–32. doi:10.1196/annals.1439.012.

Camargo, J. A., A. Alonso, and A. Salamanca. 2005. “Nitrate Toxicity to Aquatic Animals: A Review with New Data for Freshwater Invertebrates.” Chemosphere 58 (9): 1255–67. doi:10.1016/j.chemosphere.2004.10.044.

Centers for Disease Control and Prevention (CDCP). 1996. “Spontaneous Abortions Possibly Related to Ingestion of Nitrate-Contaminated Well Water.” Morbidity and Mortality Weekly Report. Vol. 45. LaGrange.

Chou, I. J., K. L. Lin, H. S. Wang, and C. J. Wang. 2006. “Evaluation of Bone Mineral Density in Children Receiving Carbamazepine or Valproate Monotherapy.” Pediatric Neurology 48 (6): 317–22. doi:10.1016/S0887-8994(98)00039-3.

Collier, A. C. 2007. “Pharmaceutical Contaminants in Potable Water: Potential Concerns for Pregnant Women and Children.” Ecohealth 4 (2): 164–71. doi:10.1007/s10393- 007-0105-5.

Curt, M. D., P. Aguado, G. Sánchez, M. Bigeriego, and J. Fernández. 2004. “Nitrogen Isotope Ratios of Synthetic and Organic Sources of Nitrate Water Contamination in Spain.” Water, Air, and Soil Pollution 151 (1-4). Springer: 135–42.

Curtis, C. J., C. D. Evans, C. L. Goodale, and T. H. E. Heaton. 2011. “What Have Stable Isotope Studies Revealed About the Nature and Mechanisms of N Saturation and Nitrate Leaching from Semi-Natural Catchments?” Ecosystems 14 (6): 1021–37. doi:10.1007/s10021-011-9461-7.

92 de Jongh, C. M., P. J. F. Kooij, P. de Voogt, and T. L. ter Laak. 2012. “Screening and Human Health Risk Assessment of Pharmaceuticals and Their Transformation Products in Dutch Surface Waters and Drinking Water.” Science of the Total Environment 427-428 (November 2015): 70–77. doi:10.1016/j.scitotenv.2012.04.010.

Dejwakh, N. R., T. Meixner, G. Michalski, and J. McIntosh. 2011. “Using 17O to Investigate Nitrate Sources and Sinks in a Semi-Arid Groundwater System.” Environmental Science & Technology 46: 745–51. doi:10.1021/es203450z

Drake, V. M., and J. W. Bauder. 2005. “Ground Water Nitrate-Nitrogen Trends in Relation to Urban Development, Helena, Montana, 1971-2003.” Ground Water Monitoring and Remediation 25 (2): 118–30. doi: 10.1111/J.1745- 6592.2005.0017.X.

Drewes, J. E., T. Heberer, T. Rauch, and K. Reddersen. 2003. “Fate of Pharmaceuticals During Ground Water Recharge.” Ground Water Monitoring & Remediation. doi:10.1111/j.1745-6592.2003.tb00684.x.

Dufour-Rainfray, D., P. Vourc’h, S. Tourlet, D. Guilloteau, S. Chalon, and C. R. Andres. 2011. “Fetal Exposure to Teratogens: Evidence of Genes Involved in Autism.” Neuroscience & Biobehavioral Reviews 35 (5): 1254–65. doi:10.1016/j.neubiorev.2010.12.013.

Ewing, R., and S. Hamidi. 2014. “Measuring Sprawl 2014.” Smart Growth America: Making Neighborhoods Great Together 45 (2): 1–45. doi:10.1111/j.1728- 4465.2014.00390.x.

Fenech, C. 2014. “Differentiation of Nitrate Sources: An Environmental Forensics Approach.” Reviews in Environmental Science and Biotechnology 13: 1–4. doi:10.1007/s11157-013-9319-x.

Fenech, C., K. Nolan, L. Rock, and A. Morrissey. 2014. “Development of a Decision- Support Tool for Identifying the Most Suitable Approach to Achieve Nitrate Source Determination.” Environmental Science. Processes & Impacts 16 (11): 2564–70. doi:10.1039/c3em00620d.

Fenech, C., L. Rock, K. Nolan, J. Tobin, and A. Morrissey. 2012. “The Potential for a Suite of Isotope and Chemical Markers to Differentiate Sources of Nitrate Contamination: A Review.” Water Research 46 (7): 2023–41. doi:10.1016/j.watres.2012.01.044.

93

Ferrer, I., C.E. Heine, and E.T. Thurman. 2004. “Combination of LC/TOF-MS and LC/Ion Trap MS/MS for the Identification of Diphenhydramine in Sediment Samples.” Analytical Chemistry 76 (5): 1437–44. doi:10.1021/ac034794m.

Focazio, M. J., D. W. Kolpin, K. K. Barnes, E. T. Furlong, M. T. Meyer, S. D. Zaugg, L. B. Barber, and M. E. Thurman. 2008. “A National Reconnaissance for Pharmaceuticals and Other Organic Wastewater Contaminants in the United States - II) Untreated Drinking Water Sources.” Science of the Total Environment 402 (2-3): 201–16. doi:10.1016/j.scitotenv.2008.02.021.

Fong, P. P. 1998. “Zebra Mussel Spawning Is Induced in Low Concentrations of Putative Serotonin Reuptake Inhibitors.” Biological Bulletin 194 (2): 143–49. doi:10.2307/1543044.

Foran, C. M., J. Weston, M. Slattery, B. W. Brooks, and D. B. Huggett. 2004. “Reproductive Assessment of Japanese Medaka (Oryzias Latipes) Following a Four- Week Fluoxetine (SSRI) Exposure.” Archives of Environmental Contamination and Toxicology 46 (4): 511–17. doi:10.1007/s00244-003-3042-5.

Green, C. T., L. J. Puckett, J. K. Böhlke, B. A. Bekins, S. P. Phillips, L. J. Kauffman, J. M. Denver, and H. M. Johnson. 2005. “Limited Occurrence of Denitrification in Four Shallow Aquifers in Agricultural Areas of the United States.” Journal of Environment Quality 37 (3): 994–1009. doi:10.2134/jeq2006.0419.

Groffman, P. M., and E. J. Rosi-Marshall. 2013. “The Nitrogen Cycle.” In Fundamentals of Ecosystem Science, 137–58. doi:10.1016/B978-0-08-091680-4.00007-X.

Hollander, M., D. A. Wolfe, and E. Chicken. 2014. Nonparametric Statistical Methods. 3rd ed. New Jersey: John Wiley & Sons, 39-55. ISBN: 9780470387375.

Isojärvi, J. I. T., E. Löfgren, K. S. T. Juntunen, A. J. Pakarinen, M. Päivänsalo, I. Rautakorpi, and L. Tuomivaara. 2004. “Effect of Epilepsy and Antiepileptic Drugs on Male Reproductive Health.” Neurology 62 (2): 547–253. doi:10.1212/01.WNL. 0000098936.46730.64.

Kasprzyk-Hordern, B., R. M. Dinsdale, and A. J. Guwy. 2009. “Illicit Drugs and Pharmaceuticals in the Environment - Forensic Applications of Environmental Data. Part 1: Estimation of the Usage of Drugs in Local Communities.” Environmental Pollution 157 (6): 1773–77. doi:10.1016/j.envpol.2009.03.017.

Kendall, C. 1998. “Tracing Nitrogen Sources and Cycling in Catchments.” In Isotope Tracers in Catchment Hydrology, edited by C. Kendall and J. J. McDonnell, 1st ed., Amsterdam, 519–76. Elsevier.

94

Kendall, C., and D. H. Doctor. 2003. “Stable Isotope Applications in Hydrologic Studies.” In Treatise on Geochemistry, edited by J. I. Drever, H. D. Holland, and K. K. Turekian, Volume 5, 319–64. Elsevier. doi:10.1016/B0-08-043751-6/05081-7.

Kendall, C., E. M. Elliott, and S. D. Wankel. 2008. “Tracing Anthropogenic Inputs of Nitrogen to Ecosystems.” In Stable Isotopes in Ecology and Environmental Science, edited by R. Michener and K. Lajtha, 2nd ed., 375–449. doi:10.1002/9780470691854.ch12.

Kummerer, K. 2008. Pharmaceuticals in the Environment. Edited by Klaus Kümmerer. Pharmaceuticals in the Environment: Sources, Fate, Effects, and Risks. 2nd ed., 3- 38. Berlin: Springer Berlin Heidelberg. doi:10.1007/978-3-540-74664-5.

LaGro, J. A. 1996. “Designing without Nature: Unsewered Residential Development in Rural Wisconsin.” Landscape and Urban Planning 35 (1): 1–9. doi:10.1016/0169- 2046(96)00314-3.

Madison, R. J., and J. O. Brunett. 1985. “Overview of the Occurrence of Nitrate in Ground Waters of the United States.” In National Water Summary 1984-Hydrologic Events, Selected Water-Quality Trends, and Ground-Water Resources: U.S. Geological Survey Water-Supply Paper 2275, 93–105. Washington, DC.

Meehan, C. W. 2005. “Geochemistry of the Southern Pocatello Aquifer and Its Implications for Aquifer Recharge and Contamination Potential.” Master’s Thesis. Idaho State University, Pocatello. http://geology.isu.edu/thesis/Meehan_Christopher_2005_ISU.pdf.

Michalski, G., J.K. Böhlke, and M. Thiemens. 2004a. “Long Term Atmospheric Deposition as the Source of Nitrate and Other Salts in the Atacama Desert, Chile: New Evidence from Mass-Independent Oxygen Isotopic Compositions.” Geochimica et Cosmochimica Acta 68 (20): 4023–38. doi:10.1016/j.gca.2004.04.009.

Michalski, G., T. Meixner, M. Fenn, L. Hernandez, A. Sirulnik, E. Allen, and M. Thiemens. 2004b. “Tracing Atmospheric Nitrate Deposition in a Complex Semiarid Ecosystem Using Δ17O.” Environmental Science and Technology 38 (7): 2175–81. doi:10.1021/es034980+.

Miller, Kathleen J., and Joseph Meek. 2006. “Helena Valley Groundwater: Pharmaceuticals, Personal Care Products, Endocrine Disruptors (PPCPs) and Microbial Indicators of Fecal Contamination.” Montana Bureau of Mines and Geology Open-File Report 532, 19. http://www.mbmg.mtech.edu/pdf-open- files/mbmg532-helenavalley.pdf.

95

Minnesota, Division of Environmental Health. 2013. “Toxicological Summary Sheet for Carbamazepine CAS: 298-46-4.” Health Risk Limits for Groundwater 651-201- 4899. http://www.health.state.mn.us/divs/eh/risk/guidance/gw/carbamazepine.pdf.

Mitchell, R. J., R. S. Babcock, S. Gelinas, L. Nanus, and D. E. Stasney. 2003. “Nitrate Distributions and Source Identification in the Abbotsford-Sumas Aquifer, Northwestern Washington State.” Journal of Environmental Quality 32 (3): 789– 800. doi:10.2134/jeq2003.0789.

