<<

An Investigation of Fertilizer-Derived Uranium in Ohio Agricultural Soils

Thesis

Presented in Partial Fulfillment of the Requirements for the Degree Master of Science in

the Graduate School of The Ohio State University

By

Adolfo Eberhard Calero, B.S.

Graduate Program in Earth Sciences

The Ohio State University

2020

Thesis Committee

Dr. W. Berry Lyons, Advisor

Dr. Nicholas T. Basta

Dr. Rattan Lal

1

Copyrighted by

Adolfo Eberhard Calero

2020

2

Abstract

The evolution of modern agricultural practices as a result of a growing global population has led to increases in the flux of agricultural chemicals and compounds to both natural waters and soils. The most studied of these impacts are the addition of nitrogen (N) and phosphorus (P) through the application of chemical fertilizers, leading to their increase in surface runoff from agricultural operations, causing eutrophication in natural waters.

Uranium (U) is a trace element that is often associated with P-rich fertilizer use, having the potential to accumulate in soils, crops, and surface waters, due to its natural occurrence in phosphate rock, along with various weathering and erosional processes that make it mobile. Studies have suggested phosphorus fertilizer is a source of uranium contamination in natural waters and soils. This study investigates chemistry of agricultural soils treated with P-fertilizer and compares them with the chemistry of agricultural soils under different management practices. This study also characterizes historical Geographic Information System (GIS) data in order to correlate uranium concentrations in Ohio soils with Ohio’s glaciated compared to non-glaciated surfaces, underlying bedrock geology, and current land use/land cover. Total U concentrations in the soils ranged from 2.7 – 5.4 µg/g, and the water-, base-, and acid-soluble concentrations of U ranged from 2.2 x 10-4 – 1.4 x 10-3, 0.09 – 0.58, and 0.27 – 0.76 µg/g, respectively. The average concentration of the sum of the U in each of the

i soluble/extractable phases gave a maximum value of 18% of the total U. Using previously published data from the USGS, there are higher U concentrations in soils from the glaciated portion of the State, as opposed to those from the unglaciated portion of

Ohio. This research clearly demonstrates that soils having fertilizer application show no statistical difference in U compared to similar soils without fertilizer application.

ii

Acknowledgments

I would like to express my sincere appreciation to Dr. Berry Lyons for his infinite patience, encouragement, and for giving me the opportunity to perform research in both

Ohio and Antarctica. I would also like to extend my deepest gratitude to Melisa Diaz for being an amazing friend, scientist, and for providing me with the opportunity to communicate my research. I would like to thank Dr. Nick Basta, for being on my committee, for helping deepen my understanding of soils and soil chemistry, and for running dichromate analyses on my samples. I would also like to thank Dr. Rattan Lal for being on my committee, and for providing me access to his plots at Waterman Farm. I am extremely grateful to Dr. Chris Gardner for all of his advice regarding statistics and data presentation, as well as for bringing me to help with research in Antarctica. Thank you to

Dr. Sue Welch for helping me with SEM analysis and for providing me with advice on any questions I had about lab work. I would also like to thank Dr. Steve Culman and

Bethany Herman for providing me with soil samples, as well as answers to any questions

I had about them. Thank you to Devin Smith for spending a couple of early mornings with me at Waterman Farm to help with sample collection. Thanks to the Trace Element

Research Lab, especially Anthony Lutton, for helping me with the analyses of my samples. Thank you to Dr. Anne Carey for the use of her lab space. Thanks to Nall

Moonilall for providing me with the locations and with all the information regarding the

iii plots at Waterman Farm. Thanks to Brianna Piergallini for helping me with my GIS maps. I would also like to thank the Geological Society of America and the Friends of

Orton Hall for providing me with funding for this thesis work. Thanks to Dr. Seth Young for encouraging me to pursue a graduate degree and to work in a great research program.