Musolff, Andreas, Sebastian Leschik, Monika Möder, Gerhard Strauch, Frido Reinstorf, and Mario Schirmer. 2009. “Temporal and Spatial Patterns of Micropollutants in Urban Receiving Waters.” Environmental Pollution 157 (11): 3069–77. doi:10.1016/j.envpol.2009.05.037.

Nestler, A., Mi. Berglund, F. Accoe, S. Duta, D. Xue, P. Boeckx, and P. Taylor. 2011. “Isotopes for Improved Management of Nitrate Pollution in Aqueous Resources: Review of Surface Water Field Studies.” Environmental Science and Pollution Research 18 (4): 519–33. doi:10.1007/s11356-010-0422-z.

Osenbrück, K., H. R. Gläser, K. Knöller, S. M. Weise, M. Möder, R. Wennrich, M. Schirmer, F. Reinstorf, W. Busch, and G. Strauch. 2007. “Sources and Transport of Selected Organic Micropollutants in Urban Groundwater Underlying the City of Halle (Saale), Germany.” Water Research 41 (15): 3259–70. doi:10.1016/j.watres.2007.05.014.

Phillips, D. L., and J. W. Gregg. 2001. “Uncertainty in Source Partitioning Using Stable Isotopes.” Oecologia 127 (2): 171–79. doi:10.1007/s004420000578.

Pohl, H., and J. Colman. 2006. Interaction Profile for: Atrazine, Deethylatrazine, Diazinon, Nitrate, and Simazine. Agency for Toxic Substances and Disease Registry (ATSDR). Atlanta: U.S. Department of Health and Human Services.

Powlson, D. S., T. M. Addiscott, N. Benjamin, K. G. Cassman, T. M. de Kok, H. van Grinsven, J. L’Hirondel, A. Avery, and C. van Kessel. 2008. “When Does Nitrate Become a Risk for Humans?” Journal of Environmental Quality 37 (2): 291–95. doi:10.2134/jeq2007.0177.

Rosi-Marshall, E. J., D. Snow, S. L. Bartelt-Hunt, A. Paspalof, and J. L. Tank. 2014. “A Review of Ecological Effects and Environmental Fate of Illicit Drugs in Aquatic Ecosystems.” Journal of Hazardous Materials 282: 18–25. doi:10.1016/j.jhazmat.2014.06.062.

Rosi-Marshall, E. J., D. W. Kincaid, H. A. Bechtold, T. V. Royer, M. Rojas, and J. J. Kelly. 2013. “Pharmaceuticals Suppress Algal Growth and Microbial Respiration and Alter Bacterial Communities in Stream Biofilms.” Ecological Applications 23 (3): 583–93. doi:10.1890/12-0491.1.

96

Schlegel, M. E., A. L. Mayo, S. Nelson, D. Tingey, R. Henderson, and D. Eggett. 2009. “Paleo-Climate of the Boise Area, Idaho from the Last Glacial Maximum to the Present Based on Groundwater δ2H and δ18O.” Quaternary Research 71 (2). University of Washington: 172–80. doi:10.1016/j.yqres.2008.09.008.

Spalding, R. F., and M. E. Exner. 1993. “Occurrence of Nitrate in Groundwater- A Review.” Journal of Environment Quality 22 (3): 392. doi:10.2134/jeq1993.00472425002200030002x.

Squillace, P. J., J. C. Scott, M. J. Moran, B. T. Nolan, and D. W. Kolpin. 2002. “VOCs, Pesticides, Nitrate, and Their Mixtures in Groundwater Used for Drinking Water in the United States.” Environmental Science & Technology 36 (9): 1923–30. doi:10.1021/es015591n.

Stackelberg, P. E., E. T. Furlong, M. T. Meyer, S. D. Zaugg, A. K. Henderson, and D. B. Reissman. 2004. “Persistence of Pharmaceutical Compounds and Other Organic Wastewater Contaminants in a Conventional Drinking-Water-Treatment Plant.” Science of the Total Environment 329: 99–113. doi:10.1016/j.scitotenv.2004.03.015.

Sui, Q., X. Cao, S. Lu, W. Zhao, Z. Qiu, and G. Yu. 2015. “Occurrence, Sources and Fate of Pharmaceuticals and Personal Care Products in the Groundwater: A Review.” Emerging Contaminants 1 (1): 14–24. doi:10.1016/j.emcon.2015.07.001.

Ternes, T. A., A. Joss, and H. Siegrist. 2004. “Scrutinizing Pharmaceuticals and Personal Care Products in Wastewater Treatment.” Environmental Science and Technology 38 (20): 392–99. doi:10.1021/es040639t.

Tu, J. 2011. “Spatially Varying Relationships between Land Use and Water Quality across an Urbanization Gradient Explored by Geographically Weighted Regression.” Applied Geography 31 (1): 376–92. doi:10.1016/j.apgeog.2010.08.001.

U.S. Census Bureau. 2010. “Pocatello City, Idaho Quick Facts.” Census of Population and Housing. http://www.census.gov/quickfacts/table/HSG010214/1664090.

US-EPA. 1992. “Specifications and Guidance for Contaminant-Free Sample Containers.” Environmental Protection Agency Report: A540/R-93/051 PB93-963316. Washington, DC: 1–26.

Verhoeven, J. T. A., B. Arheimer, C. Yin, and M. M. Hefting. 2006. “Regional and Global Concerns over Wetlands and Water Quality.” Trends in Ecology and Evolution 21 (2): 96–103. doi:10.1016/j.tree.2005.11.015.

Ward, M. H., B. A. Kilfoy, P. J. Weyer, K. E. Anderson, A. R. Folsom, and J. R. Cerhan. 2010. “Nitrate Intake and the Risk of Thyroid Cancer and Thyroid Disease.” Epidemiology 21 (3): 389–95. doi:10.1097/EDE.0b013e3181d6201d.

97

Ward, M. H., S. D. Mark, K. P. Cantor, D. D. Weisenburger, A. Correa-Villaseñor, and S. H. Zahm. 1996. “Drinking Water Nitrate and the Risk of Non-Hodgkin’s Lymphoma.” Epidemiology 7 (5): 465–71. http://www.ncbi.nlm.nih.gov/pubmed/8862975.

Welhan, J. A. 2006. “Water Balance and Pumping Capacity of the Lower Portneuf River Valley Aquifer, Bannock County, Idaho.” Idaho Geological Survey Report 06-5. Moscow: 1–36.

Welhan, J. and E. Dombrowski, 2014, Recharge Processes in the Lower Portneuf River Valley Watershed; Idaho NSF EPSCoR Annual Meeting, October 9, Boise.

Weyer, P .J., J. R. Cerhan, B. C. Kross, G. R. Hallberg, J. Kantamneni, G. Breuer, M. P. Jones, W. Zheng, and C. F. Lynch. 2001. “Municipal Drinking Water Nitrate Level and Cancer Risk in Older Women: The Iowa Women’s Health Study.” Epidemiology 12 (3): 327–38. http://www.ncbi.nlm.nih.gov/pubmed/11338313.

Wood, Warren W., and Ward E Sanford. 1995. “Chemical and Isotopic Methods for Quantifying Groundwater Recharge in a Regional, Semiarid Environment.” Groundwater 33 (3): 458–68. doi:10.1111/j.1745-6584.1995.tb00302.x.

Xue, D., J. Botte, B. De Baets, F. Accoe, A. Nestler, P. Taylor, O. Van Cleemput, M. Berglund, and P. Boeckx. 2009. “Present Limitations and Future Prospects of Stable Isotope Methods for Nitrate Source Identification in Surface and Groundwater.” Water Research 43 (5): 1159–70. doi:10.1016/j.watres.2008.12.048.

98

3 Chapter 3: Factors Influencing Public Risk Perceptions of Drinking Water Quality in the LPRV and its Effect on Behavior

Abstract

A risk perception framework was used to evaluate risk as analysis, risk as feelings, and risk as politics, to examine how risk perception affected behavior associated with drinking water quality. Actions that respondents might take in response to perceived risk by filtering, softening, or treating, regularly monitoring or testing water quality, buying bottled water, or doing nothing were evaluated. The results showed that measured water quality standards and trust in city officials do not influence public risk perceptions.

Instead three factors were more important: whether or not water was privately supplied, risks associated with unknown dangers, and education.

3.1 Introduction

Humans depend on clean water for municipal and private supplies in many cities.

The cities of Pocatello and Chubbuck in semi-arid southeastern Idaho, USA, rely on clean groundwater from the Lower Portneuf River Valley (LPRV) aquifer as the primary source of potable water. Approximately 77% of Pocatello residents get their drinking water from city services and the remaining 23% of households have their own private water supplies through wells (U.S. Census Bureau 2010). Maintaining private water supplies is the responsibility of the well owners rather than government. Thus, the perception of water quality of private well owners is important because perception drives action (Wachinger et al. 2013), including well maintenance and water quality testing

(Doria et al. 2009). Maintenance and use of public water supplies also may be influenced by risk perception, with more complex relationships between citizen and public

99 institutions that are moderated by trust in those institutions. Thus, regardless of the source of water, risk perceptions influence actions that can affect water quality.

3.1.1 Groundwater and Water Quality in Pocatello, ID

Groundwater quality around Pocatello is affected by nitrate, emerging contaminants such as antibiotics, and trichloroethylene (TCE) (CH2M-Hill 1994, Welhan et al 1996, Meehan 2005, Ohr et al. in preparation). These contaminants have been attributed to septic leachate, road salts, and runoff from agriculture and ranching operations in the LPRV (CH2M-Hill 1994, Meehan 2005), or some mixture of these

- sources (Ohr et al. in preparation). Although nitrate (NO3 ) is a naturally occurring ion within the nitrogen (N) cycle, alteration of the N cycle for crop production and urbanization can lead to N excess in the ecosystem and water quality concerns

(Cadenasso et al. 2008, Kaushal et al. 2011). High nitrate concentrations in drinking water can adversely affect human health (Addiscott & Benjamin 2004, Powlson et al.

2008), causing cyanosis, altered thyroid function, rapid pulse, and a build-up of protein in the liver, spleen, or kidney (Pohl & Colman, 2006). Children under the age of six months can develop nitrate-induced methemoglobinemia, a condition that produces too much methemoglobin, preventing oxygen uptake in the blood (Mueller & Helsel, 1996). Septic systems can introduce nitrate and other contaminants of concern, including pharmaceuticals and personal care products (PPCPs). Pharmaceuticals are introduced to ground and surface waters when they are ingested and then excreted into the environment, passing through septic/sewer systems or directly entering the environment through excretion by husbandry animals (Figure 3.1.1). PPCPs are not always removed during waste water treatment, and can be flushed into ground and surface waters from

100 septic systems, waste treatment plants or leaking sewage pipes. Even low concentrations of pharmaceuticals can lead to negative human health effects (Ternes et al. 2004, Seiler et al. 1999 & Lucia et al. 2010). Exposure to pharmaceuticals can cause interference of hormone production from endocrine disrupters (Ternes et al. 2004) and prenatal health effects, which can include autism (Dufour-Rainfray et al. 2011).