Last, I would like to thank my family for their love and support throughout my graduate experience.

iv

Vita

September 15th, 1996……………………………..Born – Miami, Florida August 2018………………………………………B.S. Geology, Florida State University August 2018 to Present…………………………..Graduate Research and Teaching Associate, School of Earth Sciences, The Ohio State University

Fields of Study

Major Field: Earth Sciences

v

Table of Contents

Abstract ...... i Acknowledgments...... iii Vita ...... v List of Tables ...... viii List of Figures ...... xiii Chapter 1. Introduction ...... 1 1a. U in Soils; General Considerations ...... 3 1b. Previous Work ...... 6 1c. Potential Human Impacts ...... 8 1d. Rationale for Research ...... 9 1e. Hypothesis ...... 10 Chapter 2. Study Areas ...... 11 2a. Waterman Farm ...... 11 2b. “Wooster” Sites ...... 12 2c. Coshocton ...... 13 Chapter 3. Methods ...... 15 3a. Cleaning Plasticware ...... 15 3b. Sampling and Processing ...... 15 3c. Extractions...... 16 3d. U-Analysis ...... 19 3e. Precision and Accuracy ...... 20 3f. pH Analysis ...... 20 3g. Loss on Ignition ...... 21 3h. XRF Analysis ...... 22 3i. GIS Maps ...... 23 3j. Scanning Electron Microscopy ...... 24 Chapter 4. Results ...... 25 4a. USGS and NURE Ohio GIS Map Observations ...... 25 4b. Loss on Ignition for Samples ...... 28

vi

4c. pH Values for Samples...... 29 4d. XRF Data – Impact of Weathering on U Concentrations in Soils ...... 30 4e. Uranium Concentrations in Soils ...... 32 4f. Uranium-Bearing Minerals in Soils ...... 39 Chapter 5. Discussion ...... 41 5a. Relationship of LOI and pH to U Concentrations in Soils...... 41 5b. Extractions – Water vs. Base vs. Acid ...... 42 5c. U Concentrations in Ohio Soils: Based on a Datamining and GIS Approach ...... 44 5d. Comparison of Agricultural Management Practices in Relation to U Concentrations in Agricultural Soils ...... 47 5e. Comparison of Ohio U Soil Chemistry to US and Global Inventories ...... 52 Chapter 6. Conclusions ...... 54 6a. Summary of Research ...... 54 6b. Future Work ...... 55 Works Cited ...... 57 Appendix A: General Tables and Figures ...... 66 Appendix B: pH values for samples ...... 126 Appendix C: Loss On Ignition Values ...... 130 Appendix D: U Concentrations ...... 135

vii

List of Tables

Table A-1 Loss on Ignition (LOI) data provided by SGS Canada ...... 111

Table A-2 Si, Al, Fe, and Mg Oxide data provided by SGS Canada ...... 112

Table A-3 Ca, K, Na, and Ti Oxide data provided by SGS Canada ...... 113

Table A-4 Mn, P, Cr, and V Oxide data provided by SGS Canada ...... 114

Table A-5 Calculated Chemical Index of Alteration (CIA) values for samples provided to SGS Canada ...... 115

Table A-6 Concentrations of standards used for ICP-MS analysis of 10% HCl and 0.1M sodium bicarbonate leaches...... 116

Table A-7 Concentrations of standards used for ICP-MS analysis of DI water leaches 116

Table A-8 Check Standard concentrations and RSD calculations for HCl and Bicarbonate extraction ICP-MS measurement precision...... 117

Table A-9 Check Standard concentrations and RSD calculations for DI water extraction ICP-MS measurement precision...... 118

Table A-10 Comparison of values from three different instances of organic matter determination...... 119

Table A-11 List of minerals identified through SEM analyses...... 120

Table A-12 Correlations between U concentrations for non-sequentially leached samples from Waterman Farm with LOI and pH...... 120

Table A-13 Correlations between U concentrations for sequentially leached samples from Waterman Farm with LOI and pH...... 120

Table A-14 Correlations between U concentrations from sequential extractions to total U (obtained from XRF) and calculated CIA values...... 121

viii

Table A-15 Approximations of mineral to grain sizes from SEM images...... 122

Table A-16 Ohio geologic units associated with the NURE and USGS datasets ...... 123

Table A-17 The grouping of original land use classifications into their simplified categories for the agricultural, developed and water land use categories ...... 124

Table A-18 The grouping of original land use classifications into their simplified categories for the undeveloped land use category ...... 125

Table B-1 pH values for Manure plots at Waterman Farm...... 126

Table B-2 pH values for Cover Crop plots at Waterman Farm...... 126

Table B-3 pH values for Fertilized plots at Waterman Farm...... 127

Table B-4 pH values for Fallow plots at Waterman Farm...... 127

Table B-5 pH values for the Coshocton site...... 128

Table B-6 pH values for the East Badger site...... 128

Table B-7 pH values for the Western site...... 129

Table B-8 pH values for the Northwest site...... 129

Table C- 1 LOI values for Cover Crop plots at Waterman Farm...... 130

Table C- 2 LOI values for Manure plots at Waterman Farm...... 130

Table C- 3 LOI values for Fallow plots at Waterman Farm...... 131

Table C- 4 LOI values for Fertilizer plots at Waterman Farm...... 131

Table C- 5 LOI values for the Northwest site...... 132

Table C- 6 LOI values for the East Badger site...... 133

Table C- 7 LOI values for the Western site...... 133

Table C- 8 LOI values for the Coshocton site...... 134

Table D-1 Total U data provided by SGS Canada...... 135 ix

Table D-2 U concentrations and statistics for 10% HCl extractions for Cover Crop plots, Waterman Farm...... 136

Table D-3 U concentrations and statistics for 10% HCl extractions for Manure plots, Waterman Farm...... 137

Table D-4 U concentrations and statistics for 10% HCl extractions for Fallow plots, Waterman Farm...... 137

Table D-5 U concentrations and statistics for 10% HCl extractions for Fertilizer plots, Waterman Farm...... 138

Table D-6 U concentrations and statistics for 10% HCl extractions for the East Badger site...... 138

Table D-7 U concentrations and statistics for 10% HCl extractions for the Western location...... 139

Table D-8 U concentrations and statistics for 10% HCl extractions for the Northwest location...... 139

Table D-9 U concentrations and statistics for 10% HCl extractions for Coshocton...... 139

Table D-10 U concentrations for DI Water extractions, Cover Crop plots, Waterman Farm...... 140

Table D-11 U concentrations for DI Water extractions, Manure plots, Waterman Farm...... 140

Table D-12 U concentrations and statistics for DI Water extractions, Fallow plots, Waterman Farm...... 140

Table D-13 U concentrations and statistics for DI Water extractions, Fertilizer plots, Waterman Farm...... 141

Table D-14 U concentrations for DI Water extractions, East Badger site...... 141

Table D-15 U concentrations for DI Water extractions, Western site...... 141

Table D-16 U concentrations for DI Water extractions, Northwest site...... 141

Table D-17 U concentrations for 0.1 M sodium bicarbonate extractions, Cover Crop plots, Waterman Farm...... 142 x

Table D-18 U concentrations for 0.1 M sodium bicarbonate extractions, Manure plots, Waterman Farm...... 142

Table D-19 U concentrations and statistics for 0.1 M sodium bicarbonate extractions, Fallow plots, Waterman Farm...... 142

Table D-20 U concentrations and statistics for 0.1 M sodium bicarbonate extractions, Fertilizer plots, Waterman Farm...... 143

Table D-21 U concentrations for 0.1 M sodium bicarbonate extractions, East Badger site...... 144

Table D-22 U concentrations for 0.1 M sodium bicarbonate extractions, Western site. 144

Table D-23 U concentrations for 0.1 M sodium bicarbonate extractions, Northwest site...... 144

Table D-24 U concentrations for 10% HCl sequential extractions, Cover Crop plots, Waterman Farm...... 144

Table D-25 U concentrations for 10% HCl sequential extractions, Manure plots, Waterman Farm...... 145

Table D-26 U concentrations for 10% HCl sequential extractions, Fallow plots, Waterman Farm...... 145

Table D-27 U concentrations for 10% HCl sequential extractions, Fertilized plots, Waterman Farm...... 145