Figure 3.1.1 Pathways for input and transport of pharmaceuticals into drinking water. Both human and. veterinary pharmaceuticals can contribute to drinking water contamination (adapted from Figure 1.1 in Kummerer 2004).

3.1.2 Risk perception

A person’s response to a potential threat, such as drinking water contamination, depends on their perception of risk associated with that threat (Peters et al. 2004).

Perception is the process of filtering information to create an understanding of uncertain impacts of the surrounding world (Wachinger et al. 2013). Risk perception is how that understanding pertains to characteristics or severity of various threats. In the psychological literature, risk perception integrates three different risk types: risk as

101 feelings, risk as analysis, and risk as politics (Slovic et al. 2004). Risk as feelings refer to instinctive and intuitive reactions to danger, which is the predominant method by which humans evaluate risk (Peters et al. 2004, Slovic et al. 2004, Slovic and Peters 2006). Risk as analysis uses scientific research, logic, and reason to form perceptions (Slovic et al.

2004). Risk as politics accounts for a person’s point of view or worldview, which affects their risk judgments (Slovic 1999). For instance, a person with a hierarchy worldview would trust experts or authorities to properly manage water quality, which could decrease their risk perception (Slovic 1999). Trust in government officials also acts with the bounds of risk as politics; if someone has less trust in government then they are more likely to have an increased risk perception (Robinson et al. in press).

Risk perceptions also vary based on knowledge and experience. For instance, experts and the untrained public have differing opinions of risk severity because their dominant methods of perception formation differ (Slovic 1987). Experts tend to quantify risk based on analysis of technical reports (e.g., annual fatalities). Untrained individuals often assess risk based on feelings. While experts might assess risk primarily based on two variables, the number of lives at stake and probability of injury, untrained individuals often rely on more factors (Slovic 1987). Risk perception for the public can be categorized into two categories of risk as feelings: 1) dread risk- the feeling of dread, lack of control, perceived potential catastrophes, and risk versus benefit, and 2) unknown risk- risk that is new, unknown, unobservable, or has delayed harmful effects (Peters et al. 2004). Emotions and rational thought interact to form a person’s risk perception. This cognitive-emotive process is influenced by social and political demographics, past experience, likelihood of risk occurrence, and information from the social environment

102

(Sjoberg 2000, Peters et al. 2004, Slovic et al. 2004, Stoutenbourough et al. 2013,

Stoutenborough et al. 2014).

3.1.3 1.4 Risk Perception of Water Quality

Risk perception associated with drinking water quality is affected by these same cognitive-emotive processes (Doria 2010). Taste or smell of the water, trust in institutions, political and social demographics, perceived control, and experience are key factors associated with water quality risk perception (Johnson 2003, Jones et al. 2006,

Doria 2009, Hu et al. 2011).

Sensory information, such as color and flavor, affect public perception of risk associated with city water (Doria et al. 2005, Doria 2010). Personal concern about drinking water also depends on an individual’s perceptions of serious local environmental health problems and by a sense of how much personal control they have over their own health (Johnson 2003). In addition, those who remember past water-related health problems tend to have higher risk perceptions because prior personal experience provides context for analyzing new information (Griffen & Dunwoody 2000, Doria et al. 2005).

Social information about water quality is expected to influence risk perception because we learn from direct experience and from the experiences of others (Rosenstock et al.

1988).

Along with social information, technical information would be expected to influence risk perceptions. Due to the Safe Drinking Water Act Amendments of 1996, city water departments are required to give customers annual water quality reports. This reporting is assumed to affect customers’ risk perceptions of water quality. However, reporting scientific and technical information about water standard violations did not

103 increase risk perceptions in New Jersey even when violations were clearly marked

(Johnson 2003). Furthermore, media reporting of potential water quality problems has not been clearly linked to increases in public risk perception (Sjoberg 2000). One possible reason that people are not strongly influenced by water quality reporting is that trust in government is poorly correlated with risk perception associated with water quality

(Johnson 2003, Doria et al. 2005).

Water quantity and quality are inherently linked concepts, and thus, the insights based on risks associated with water scarcity may inform risk perceptions of water quality. Risk perception studies in other semi-arid environments have found that women tend to be more concerned about water quantity (Larson et al. 2010) and that concern varied by ethnicity, ecological worldview, and political identification (Larson et al.

2011). These groups may also be more concerned about water quality in the LPRV because local droughts may affect water quality and quantity in the future.

Here the main factors influencing risk perception were assessed. These factors ultimately drive self-described actions to alter (or attempt to alter) water quality. It was hypothesized that actual water quality, feelings of concern, and worldview or trust in government may be important predictors of risk perception and action, as outlined below:

H1: Neighborhoods that have elevated nitrate-N concentrations will have

increased risk perceptions and will be more likely to alter their drinking water.

High nitrate concentrations will encourage water filtration, treatment, or testing,

or will lead people to buy bottled water.

104

H2: Respondents who are more concerned about pharmaceuticals in their drinking

water or are more concerned about health issues due to water quality will have

increased risk perceptions and be more likely to alter their drinking water.

H3: Respondents with less trust in city government will be more likely to have

increased risk perception and therefore will be more likely to alter their drinking

water.

3.2 Methods

To analyze risk perceptions, a randomized mail survey of residents within the

Pocatello city limit was used. Approximately 1,500 surveys were mailed and there were close to 375 responses, representing ~1% of the cities’ adult population. The demographics for the cities of Chubbuck and Pocatello and of the public survey respondents can be found below in Table 3.2.1 and Table 3.2.2, respectively. To account for the fact that survey results do not perfectly reflect distributions of the general public, demographics were included to account for any bias that may have occurred.

Respondents’ perception of risk of contaminated drinking water, trust in city government, social and political demographics, knowledge of the groundwater system, and their current source of drinking water was analyzed. All three risk perception concepts - feelings, analysis, and politics- were used to model respondents’ risk perceptions of water quality and their behavioral responses. Risk as feelings were assessed using two different questions, respondents’ concern with health issues due to pollution and also with pharmaceuticals in groundwater. Water quality measurements across the Pocatello region were compiled to assess risk perceptions stemming from risk as analysis. Nitrate-N concentrations [in milligrams per liter] were assigned to each respondent based on their

105 reported neighborhood affiliation (no addresses were collected with each survey). Nitrate concentrations were measured in city well waters by the City of Pocatello Water

Department at least quarterly. Concentrations were averaged for the years 2013-2015, and then spatially interpolated using ordinary kriging with a normal score transformation and a stable semivariogram function with a nugget effect of 0.4, and a lag size of 265 m.

The interpolated nitrate was then averaged by neighborhood to link water quality information to survey responses. Risk as politics was assessed by testing respondents for a hierarchy worldview, where trust in city officials used an analogous 5-point scale from no trust to complete trust. The effects of risk perception on behavior were then analyzed.

Respondents reacted to perceived risk through a range of behaviors, including filtering, softening, or treating water, regularly monitoring or testing water quality, buying bottled water, or doing nothing.

Table 3.2.1 Combined demographics for the cities of Pocatello and Chubbuck based on the 2010 U.S. Census. Residents relying on city water supplies in Chubbuck could not be determined, and therefore, data for the city vs. private water supply is provided for Pocatello residents only.

Table 3.2.2 Demographics of survey respondents

106

Action was assessed from one question, which prompts, “Which of the following do you do for your household water needs?” Possible answers were nothing, filter water, soften water, treat water (other than filtering or softening), test water, or buy bottled water. Action was potentially influenced by risk perception or a suite of demographic and non-demographic variables. Two questions pertained to drinking water sources. The first asked the respondents if they are on city or well water and the second assessed their knowledge of where the city gets its drinking water (i.e. river or groundwater). Another question asked whether the household was using septic or sewage to meet their wastewater needs. In each of these questions, respondents rated their level of concern on a 5-point Likert scale. Ideology and party affiliation were rated on their degree of identification, very liberal to very conservative and strong democrat to strong republican.

Demographic identifiers for age, education, marital status, gender, and income were also used in this study. The verbatim survey questions are compiled in Appendix G.

Expected relationships between survey responses, risk perception, and action are summarized below (Figure 3.2.1Error! Reference source not found.). Concern level and nitrate-N (mg/L) are expected to influence risk perceptions of water quality, which in turn are expected to affect individual actions (filtering, softening, treating, testing, buying bottled water, or nothing). In addition, demographics and other independent variables are expected to directly influence individual actions.

107

Figure 3.2.1 Independent variables used in maximum-likelihood probit model for testing the relationship between these controls, risk perception and dependent action variables (filtering, softening, treating, testing, buying bottled water, or nothing)

The independent variables included both coded variables (5-point concern and political affiliation) and continuous data. The dependent variables were binary (i.e., neither continuous nor ordered) and thus a maximum likelihood probit model was used in

STATA for all models, excluding the model used for respondents who test their water. A complementary log-log regression was used for the respondents who test their water because there existed rare-event bias in the probit model. The probit model is ideal for handling dichotomous variables because it does not assume a linear relationship between variables and does not assume all independent variables are continuous. These models were designed by Dr. James Stoutenborough and co-implemented and analyzed by

Courtney Ohr.

3.3 Results

Figure 3.3.2-Figure 3.3.4 represent regression analysis results with boxes around the grouped independent variables or risk perception and ovals around the resulting action.

All continuous variables show the sign of the relationship (+/-). For example, a (+) means

108 that as the independent variable increases, there is an increased likelihood that a person will take that type of action.

3.3.1 Softening and Filtering

Table 3.3.1 Factors influencing risk perception that lead to softening and filtering. The strongest controls here are demographic. The factor “wastewater system” is positive when the respondent has a residential septic system and negative when the respondent is using city sewer.

Figure 3.3.1 Significant factors that influence risk perception for survey respondents who soften and filter their water. All continuous variables show the sign of the relationship (+/-).

As age increases, so does the likelihood of softening or filtering water. Within these models, neither health concern nor nitrate-N concentrations have a significant

109 relationship with action. Instead social and political demographics were the key variables that drove the use of water softeners.

3.3.2 Testing and Treating

Table 3.3.2 Factors influencing risk perception that lead to testing and treating water. The strongest controls are related to education, concern, and whether the respondent’s primary water source was private or municipal.

Figure 3.3.2 Significant factors that influence risk perception for survey respondents who test and treat their water. All continuous variables show the sign of the relationship (+/-).

Drinking water from a private well supply, increased concern about pharmaceuticals, and increased education were the strongest predictors of testing and treating water. Residents

110 drinking municipal water tested and treated their water significantly less often than residents drinking private well water.

3.3.3 Buying Bottled Water

Table 3.3.3 Factors influencing risk perception that lead to buying bottled water

Figure 3.3.3 Significant factors influencing risk perception for survey respondents who buy bottled water. All continuous variables show the sign of the relationship (+/-).

111

Respondents who were located in neighborhoods with lower nitrate-N concentrations were more likely to buy bottled water. They also tended to be younger in age, female, identified as democratic, and were more educated.

3.3.4 Respondents who reported doing nothing to their water supply

Table 3.3.4 Factors influencing risk perception that lead to respondents doing nothing to their water supply

112

Figure 3.3.4 Significant factors influencing risk perception for survey respondents who do nothing. All continuous variables show the sign of the relationship (+/-).