Table D-28 U concentrations for 10% HCl sequential extractions, East Badger site..... 146

Table D-29 U concentrations for 10% HCl sequential extractions, Western site...... 146

Table D-30 U concentrations for 10% HCl sequential extractions, Northwest site...... 146

Table D-31 Comparison of U concentrations from sequential extractions to total U concentrations from XRF analyses...... 147

Table D-32 Percentage of total U leached from each sample through sequential extractions...... 147

Table D-33 Mean and Median comparisons for Waterman Farm extraction values ...... 148

xi

Table D-34 Range of each sequential extraction at Waterman and Wooster sites, compared to the amount of total U extracted...... 148

Table D-35 Comparison of totals for U concentration data for the NURE and USGS datasets in Ohio...... 148

Table D-36 Comparison of U concentration data from the NURE and USGS datasets, as well as Barnes, 2020, et al. in the glaciated and unglaciated areas of Ohio...... 149

Table D-37 Comparison of data from visually identified points from the USGS Top 5cm data set in regard to their land cover...... 149

Table D-38 Comparison of U concentrations per geologic unit and dataset in Ohio. .... 150

Table D-39 Southwestern Ohio glacial till U data (Barnes, et al., 2020)...... 150

Table D-40 U concentrations in different types of P-mineral fertilizers from their respective countries, adapted from Kratz, et al. (2008) ...... 151

Table D-41 U concentrations in different types of P-containing mineral compound fertilizers from their respective countries, adapted from Kratz, et al. (2008) ...... 152

xii

List of Figures

Figure 1. Field at Waterman Farm showing plots for Fallow, Compost, Manure, and Cover Crop management practices...... 66

Figure 2 Field at Waterman Farm showing plots with different rates of Fertilizer and Mulch application...... 67

Figure 3 East Badger Field in Wayne County, OH (Fulford & Culman, 2018) ...... 68

Figure 4 Western Agricultural Research Station Field in Clark County, OH (Fulford & Culman, 2018)...... 69

Figure 5 Northwest Agricultural Research Station in Wood County, OH (Fulford & Culman, 2018)...... 70

Figure 6 U concentrations from Coshocton samples...... 71

Figure 7 Pie charts showing the distribution of total U in samples from Waterman Farm...... 72

Figure 8 Pie charts showing the distribution of total U in samples from the East Badger location in Wayne County, OH...... 73

Figure 9 Pie charts showing the distribution of total U in samples from the Northwest location in Wood County, OH...... 74

Figure 10 Pie charts showing the distribution of total U in samples from the Western location in Clark County, OH...... 75

Figure 11 Box plot showing U concentrations at Waterman Farm from samples that were leached using only 10% HCl...... 76

Figure 12 Box plot showing U concentrations at Waterman Farm for samples that were leached with DI Water as part of the sequential extraction process...... 77

Figure 13 Box plot showing U concentrations at Waterman Farm for samples that were leached with 0.1 M sodium bicarbonate solution as part of the sequential extraction process...... 78

Figure 14 Box plot showing U concentrations at Waterman Farm for samples that were leached with 10% HCl as part of the sequential extraction process...... 79 xiii

Figure 15 U concentrations from non-sequential HCl extractions for “Wooster” soils. .. 80

Figure 16 U concentrations from DI Water extractions for “Wooster” soils ...... 81

Figure 17 U concentrations from 0.1M sodium bicarbonate extractions for “Wooster” soils ...... 82

Figure 18 U concentrations from the sequential HCl extractions for “Wooster” soils ..... 83

Figure 19 U concentrations at topsoil depth (0-5cm) throughout the contiguous US (Smith, et al., 2014)...... 84

Figure 20 Location of sample sites used in this investigation over their respective counties in Ohio...... 85

Figure 21 Locations of sampled sites over the glaciated and unglaciated portions of Ohio ...... 86