Survey respondents who do nothing to alter their water supply have less income, are single, and are the least concerned about pharmaceuticals in drinking water.

3.4 Discussion

3.4.1 Risk as analysis

City residents who drink from the public water supply may rely on municipal water quality reports, but none of the model results suggested these reports affect action. Risk perception does not seem to be the main driver of bottled water consumption in Pocatello.

One might expect that residents drinking municipal water in areas with the lowest nitrate-

N concentrations would be less concerned about water quality risks, but these residents are most likely to use bottled water.

3.4.2 Risk as feelings

Interestingly, those who are less concerned about health issues due to water quality are also more likely to treat their water, but not necessarily test their water. It is possible that their lack of concern reflects that they are already treating their water. Even if they were not concerned about general health issues related to water quality, respondents who

113 reported testing or treating their water were concerned about pharmaceuticals in water supplies. The first study of pharmaceuticals in Pocatello was completed after these survey results were collected. Risk perception associated with pharmaceutical contamination may be due to unknown risks rather than from known exposure (Peters et al. 2004). If this survey was the first time individuals were exposed to the idea of pharmaceutical contamination, then "perceived unknown risks" from a new substance could increase overall risk perceptions. Alternatively, if respondents already knew about possible side effects from continuous, low level exposure of pharmaceuticals, then perception could be influenced by dread risks (Sjoberg 2000, Stoutenborough 2015). Some pharmaceuticals are prevalent and persistent within natural systems, but are difficult to observe without expensive technical equipment. This can lead to a feeling of lack of control over one’s own health or water quality, which increases dread risk and overall risk perception.

Negative effects from different types of pharmaceutical exposure are poorly understood, difficult to observe, and may be delayed following exposure; all of these factors lead to increased unknown risk (Sjoberg 2000, Peters et al. 2004, Stoutenborough et al. 2013).

The sub-population that self-identified as doing nothing was the least likely to be concerned about pharmaceuticals in their drinking water. These respondents may not be concerned with unknown risks, or they do perceive some risks, but are unable to afford to do anything about them.

3.4.3 Risk as politics

Residents with city supplied water may trust that public water supplies will meet drinking water standards whether or not they review annual drinking water quality reports, but these survey results cannot address those possibilities.

114

3.4.4 Other independent variables

Respondents who reported testing and/or treating their water had increased education and were on private well-supplied water. Because respondents who were more likely test and/or treat their water were on private well- supplied water, this finding may indicate that private well owners know that they are responsible for maintaining water quality for their private wells. This may be attributed as success of past city initiatives to educate private well owners of their responsibilities for maintaining their own water quality.

People softening water do not report a perception of significant water quality risks.

Instead those with larger incomes were more likely to soften their water, and economic drivers were the strongest predictor of this action. Key explanatory variables that predict whether people will filter their water could be missing because age was the only significant predictor. The distinction between filtering and treating water was unclear based on responses to an open-ended question that allowed respondents to specify the type of filtering system that they used. Some respondents noted use of an activated carbon (Brita-like) system that improves taste, but does not remove most pharmaceuticals or nitrate. Others specified that filtering referred to a reverse osmosis system, which lowers nitrate-N concentrations. It may be expected that respondents who use activated carbon filters might be more similar to those who use softeners, whereas those using reverse osmosis systems might be more similar to those who reported “treating” their water.

Though this survey population is skewed towards older respondents, these results are consistent with other research that has shown that younger populations and women are more likely to use bottled water (Hu et al. 2011). Political affiliation was also a strong

115 predictor of one action: Democrats were more likely to use bottled water. The survey did not specify how often the respondent used bottled water. If the question had specified whether respondents drank bottled water as their primary water source, survey results might have differed.

3.5 Conclusion

It can be difficult to discern factors that influence risk perceptions for drinking water and how they in turn drive actions. In this study, the hypothesis that measured water quality was the major predictor of risk perceptions and actions that would alter water quality.

Instead demographics, whether respondents’ primary water source was private or municipal, and concern about pharmaceuticals or water quality were important factors for those who take action to improve their water quality. These observations support the 2nd hypothesis: risk perception increased when there was more concern about pharmaceuticals in drinking water, which in turn drove the actions of testing and treating water. Finally, trust in city government was not a significant predictor of risk perception or action in any model. In addition to these assessments, those respondents who drink private well water were significantly more likely to test their water quality. Past education measures that encouraged testing of private water supplies may have contributed to this response, but this possibility was not confirmed in the survey.

3.6 Acknowledgments

This publication was made possible by the NSF Idaho EPSCoR Program and by the

National Science Foundation under award number IIA-1301792. Its contents are solely the responsibility of the authors and do not necessarily represent the official views of

NSF.

116

3.7 References

Addiscott, T. M., and N. Benjamin. 2004. “Nitrate and Human Health.” Soil Use and Management 20 (2). Wiley Online Library: 98–104. doi:10.1111/j.1475- 2743.2004.tb00344.x.

Cadenasso, M. L., S. T A Pickett, P. M. Groffman, L. E. Band, G. S. Brush, M. F. Galvin, J. M. Grove, et al. 2008. “Exchanges across Land-Water-Scape Boundaries in Urban Systems: Strategies for Reducing Nitrate Pollution.” Annals of the New York Academy of Sciences 1134: 213–32. doi:10.1196/annals.1439.012.

CH2M-Hill. 1994. “Hydrogeology and Assessment of TCE Contamination in the Southern Portion of the Pocatello Aquifer.” Phase I Aquifer Management Plan Final Report, 1–100.

Doria, M. F. 2006. “Bottled Water versus Tap Water: Understanding Consumer’s Preferences.” Journal of Water and Health 4 (2): 271–76. doi:10.2166/wh.2006.008.

Doria, M. F. 2010. “Factors Influencing Public Perception of Drinking Water Quality.” Water Policy 12 (1): 1–19. doi:10.2166/wp.2009.051.

Doria, M. F., N. Pidgeon, and P. R. Hunter. 2005. “Perception of Tap Water Risks and Quality: A Structural Equation Model Approach.” Water Science and Technology : A Journal of the International Association on Water Pollution Research 52 (8): 143– 49. http://www.ncbi.nlm.nih.gov/pubmed/16312961.

Doria, M. F., N. Pidgeon, and P. R. Hunter. 2009. “Perceptions of Drinking Water Quality and Risk and Its Effect on Behaviour: A Cross-National Study.” Science of the Total Environment 407 (21). Elsevier B.V.: 5455–64. doi:10.1016/j.scitotenv.2009.06.031.

Johnson, B. B. 2003. “Do Reports on Drinking Water Quality Affect Customers’ Concerns? Experiments in Report Content.” Risk Analysis 23 (5): 985–98. doi:10.1111/1539-6924.00375.

Jones, A. Q., C. E. Dewey, K. Doré, S. E. Majowicz, S. A. McEwen, D. Waltner-Toews, M. Eric, D. J. Carr, and S. J. Henson. 2006. “Public Perceptions of Drinking Water: A Postal Survey of Residents with Private Water Supplies.” BMC Public Health 6: 94. doi:10.1186/1471-2458-6-94.

Kaushal, S. S., P. M. Groffman, L. E. Band, E. M. Elliott, C. A. Shields, and C. Kendall. 2011. “Tracking Nonpoint Source Nitrogen Pollution in Human-Impacted Watersheds.” Environmental Science and Technology 45 (19): 8225–32. doi:10.1021/es200779e.

117

Larson, K. L., A. Wutich, D. White, Ti. A. Munoz-Erickson, and S. L. Harlan. 2011. “Multifaceted Perspectives on Water Risks and Policies: A Cultural Domains Approach in a Southwestern City.” Human Ecology Review 18 (1): 75–87.

Larson, K. L., D. C. Ibes, and D. D. White. 2010. “Gendered Perspectives About Water Risks and Policy Strategies: A Tripartite Conceptual Approach.” Environment and Behavior 43 (3): 415–38. doi:10.1177/0013916510365253.

Meehan, C. W. 2005. “Geochemistry of the Southern Pocatello Aquifer and Its Implications for Aquifer Recharge and Contamination Potential.” Idaho State University. http://www.bannockcounty.us/groundwater/wp- content/uploads/Meehan-Thesis.pdf.

Mueller, D. S., and D. R. Helsel. 1996. “Nutrients in the Nation’s Waters: Too Much of a Good Thing?” U.S. Geological Survey Circular, no. 1136: 24.

Peters, E. M., B. Burraston, and C. K. Mertz. 2004. “An Emotion-Based Model of Risk Perception and Stigma Susceptibility: Cognitive Appraisals of Emotion, Affective Reactivity, Worldviews, and Risk Perceptions in the Generation of Technological Stigma.” Risk Analysis 24 (5): 1349–67. doi:10.1111/j.0272-4332.2004.00531.x.

Pohl, H., and J. Colman. 2006. “Interaction Profile for: Atrazine, Deethylatrazine, Diazinon, Nitrate, and Simazine.” U.S. Department of Health and Human Services. Agency for Toxic Substances and Disease Registry, no. August: 1–68.

Powlson, D. S., T. M. Addiscott, N. Benjamin, K. G. Cassman, and T. M. de Kok. 2008. “When Does Nitrate Become a Risk for Humans?” Journal of Environment Quality 37 (2): 291. doi:10.2134/jeq2007.0177.

Robinson, Scott E., James W. Stoutenborough, and Arnold Vedlitz. Under Contract. Understanding Trust in Government: Environmental Sustainability, Fracking, and Public Opinion in American Politics. New York: Routledge.

Rosenstock, I. M., V. J. Strecher, and M. H. Becker. 1988. “Social Learning Theory and the Health Belief Model.” Health Education Quarterly 15 (2): 175–83. doi:10.1177/109019818801500203.

Seiler, R. L., S. D. Zaugg, J. M. Thomas, and D. L. Howcroft. 1999. “Caffeine and Pharmaceuticals as Indicators of Waste Water Contamination in Wells.” Groundwater 37 (3): 405–10. doi:10.1111/j.1745-6584.1999.tb01118.x.

Sjoberg, L. 2000. “Factors in Risk Perception.” Risk Analysis 20 (1): 1–11. doi:10.1111/0272-4332.00001.

Slovic, P. 1987. “Perception of Risk.” Science 236 (4799): 280–85. doi:155.97.178.73.

118

Slovic, P. 1999. “Trust, Emotion, Sex, Politics, and Science: Surveying the Risk- Assessment Battlefield.” Risk Analysis 19 (4): 689–701. doi:10.1111/j.1539- 6924.1999.tb00439.x.

Slovic, P., M. L. Finucane, E. Peters, and D. G. MacGregor. 2004. “Risk as Analysis and Risk as Feelings: Some Thoughts about Affect, Reason, Risk, and Rationality.” Risk Analysis 24 (2): 311–22. doi:10.1111/j.0272-4332.2004.00433.x.

Slovic, P., and E. Peters. 2006. “Risk Perception and Affect.” Current Directions in Psychological Science 15 (6): 322–25. doi:10.1177/03063127067078012.

Stoutenborough, J. W., S. G. Sturgess, and A. Vedlitz. 2013. “Knowledge, Risk, and Policy Support: Public Perceptions of Nuclear Power.” Energy Policy 62. Elsevier: 176–84. doi:10.1016/j.enpol.2013.06.098.

Stoutenborough, J. W., and A. Vedlitz. 2014. “The Effect of Perceived and Assessed Knowledge of Climate Change on Public Policy Concerns: An Empirical Comparison.” Environmental Science and Policy 37. Elsevier Ltd: 23–33. doi:10.1016/j.envsci.2013.08.002.

Ternes, T. A., A. Joss, and H. Siegrist. 2004. “Peer Reviewed: Scrutinizing Pharmaceuticals and Personal Care Products in Wastewater Treatment.” Environmental Science & Technology 38 (20): 392A – 399A. doi:10.1021/es040639t.

U.S. Census Bureau. 2010. “Pocatello City, Idaho Quick Facts.” Census of Population and Housing. http://www.census.gov/quickfacts/table/HSG010214/1664090.

Wachinger, G., O. Renn, C. Begg, and C. Kuhlicke. 2013. “The Risk Perception Paradox- Implications for Governance and Communication of Natural Hazards.” Risk Analysis 33 (6): 1049–65. doi:10.1111/j.1539-6924.2012.01942.x.

Welhan, J., C. Meehan, and T. Reid. 1996. “The Lower Portneuf River Valley Aquifer: A Geologic/hydrologic Model and Its Implications for Wellhead Protection Strategies.” In Final Report for the EPA Wellhead Protection Demonstration Project and the City of Pocatello Aquifer Geologic Characterization Project, 131.

119

4 Conclusions and Remarks

4.1 Summary of thesis work

4.1.1 Spatial patterns and sources of Nitrate-N

In this project, I identified hotspots of nitrate-N groundwater contamination in the

LPRV and identified the main sources of this nitrate contamination. I hypothesized that the main source of nitrate contamination was septic systems. The results of the triple isotope signatures, PPCP tracers, nitrate concentrations, and major anions indicated that septic systems were possible or likely sources of nitrate for about 2/3 of the wells sampled in 2015. This multi-tracer approach identified fertilizer sources for two wells.

Forty-one wells were definitively confirmed as septic-impacted. Of the 41 wells with clear septic sources, 11 had nitrate concentrations over the EPA MCL. Twenty-two of the remaining wells are likely septic-impacted, but there is the possibility of other sources.

There was insufficient information to define nitrate sources for the remaining 35 wells, but septic sources could not be rejected as a possibility for these wells.

I identified two main nitrate hotspots in the LPRV: one on the eastern side of the valley and another on the western side. When the nitrate concentration from wells located within both hotspots were compared to septic density, there was not an apparent relationship, regardless of the method by which septic density was estimated.

4.1.2 Public risk perceptions that drive action toward water quality improvement

In this study, I identified the main factors that influenced risk perceptions and action for modifying water quality at the residential scale. Public perceptions of drinking water influence water quality improvement measures on a residential level. The cities of

120

Chubbuck and Pocatello, Idaho are mid-sized cities whose development at exurban boundaries may affect drinking water supplies. Public risk perceptions have implications for policy and management, and are affected by historical water quality issues, a current nitrate problem for private water supplies, and the potential of increased water quality issues for city and private supplied drinking water. Therefore, I hypothesized that residents who had city-reported elevated nitrate would be more likely to utilize residential modification methods that improve water quality.

Within this study, I discovered that contrary to my hypothesis, measured water quality is not the major predictor of risk perceptions and actions that would improve water quality. Instead, three factors were important for those who improve their water quality: privately supplied water, concern associated with pharmaceutical contamination, and education. I found that modifying water content for taste was not driven by increased risk perceptions. I also saw that trust in city government did not increase or decrease risk perceptions.

4.1.3 Integrating public risk perceptions with measured water quality

Survey results indicated that residents using well water were significantly more likely to test their water quality. However, measured water quality reported to the public did not appear to alter risk perceptions in this study. Surveys were not conducted in some of the areas with high nitrate-N clustering associated with septic-impacted private drinking wells because they fell outside of the City limits. It would be useful to evaluate whether these residents are also unlikely to take action to improve their water quality.

121

4.2 Opportunities for future work

In the LPRV, the effects of seasonality on groundwater contamination and possible well capture zones need further analysis to understand the full implication of septic systems on drinking water. To accomplish this goal, a simple model of groundwater nitrate transport could be applied with software like IDRIS (Lasserrer et al. 1999) or, with more data, a complete and in-depth model could be applied using MT3D-MODFLOW

(Prommer et al. 2003). However, it would be ideal to identify heterogeneities in the flow system before creating a full transport model. Future studies should also include validation of septic density by using actual permitting records.

Survey respondents should be asked about their sensory perceptions of their water quality (e.g., taste and smell) to assess whether behaviors are driven primarily by risk perceptions or simply due to taste preferences. Future questions should also properly identify nitrate as a specific contaminant of concern, similar to the question asking specifically about pharmaceuticals. The questionnaire should also clarify the difference between treating and filtering of drinking water. It may also be beneficial to include in- person surveys concurrently with well sampling. This would allow a more direct analysis of measured water quality affecting risk perception and action.

Other areas of Idaho should be studied utilizing this same interdisciplinary approach. However, both the isotope and PPCP analyzes are expensive and require scientific expertise (Fenech et al. 2014). Therefore, future work could substitute microbiological identification of fecal indicator bacteria instead of isotopes and PPCPs.

This approach would give the same identification of manure/septic that isotopes provided for this study while reducing cost (Fenech et al. 2014). Fecal indicators would not be able

122 to differentiate sewage and manure contamination. However, if the study is within an area that clearly indicates septic, manure, or another source as the dominant nitrate source, then this should not be a problem. Future studies should use the decision-support tool created by Fenech et al. (2014), along with the results presented here as guidance to identify the best nitrate sourcing method for individual projects. Another useful decision- support tool that should be considered for future works is presented by Almasri (2006).

This framework includes modeling nitrate fate and transport in both the unsaturated and saturated zone, and management options for both environmental and economic aspects.

If pharmaceuticals are considered for future studies in the Idaho area, they should include sulfamethoxazole and sulfadimethoxine because they have been detected in drinking water within this study, as well as in Washington and Cassia Counties (Batt et al. 2006, Schorzman & Baldwin 2009). Sucralose was also found to occur frequently in this study and to my knowledge, has not been tested for elsewhere in Idaho. It should, therefore, also be considered for future analysis.

123

4.3 References

Almasri, Mohammad N. 2006. “Nitrate Contamination of Groundwater: A Conceptual Management Framework.” Environmental Impact Assessment Review 27 (3): 220– 42. doi:10.1016/j.eiar.2006.11.002.

Doria, M. F. 2006. “Bottled Water versus Tap Water: Understanding Consumer’s Preferences.” Journal of Water and Health 4 (2): 271–76. doi:10.2166/wh.2006.008.

Prommer, H., D. A. Barry, and C. Zheng. 2003. “MODFLOW/MT3DMS Based Reactive Multicomponent Transport Modeling.” Ground Water 41 (2): 247–57. doi:10.1111/j.1745-6584.2003.tb02588.x.

Schorzman, K., and J. Baldwin. 2009. “Nitrate and Emerging Contaminants Evaluation of Springdale, Idaho: Cassia County Nitrate Priority Area.” Ground Water Quality Technical Report 1-50. Boise.

124

5 Appendices:

5.1 Appendix A: PPCPs and Detection Limits

CAS No. Analyte Detection Limit (ng/L)

103-90-2 Acetaminophen 5.0 51022-70-9 Albuterol 7.5 58-08-2 Caffeine 25 298-46-4 Carbamazepine 1.0 56-75-7 Chloramphenicol 20 486-56-6 Cotinine 5.0 134-62-3 DEET 5.0 147-24-0 Diphenhydramine 5.0 76-57-3 Codeine 5.0 25812-30-0 Gemfibrozil 1.0 15687-27-1 Ibuprofen 5.0 22204-53-1 Naproxen 5.0 611-59-6 Paraxanthine (1,7) 50

56296-78-7 DimethylxanthineFluoxetine ) 8.0 56038-13-2 Sucralose 25 122-11-2 Sulfadimethoxine 1.0 57-68-1 Sulfamethazine 1.0 723-46-6 Sulfamethoxazole 2.5 467-15-2 Norcodeine 2.5 72-14-0 Sulfathiazole 5.0 93413-62-8 0-DesmethylVenlafaxine 5.0 101-20-2 Triclocarban 10 3380-34-5 Triclosan 50 738-70-5 Trimethoprim 5.0 81-81-2 Warfarin 25 93413-69-5 Venlafaxine 2.5

125

5.2 Appendix B: Clean Hand/Dirty Hand Jobs

Field sampling reduced the possibility of introducing contaminants by assigning tasks that are designated for one person as Clean Hands (CH) and another person as Dirty

Hands (DH). These tasks can sometimes overlap which means it can be handled by either

CH or DH as long as care is taken to prevent sample contamination.

Both CH and DH

 Wear disposable, powderless gloves during all sampling occupations while changing gloves frequently. CH tasks

Takes care of all equipment that comes into contact with the sample:

 Handles and labels sample bottles.  Handles sample discharge tubing.  Prepares the clean workspace inside the vehicle for sampling.  Filter samples. DH tasks

Takes care of all tasks that involve contact with potential sources of contamination.

 Handles all equipment used at multiple sampling sites: YSI equipment, flow cell, hoses and buckets for measuring discharge.  Handles all tools such as wrenches and pliers.  Handles solution for calibration and calibrates YSI equipment.  Records all field measurements.

126

5.3 Appendix C: Sampling and Cleaning Procedures

This appendix includes field sampling and lab processing protocols.

5.3.1 Pre-field preparation and cleaning

Before you start the pre-field bottle prep, check that you have:

500mL HDPE bottles 5 colors of sample tape Sharpie Ziploc bags 60mL HDPE bottles Gloves

Pre-label all 5 60-mL HDPE subsample bottles with DIFFERENT colors that are used consistently for each of the 5 bottles:

Nitrate: PVNnit.1 (red) Cations: PVNcat.1 (blue) Anions: PVNan.1 (orange) DIC: PVNdic.1 (green) Isotopes: PVNiso.1 (yellow)

For all 500mL field bottles:

 Bring to Lohse lab after coordinating with Kitty or lab manager about space/timing constraints.  Put on gloves.  Triple rinse each bottle with 18 mega-ohm DI water. Dump waste water.  Then fill with DI water and cap. Bring to geochem lab, so they are ready for field campaigns.

60mL subsample bottles (nitrate, anions, DIC, cation and isotopes:

 Bring to Lohse lab after coordinating with Kitty or lab manager about space/timing constraints.  Put on gloves.  Triple rinse each bottle with 18 mega-ohm DI water. Dump waste water.

127

 Then leach the sample bottles by filling with DI water and cap. Bring to geochem lab, so they are ready for post-field campaign filtering.  They should leach for at least 24 hours before sampling.

For the 1L pharmaceutical amber sample bottles:

 While wearing gloves, label half of the sample bottles sequentially, these will contain the post-field filtered samples. Start with PVRrx.1 and label to PVRrx.120  Avoid all possible sources of caffeine and pharmaceuticals 24 hours prior to this labeling process.  Do not open the bottles. The bottles arrive certified clean from the manufacturer and should not be leached or rinsed or opened.

Before you start the pre-field YSI calibration, check that you have:

YSI ProPlus Spare AA batteries (ideally Energizer industrial or Duracell ProCells) Small Phillip’s head screwdriver YSI multiparameter probe cable YSI EC/T probe YSI pH probe pH calibration solution (pH 7 and 10) EC calibration solution (EC>=1,000 uS/cm) Threaded calibration cap for YSI sensors

Calibrate the YSI pH and EC probes

 First, do a visual check for moisture, corrosion, threads, O-rings. Check ports to ensure that they aren’t blocked or building up any biofilms, etc.  If replacing any parts, first clean in a circular motion, parallel to the threads to remove dirt and moisture before replacing any parts (rubbing alcohol or canned air can help to dry things out, if needed).  Replace O rings every couple of years, and maintain with Crytox silicone seal (just a little bit needed)  When inserting instruments, ensure that pins are lined up and keep an eye on glass bulbs (if using) to avoid breakage. It is very important that you do no force the connection between the cable and instrument reader! The connection is very easy to break.  Store sensors in MOIST conditions (but not wet: KCl in pH probe will diffuse out too quickly and lessen its life, storing at pH 4 also an option) 1. If data set-up appears, ensure that date format MM/DD/YY is selected and set the correct date.

128

2. Temperature: small silver thermistor on probe. OK to rub/scrape off build-up. Can’t calibrate directly, but can check with an ice bath. If it’s off by >2C, we’ll send it in. This is usually not a problem. 3. Conductivity: this sensor has an extremely linear response, so we can calibrate at high conductivity sensitivity (look in manual for further explanation). 5.3.2 Field Sampling

Before heading to the field, check that you have: Not consumed caffeine, ibuprofen, or cigarettes (for pharma and N contamination) in the last 24 hrs.

Intended sampling locations for the day and back-up plans Field data sheets, indicating whether you’re on the number or letter team (if there are 2 teams) 1L pharm. sampling bottles 500mL multi-sample bottles Cooler Ice packs Ziploc bags Gloves powderless, nitrile Well tapes Plenty of tubing Garden hose (25 ft.) Y-connector Hose to tubing connectors Tubing connector reducers Tubing clamps Plastic for ground Tool kit: wrenches, Phillip’s head screwdriver (jeweler’s and regular size)

Bucket Cell phone Sampling partner ISU ID on you Sample t-shirt or hat? Field book with map, contact list Labels for bottles Sharpie Pen Flashlight YSI multiparameter probe ProPlus controller Extra AA batteries

Procedures

129

 Check to ensure site is proper sampling location based on list of intended samples for the day  Gain permission to sample at well o If permission cannot be gained, move on to back-up sampling location as identified on sampling list.  Connect a garden hose to the sampling point (outdoor spigot or a faucet on house exterior) and flush the lines for at least 10 minutes to purge the pump, plumbing, and at least some of the well casing.

 Begin to fill out sampling sheet!  If the YSI probe is not already attached to the flow-through cell, attach it and connect to the ProPlus controller.  Describe well and site conditions in field notes  Spread clean plastic sheet around well location to keep tubing and sampling equipment clean.  Put on gloves  Pull one amber 1L and one 500mL HDPE sample bottle from the cooler and label with site identification, sample designation, date and time. Format is site code (PV), sample type (N or R, for nutrients and Rx (pharmaceutical)), and YYMMDD.# (or .A/B/C/D, etc. if 2nd field team also sampling that day i.e. 1st field team will use YYMMDD.1 for the first sample and the second field team will use YYMMDD.A for the first sample).  Temporarily shut off the water flow, disconnect the garden hose and insert the sample tubing connections upstream of the garden hose as shown in figure below:

130

o Use hose to tubing connector a T-connector and reducers, as needed. This will be needed to slow discharge entering the flow cell and sampling lines. Always use unless there is an outdoor faucet available. o There must be no water-storage tanks, holding or pressurization tanks, or chemical disinfection or water softening systems upstream of the tap or faucet to which the sample tubing will be attached. If at all possible, connect to an outdoor pump spigot, such as where a garden house would be attached for lawn irrigation. o If an outdoor pump spigot connection is not available, check if there is a water faucet on the house exterior. If neither exists, the site should not be sampled. o Connect Tygon tubing from the T-connector to arm (a) of a Y-connector (place a tubing clamp on this arm to help control flow rate/pressure to the flow cell): (b)

(a) -<

(c)

131

 Connect a short length of Tygon tubing to arm (b) of the Y-connector and close it off with a tubing clamp. (This clamp can be reused between sampling sites.) You will eventually collect sampled water from this tubing.  Connect another short piece of tubing from arm (c) of the Y-connector to the inlet (bottom) port of the flow-through cell. This tubing can remain the same during the sampling campaign and does not need to be replaced because it does not come into contact with the water being sampled. o Protect the discharge end of sample tubing (b) from accidental contamination with a plastic end cap or black electrical tape prior to sampling. o The tubing on arm (b) and (c) should only be 12-18 inches long and protected from exposure to direct sunlight to minimize heating of the water being measured and sampled.  Connect another piece of tubing to the outlet (top) port of the flow-through cell (the waste discharge port). This tubing can remain the same during the sampling campaign and does not need to be replaced because it does not come into contact with the water being sampled.  Turn on the pump or faucet and continue flushing the pump, well casing and sample lines. o If necessary adjust the flow rate to the Y-connector with a hose clamp on the Tygon line. o Run the sample continuously until the YSI meter output reaches the following EC, pH, DO and T stability criteria:

o Record the pumping rate with a 5-gallon bucket and stopwatch with the water discharging from the garden hose outlet. o Let three times the equivalent volume of the flow cell and tubing run through the Y-connector and sampling system.  Dump DIW water from sampling bottles if they were pre-leached; this can also go down the drain or in the bucket with the discharge.

132

 Rinse 500mL sample bottles three times with sample water immediately after the above measurements have been recorded. Then fill the sample bottles completely, leaving no air/headspace in the top of the bottles. Triple-rinse a clean 1L sampling bottle prior to collecting the sample. Fill both bottles completely, leaving no air in the headspace. o Do not let the rate of flow exceed 500 mL/min while filling sample bottles; use the tubing clamp, if necessary but be careful to avoid air- bubble formation downstream of the constriction.  For quality control, collect one sample in triplicate for every 10th sample site. o Collect 1 extra sample in both the 500mL and 1L bottles following the same procedures above. Label these with the same code plus a “dup” on the end and note on the field notes that you collected a duplicate sample. o Also label one set of additional bottles with the same code plus a “mat” on the end. Then remove the cap and reseal this set of additional bottles, handling them in the same way that you handle the samples, but do not fill them in the field. You will fill these with DI water back in the lab and then filter them there according to normal procedures.  Disconnect all tubing and rinse the YSI probes and flowthrough cell with DI water.  Place used tubing and connectors in a DI bath.  Place both sample bottles in a large Ziploc bag and store in the cooler at 4°C in cooler as quickly as possible after collecting sample.  Before leaving site, check that o field sampling sheet is completely filled out o labels on bottles match the labels on the field sheet.  Place sample bottles in lab refrigerator and immediately begin filtering the samples on at a time.

133

5.3.3 Laboratory Processing of Samples

Subsamples from 500 mL bottles

Before pulling samples from refrigerator, check that you have:

Field data sheets Laboratory processing sheets Pre-processed Ziploc bag with 5 pre-labeled 60mL aliquot bottles 0.45um nylon filter (Whatman Puradisc 25mm 0.45um nonsterile) new 50mL syringe lab coat safety goggles Pen Nitrile gloves ultrapure nitric acid (Lohse lab providing during this pilot project)

In refrigerator: 500mL multisample bottles

Filtering Procedures

 Ensure that you have a clean workspace.  Prep all equipment and laboratory sheets before pulling 500 mL samples from refrigerator.  Begin to fill out laboratory sheet based on field sheet for the first samples you will be processing.  Put on gloves.  Open syringe and filter, and the bag of 5 pre-labeled 60mL bottles.  Dump out the leachate DI water from the 60mL bottle, but NOT the pre- conditioned nitric acid from the cation sample bottle.  Change gloves.  Suck in a 50mL sample from the wide-mouthed subsample bottle,  Attach filter to the end of the syringe  Place ~5mL of filtered sample into each 60mL sample bottle. Rinse the bottle and pour out the 5mL filtered sample. Repeat this three times total.  Then fill the 60mL bottle with filtered sample until there is no air/minimal headspace in the bottle. Cap it. o If your filter becomes clogged, replace it. Be sure to extrude a small amount of sampled water as waste through the filter before resuming the filtration.  Repeat with the remaining bottles EXCEPT with the cation bottle; leave enough space in this bottle to add 3 drops of the ultrapure nitric acid.  Place the used filter and syringe in the waste bin.

134

 Place the isotope samples in one Ziploc bag in the freezer, the nitrate and anion samples in another Ziploc bag, and finally place the cation and DIC samples in a 3rd Ziploc bag. All filtered samples, EXCEPT the DIC and Nutrients, should be placed immediately in the freezer.  The DIC and Nutrient samples should be dropped off at Kitty’s lab ASAP.

Pharmaceutical samples from 1L bottles:

Before pulling samples from refrigerator, check that you have:

Field data sheets

Laboratory processing sheets

0.7um glass fiber filter lab coat safety goggles Pen Nitrile gloves Clean 1L pharmaceutical HDPE bottles

In refrigerator: 1 L pharma field sampling bottles

Filtering Procedures

 If possible, caffeine consumption should be avoided the day before and the day of extractions so contamination of the samples does not occur. The use of hand creams, personal products or anti-bacterial soaps that may contain analytes on the target list should be avoided.  Ensure that you have a clean workspace.  Prep all equipment and laboratory sheets before pulling 1L samples from refrigerator.  Begin to fill out laboratory sheet based on field sheet for the first samples you will be processing.  Put on gloves.  Suck in a 50mL sample from the wide-mouthed subsample bottle,  Attach filter to the end of the syringe  Place ~5mL of filtered sample into each 60mL sample bottle. Rinse the bottle and pour out the 5mL filtered sample. Repeat this three times total.  Then fill the 1L bottle with filtered sample until there is no air/minimal headspace in the bottle. Cap it. o If your filter becomes clogged, replace it. Be sure to extrude a small amount of sampled water as waste through the filter before resuming the filtration.

135

 Repeat with the remaining bottles.  Place filtered sample bottle in Ziploc freezer bag.  Ensure that laboratory sheets are completely filled out.  Freeze water samples immediately after filtering.  Used bottles will be washed and stored for future projects that do not have the same low detection limits after sampling campaign is complete.

136

5.3.4 EXTRA Laboratory Preparation and Cleaning

*(for reference and post-field campaign bottle cleaning only) we are purchased clean bottles and tubing for all samples for this pilot project* References: USGS, EPA, American Public Health Association. 1995. Standard methods for the examination of water and wastewater. 19th ed. APHA, Washington, DC.

Equipment

Storage bags Aluminum foil 3-4 washbasins Soft brushes HCl acid Powderless nitrile gloves Nonphosphate laboratory solution

Procedures

 Cover cleaning area with clean plastic sheet.  Wear apron, glasses, and powderless gloves, switch pairs after each cleaning step.  Prepare nonphosphate laboratory solution of 0.1-0.2 percent (v/v) (if oily can use a higher concentration).  Prepare the acid solution, using a 0.1% v/v dilution of ACS trace-element-grade hydrochloric acid (HCl) in DIW. o Add the acid to water, not water to acid.  Label intended solution on washbasins and wash bottles.  Prepare pump and flow-through cell. Disassemble so that all parts are accessible for cleaning.  Change gloves  Clean washbasins and wash bottles o Fill washbasin and standpipes with 0.2 percent solution of nonphosphate laboratory-grade detergent. o Place wash bottles, scrub brushes and any other smaller items which will be used for cleaning equipment into the washbasin. Soak for 30 minutes. o Scrub exterior and interior basins and standpipes with soft scrub brushes. o Fill wash bottles with soapy solution and shake o Change gloves o Remove detergent from all items by rinsing with tap water o Change gloves o Rinse washbasins with DIW

137

o Pour 0.1% (v/v) HCl acid solution in washbasins, standpipes, and wash bottles. Let soak for 30 minutes. (exclude any items that have metal components) o Change gloves o Rinse with DIW o Change gloves  Place small equipment parts in washbasin labeled for detergent (include flowthrough cell). o Fill with a 0.1- to 2-percent solution of nonphosphate laboratory detergent. o Soak for 30 minutes, make sure tubing is filled with solution so it stays submerged o Scrub exterior and interior of equipment with soft brush, make sure to remove all residue and pay attention to grooves and hard to reach spots where particles can accumulate o Change gloves o Rinse all equipment with warm tap water o Change gloves  If equipment is made of glass or fluorocarbon polymer (other plastics) it needs to soak in the 0.1% HCl solution (exclude flowthrough) o Place the nonmetal equipment and tubing into the washbasin that was labeled earlier as HCl solution. o Fill basin with diluted HCl solution and let sit for 30 minutes o During the soaking process swirl acid solution several times to help the removal of encrusted minerals. o After 30 minutes, move to DIW basin. o Change gloves  Place the equipment in the washbasin labeled DIW (include flow-through) o Rinse exterior and interior with DIW and place on a clean surface for drying. o Rinse water is safe to pour down drain. o Change gloves  When dry, place into sealable storage bags or other transport container to minimize contamination.

138

5.4 Appendix D: Total dissolved solids, anions, and nitrate-N analyses

Below are the histograms used to define quantile thresholds (75%) defining septic- impacted wells. Also, included are some temporal anion figures not included in the thesis body.

5.4.1 Total dissolved solids (TDS) and nitrate-N

TDS (mg/L) Nitrate-N (mg/L)

TDS (mg/L) Nitrate-N (mg/L)

Mean 432.61 Mean 3.36 Std Dev 184.72 Std Dev 4.73 Std Err Mean 18.57 Std Err Mean 0.47 Upper 95% Mean 469.46 Upper 95% Mean 4.30 Lower 95% Mean 395.77 Lower 95% Mean 2.42 N 99 N 100

139

5.4.2 Anions

Chloride (mg/L) Sulfate (mg/L)

Mean 102.27 Mean 47.44

Std Dev 132.86 Std Dev 44.07

Std Err Mean 13.29 Std Err Mean 4.41

Upper 95% Mean 128.63 Upper 95% Mean 56.19

Lower 95% Mean 75.91 Lower 95% Mean 38.70

N 100 N 100

140

The following shows temporal changes of cations measured in at least three different time periods. Each code refers to a well with prefixes MA referring to the main aquifer, WB referring to the West Bench, EA the eastern aquifer, and MC to Mink Creek. Few consistent patterns emerge from these wells. Spatial variability appears to exceed temporal trends in most cases.

141

142

143

5.5 Appendix E: Cation samples

Cations were not analyzed for samples taken for the 2015 sampling campaign, however they were collected and prepared.  Samples reserved for cation analysis were put in 60 mL HDPE bottles. The cation bottles were filled with filtered sample water, then two drops of ultrapure nitric acid were added, and the bottles were placed in the freezer until ready for analysis. 5.6 Appendix F: Septic-impacted wells

The following include the sample IDs for wells that had nitrate sources identified as mixed sources, but later identified as septic-impacted because they had nitrate- N concentrations over 4 mg/L. o Soil N and manure/septic: PVNnit.5, PVNnit.8, PVNnit.11, PVNnit.49, PVNnit.59, PVNnit.79, PVNnit.81. o Soil N, Manure/Septic, and Fertilizer: PVNnit.82 o There are 18 wells with high TDS/high nitrate, and of these wells, three (nit.53, nit.80, nit.82) have not been identified as septic-impacted. Only one well had high TDS/low nitrate, suggesting denitrification occurred, and this was the only indication that denitrification altered nitrate - 2- concentrations. Cl and SO4 are used as septic tracers because high concentrations are believed to come from water softeners. There were 16 wells that had sources of nitrate including ammonium from fertilizer/precipitation, soil N, and manure septic, and also had nitrate that increased with increase Cl-. Of the 16 wells, 5 (nit.6, nit.23, nit.67, nit.73, nit.99)

144

5.7 Appendix G: Additional information about groundwater, snowpack, and rainfall isotopic signatures

The following figure shows the isotopic signatures for rainfall, snowpack, and groundwater samples shown in red, green, and blue, respectively.

Mink Creek/Gibson-Jack Southern/Middle Aquifer

Landf Landf Landfill Northern Aquifer

145

5.8 Appendix H: Survey Questions

146

147

148

149

5.9 Appendix I: 2015 Sampling Results

5.9.1 2015 Results for nitrate-N, DIC, isotopes, anions

18 15 2 18 Sampling Id Previous Location NO3-N DIC ONO3 NNO3 HH2O OH2O Chloride Sulfate Sample Date sample Id (mg/L) (mg/L) (‰) (‰) (‰) (‰) (mg/L) (mg/L) PVNnit.1 Mink Creek 0.32 64.17 -7.94 7.20 -125.78 -16.68 29.90 14.68 5/12/2015 PVNnit.2 MC-121 Mink Creek 4.80 62.66 -127.70 -16.69 336.64 69.20 5/12/2015 PVNnit.3 West Bench 0.28 43.82 -127.62 -16.92 22.03 7.86 5/14/2015 PVNnit.4 West Bench 0.23 51.72 -135.29 -18.01 49.31 19.27 5/14/2015 PVNnit.5 EA-02 East Aquifer 16.38 67.26 -5.88 6.63 -128.55 -16.79 385.45 121.00 5/18/2015 PVNnit.6 EA-03 East Bench 16.37 42.13 -5.97 7.87 -131.72 -17.06 323.81 132.91 5/18/2015 PVNnit.7 East Bench 0.37 70.16 -142.61 -18.48 93.15 93.02 5/18/2015 PVNnit.8 East Aquifer 10.23 54.46 -5.85 7.03 -130.09 -16.89 241.50 74.08 5/18/2015 PVNnit.9 East Bench 1.32 57.50 -135.92 -17.83 296.66 106.67 5/19/2015 PVNnit.10 East Bench 1.63 49.67 -8.24 3.38 -137.73 -17.82 163.46 61.69 5/19/2015 PVNnit.11 East Bench 5.85 32.56 -10.35 5.83 -133.64 -16.93 945.67 255.56 5/19/2015 PVNnit.12 East Bench 10.55 39.56 -7.84 6.39 -134.71 -17.23 339.29 181.84 5/19/2015 PVNnit.13 Main 2.97 74.51 -5.96 6.37 -126.84 -16.53 39.49 28.49 5/20/2015 Aquifer PVNnit.14 West Bench 0.49 29.04 -132.93 -17.53 31.87 14.45 5/20/2015 PVNnit.15 West Bench 1.90 38.00 -3.35 3.17 -130.43 -16.73 77.13 29.60 5/22/2015 PVNnit.16 MA-67 Main 2.45 71.33 -4.95 4.20 -126.09 -16.33 62.60 42.49 5/22/2015 Aquifer PVNnit.17 WB-20 West Bench 0.43 26.87 -130.79 -16.91 14.10 6.08 5/22/2015 PVNnit.18 WB-12 West Bench 0.95 44.36 -125.23 -15.82 23.78 10.44 5/22/2015 PVNnit.19 MC-02 Mink Creek 0.57 34.26 -129.62 -16.94 28.57 9.48 5/22/2015

150

18 15 2 18 Sampling Id Previous Location NO3-N DIC ONO3 NNO3 HH2O OH2O Chloride Sulfate Sample Date sample Id (mg/L) (mg/L) (‰) (‰) (‰) (‰) (mg/L) (mg/L) PVNnit.20 Main 1.90 68.10 -4.24 4.20 -128.54 -17.25 39.01 33.33 5/23/2015 Aquifer PVNnit.21 RGI-05 Main 1.86 66.99 -3.64 2.37 -124.21 -15.96 42.62 40.39 5/23/2015 Aquifer PVNnit.22 Main 1.71 66.37 -4.43 1.86 -124.47 -16.15 41.93 22.94 5/25/2015 Aquifer PVNnit.23 MA-05 Main 1.40 66.60 -6.38 5.12 -123.58 -15.76 111.14 21.96 5/25/2015 Aquifer PVNnit.24 MC-103 Mink Creek 1.51 46.27 -128.32 -15.97 59.95 45.44 5/25/2015 PVNnit.25 MA-108 Mink Creek 2.12 70.77 -5.58 -0.75 -123.96 -16.03 136.78 49.34 5/25/2015 PVNnit.26 Mink Creek 3.83 71.17 -7.77 2.22 -124.87 -16.03 18.97 5.26 5/26/2015 PVNnit.27 MC-128 Mink Creek 0.24 39.77 -127.06 -16.02 52.24 23.39 5/26/2015 PVNnit.28 Mink Creek 2.08 63.55 -6.36 8.26 -122.40 -15.62 32.98 14.29 5/26/2015 PVNnit.29 Mink Creek 0.83 31.94 -127.47 -15.90 88.72 43.63 5/26/2015 PVNnit.30 MC-114 Mink Creek 1.87 52.09 -4.83 3.91 -129.42 -16.57 49.65 22.99 5/26/2015 PVNnit.31 Mink Creek 1.99 66.91 -127.95 -17.47 261.56 120.47 5/26/2015 PVNnit.32 Mink Creek 11.24 69.84 -7.64 4.77 -129.87 -17.24 246.26 85.41 5/27/2015 PVNnit.33 MC-125 Mink Creek 18.18 77.37 -5.79 7.58 -129.35 -17.02 136.12 51.30 5/27/2015 PVNnit.34 Mink Creek 15.97 75.22 -9.99 1.04 -132.70 -17.82 127.68 41.23 5/27/2015 PVNnit.35 MC-117 Mink Creek 3.95 60.44 -5.46 4.46 -129.53 -17.69 41.55 20.55 5/28/2015 PVNnit.36 Mink Creek 1.65 62.34 -1.29 15.34 -128.50 -17.37 38.30 32.12 5/28/2015 PVNnit.37 Main 1.78 68.92 -2.06 0.99 -128.72 -16.97 39.63 36.22 5/28/2015 Aquifer PVNnit.38 RGI-13 Main 1.94 67.93 -4.92 1.95 -128.29 -17.79 43.18 37.78 5/28/2015 Aquifer PVNnit.39 RGI-14 Main 2.08 68.25 -4.60 3.66 -130.52 -17.91 126.56 54.06 5/28/2015 Aquifer

151

18 15 2 18 Sampling Id Previous Location NO3-N DIC ONO3 NNO3 HH2O OH2O Chloride Sulfate Sample Date sample Id (mg/L) (mg/L) (‰) (‰) (‰) (‰) (mg/L) (mg/L) PVNnit.40 RGI-18 Main 2.10 69.30 -6.49 2.62 -126.88 -16.44 285.18 37.44 5/28/2015 Aquifer PVNnit.41 MA-06 Mink Creek 2.75 82.87 -4.91 5.98 -129.56 -17.79 30.51 16.14 5/29/2015 PVNnit.42 Mink Creek 18.44 62.61 -6.58 4.95 -135.13 -18.28 28.31 17.21 5/29/2015 PVNnit.43 MC-130 Mink Creek 0.07 70.64 -128.97 -17.41 12.71 32.55 5/29/2015 PVNnit.44 Mink Creek 0.20 91.51 -130.47 -17.58 239.27 181.40 5/29/2015 PVNnit.45 Mink Creek 0.19 51.45 -138.06 -18.74 281.59 21.63 5/29/2015 PVNnit.46 Mink Creek 6.93 58.84 -5.52 9.47 -133.83 -17.69 13.86 18.20 5/29/2015 PVNnit.47 Mink Creek 22.98 97.26 -127.81 -16.81 88.97 50.87 5/29/2015 PVNnit.48 Mink Creek 0.00 47.36 -136.45 -17.98 92.04 119.88 6/1/2015 PVNnit.49 Mink Creek 11.86 66.13 -7.37 8.07 -127.51 -16.46 21.03 16.51 6/1/2015 PVNnit.50 Mink Creek 1.62 59.33 -4.46 1.64 -128.29 -16.88 45.38 18.59 6/1/2015 PVNnit.51 Mink Creek 1.06 47.98 -130.60 -17.22 258.95 91.31 6/1/2015 PVNnit.52 West Bench 1.24 45.77 -126.84 -16.86 75.57 32.98 6/1/2015 PVNnit.53 West Bench 2.87 50.37 -133.37 -17.59 172.58 60.21 6/1/2015 PVNnit.54 WB-11 West Bench 0.42 37.42 -130.56 -16.60 62.08 77.16 6/2/2015 PVNnit.55 West Bench 3.15 32.65 -6.85 3.17 -133.29 -17.47 61.46 26.79 6/2/2015 PVNnit.56 Mink Creek 0.70 56.39 -132.50 -17.53 18.33 9.50 1/0/1900 PVNnit.57 West Bench 0.00 55.81 -129.22 -17.29 80.14 58.58 6/3/2015 PVNnit.58 West Bench 0.24 47.39 -126.67 -17.24 133.58 91.34 6/3/2015 PVNnit.59 West Bench 4.94 35.92 -3.84 7.10 -130.44 -17.08 406.34 146.46 6/3/2015 PVNnit.60 West Bench 3.06 47.93 -134.18 -17.78 34.62 31.35 6/3/2015 PVNnit.61 West Bench 6.37 48.43 -8.28 5.84 -131.55 -17.00 41.84 36.27 6/3/2015 PVNnit.62 MA-54 Main 1.73 66.93 -124.59 -16.43 42.59 36.18 6/4/2015 Aquifer PVNnit.63 Main 1.62 69.09 -124.66 -16.28 11.14 10.35 6/4/2015 Aquifer

152

18 15 2 18 Sampling Id Previous Location NO3-N DIC ONO3 NNO3 HH2O OH2O Chloride Sulfate Sample Date sample Id (mg/L) (mg/L) (‰) (‰) (‰) (‰) (mg/L) (mg/L) PVNnit.64 MA-55 Main 1.66 69.55 -5.51 4.32 -124.77 -16.45 42.14 35.57 6/4/2015 Aquifer PVNnit.65 MA-42 Main 1.73 68.48 -125.30 -16.56 41.64 35.72 6/4/2015 Aquifer PVNnit.66 MA-43 Main 1.67 67.65 -6.73 3.27 -125.00 -16.50 17.14 8.98 6/4/2015 Aquifer PVNnit.67 MA-31 Main 2.01 72.43 -6.68 6.45 -121.89 -15.21 57.49 43.37 6/4/2015 Aquifer PVNnit.68 WB-08 West Bench 0.95 47.10 -5.31 2.13 -125.27 -16.77 22.67 14.38 6/4/2015

PVNnit.69 Main 2.40 73.59 -7.65 3.10 -124.33 -16.50 43.05 36.05 6/4/2015 Aquifer PVNnit.70 MA-115 Main 0.59 50.29 -131.11 -17.42 13.98 8.82 6/5/2015 Aquifer PVNnit.71 MA-32 Main 2.47 74.33 -5.19 3.35 -124.21 -16.18 44.52 36.80 6/5/2015 Aquifer PVNnit.72 West Bench 0.43 23.58 -131.39 -17.52 18.63 7.27 6/5/2015 PVNnit.73 Main 1.92 70.24 -4.25 4.76 -126.06 -16.64 59.55 42.72 6/5/2015 Aquifer PVNnit.74 West Bench 0.47 25.76 -130.88 -17.39 40.91 34.37 6/8/2015 PVNnit.75 Main 2.43 73.47 -5.17 3.04 -126.23 -16.61 19.13 9.85 6/8/2015 Aquifer PVNnit.76 Main 1.88 69.79 -126.59 -16.66 513.58 137.83 6/8/2015 Aquifer PVNnit.77 West Bench 0.65 28.96 -128.32 -16.93 174.04 57.85 6/8/2015 PVNnit.78 East Bench 2.79 59.83 -11.33 5.44 -135.43 -17.83 107.15 74.53 6/9/2015 PVNnit.79 East Aquifer 18.35 44.79 -5.77 7.72 -131.43 -17.13 126.63 53.06 6/9/2015 PVNnit.80 East Bench 2.96 35.66 -131.80 -16.88 71.54 74.01 6/9/2015 PVNnit.81 East Aquifer 9.18 37.02 -5.15 5.48 -135.07 -17.62 61.64 44.51 6/9/2015

153

18 15 2 18 Sampling Id Previous Location NO3-N DIC ONO3 NNO3 HH2O OH2O Chloride Sulfate Sample Date sample Id (mg/L) (mg/L) (‰) (‰) (‰) (‰) (mg/L) (mg/L) PVNnit.82 East Bench 4.33 41.86 -4.10 2.73 -126.82 -15.44 23.85 18.11 6/9/2015 PVNnit.83 Main 2.40 68.85 -126.45 -16.26 53.30 40.85 6/10/2015 Aquifer

PVNnit.84 Main 2.24 70.15 -126.18 -16.41 2.81 2.68 6/10/2015 Aquifer PVNnit.85 Main 2.22 71.09 -126.26 -16.36 134.83 52.79 6/10/2015 Aquifer PVNnit.86 WB-23 West Bench 0.23 13.18 -131.71 -17.82 32.65 15.62 6/10/2015 PVNnit.87 DEQ-114 West Bench 1.61 32.04 -132.12 -17.16 34.44 16.42 6/10/2015 PVNnit.88 West Bench 0.10 62.84 -143.51 -18.99 38.00 12.77 6/11/2015 PVNnit.89 West Bench 0.00 58.46 -142.05 -18.77 39.48 30.44 6/11/2015 PVNnit.90 West Bench 0.00 61.88 -142.43 -18.80 83.51 50.28 6/11/2015 PVNnit.91 Main 1.89 66.30 -126.32 -16.53 179.55 150.39 6/16/2015 Aquifer PVNnit.92 Main 2.51 66.26 -3.06 2.84 -126.17 -16.40 79.78 63.37 6/16/2015 Aquifer PVNnit.93 East Bench 0.00 56.99 -130.86 -15.84 39.90 32.31 6/16/2015 PVNnit.94 Main 2.24 69.05 -123.10 -15.94 44.35 36.91 6/17/2015 Aquifer PVNnit.95 MA-33 Main 2.53 71.90 -5.22 4.00 -123.85 -15.69 60.47 43.07 6/17/2015 Aquifer PVNnit.96 Main 1.21 66.60 -123.27 -16.11 44.25 36.23 6/17/2015 Aquifer PVNnit.97 East Bench 2.71 69.85 -4.44 -2.05 -123.98 -15.89 37.62 27.68 6/17/2015 PVNnit.98 Main 1.58 66.35 -6.35 4.54 -124.01 -15.97 45.27 38.75 6/18/2015 Aquifer

154

18 15 2 18 Nitrate_ID Legacy_Id Location NO3-N DIC ONO3 NNO3 HH2O OH2O Chloride Sulfate Date (mg/L) (mg/L) (‰) (‰) (‰) (‰) (mg/L) (mg/L) PVNnit.99 Main 1.54 64.11 -6.57 6.42 -124.88 -16.18 0.00 0.00 6/18/2015 Aquifer PVNnit.100 Main 1.46 66.57 -122.90 -15.72 0.00 0.00 6/18/2015 Aquifer

5.9.2 2015 Results for PPCPS

RX_ALT Carbamazepine DEE Diphen- Codeine Ibuprofen Fluoxetine Sucralose Sulfadimethoxine Sulfamethoxazole ID (ng/L) T hydramine (ng/L) (ng/L) (ng/L) (ng/L) (ng/L) (ng/L) (ng/L) (ng/L) 19 9.30 1.00 2.30 14.00 20 38.00 12 17.00 115.00 98.00 8 1.00 134 14.00 19.00 7 29.00 144 17.00 126 16.00 2.00 177 33.00 14 45.00 22.00 7 8.50 88 6.10 180 39.00 41 18.00 53 104.00 14.00 186 13.00 155.00 152.00 27.00 78.00

155

RX_ALT Carbamazepine DEE Diphen- Codeine Ibuprofen Fluoxetine Sucralose Sulfadimethoxine Sulfamethoxazole ID (ng/L) T hydramine (ng/L) (ng/L) (ng/L) (ng/L) (ng/L) (ng/L) (ng/L) (ng/L) 156 44.00 74.00 126 2.00 90 62.00 62 64.00 6.00 39 56.00 174 656.00 35 618.00 181 37.00 163 57.00 56 255.00 21 36.00 182 9.70 37.00 97 53.00 47 19.00

156