Figure 22 Locations of sampled sites over Ohio’s bedrock geology...... 87

Figure 23 Locations of sampled sites over Ohio’s general land use/cover...... 88

Figure 24 Locations of sampled sites over Ohio’s geological units ...... 89

Figure 25 Map depicting the location and concentration range of samples taken from the NURE dataset...... 90

Figure 26 “Heat map” showing the ranges of U concentrations throughout the extent of the NURE dataset using Nearest Neighbor interpolation...... 91

Figure 27 Map showing the amount of NURE data that is contained within both the glaciated and unglaciated areas of Ohio...... 92

Figure 28 Map showing the NURE dataset over Ohio’s bedrock geology...... 93

Figure 29 Map showing the NURE dataset over Ohio’s general land use/cover...... 94

Figure 30 Map showing the NURE dataset over Ohio’s geologic units...... 95

Figure 31 Map showing the extent of the USGS topsoil dataset, including the ranges in U concentrations, over Ohio...... 96

xiv

Figure 32 “Heat map” of USGS topsoil U concentrations over Ohio using Nearest Neighbor interpolation...... 97

Figure 33 Map showing USGS topsoil U concentrations over Ohio’s glaciated and unglaciated areas...... 98

Figure 34 Map showing USGS topsoil U concentrations over Ohio’s bedrock geology . 99

Figure 35 Map showing USGS topsoil U concentrations over Ohio’s general land use/cover...... 100

Figure 36 Map showing USGS topsoil U concentrations over Ohio’s geologic units. .. 101

Figure 37 Monazite from sample EB-311 (East Badger Site) ...... 102

Figure 38 Zircon from sample WE-202 (Western Site)...... 103

Figure 39 Cerianite identified through spot chemical analysis from sample WE-202 (Western Site)...... 104

Figure 40 Zircon from sample NW-303 (Northwest Site)...... 105

Figure 41 Zircon from sample FA 1-2 0-5cm (Fallow plot at Waterman Farm)...... 106

Figure 42 Monazite from sample FERT 5-5 0-5cm (Fertilized plot at Waterman Farm) ...... 107

Figure 43 Monazite from sample FA 1-2 0-5cm (Fallow plot at Waterman Farm)...... 108

Figure 44 Trident Internal Standard Kit “T-Piece”...... 109

Figure 45. Graph comparing LOI and dichromate organic carbon values...... 110

xv

Chapter 1. Introduction

Since the 1950’s, the use of P-rich fertilizer derived from the mining of phosphate rock has been a major contribution to increases in crop yields (MacDonald, et al., 2011; Van

Kauwenbergh, 2010). Uranium-series nuclides (uranium (U), thorium (Th), protactinium

(Pa), radium (Ra), radon (Rn), etc.) are known to be present in association with phosphate rock deposits of marine origin (Burnett & Kim, 1985). The wide-scale use of fertilizers has created an increasing concern in their potential to act as non-point source pollutants in soils and waters, particularly those near to or acting as runoff sites from agricultural operations. Runoff from fertilized landscapes has often resulted in the eutrophication of natural waters, where nitrogen (N) and phosphorus (P) delivery into waterways accelerates normal plant growth and promotes excess and toxic algal blooms, which enhance decomposition, resulting in oxygen depletion in natural waters (Kling, et al.,

2014; Michalak, et al., 2013; Otero, et al., 2005; Pelley, 2016). The impact of nitrogen and phosphorus additions to landscapes have long been a focus of agricultural research

(Goyette, et al., 2018; Kling, et al., 2014; Liesch, et al., 2015; Otero, et al., 2005;

Spalding & Sackett, 1972; Suh & Yee, 2011), but little information exists on the addition, accumulation and mobility other constituents besides N and P compounds, such as potentially toxic metals. Like N and P, annual additions of fertilizer may also lead to the long-term accumulation of U in soils and waters (Goyette, et al., 2018).

1

In nature, uranium is present in low concentrations in rocks, soils, and water. On average, uranium occurs in the continental crust at 2.7 μg/g (Rudnick & Gao, 2003), while typical soil background levels range from 0.3 – 11.9 μg/g (Alloway, 1995; UNSCEAR, 1993).

Uranium is present as a trace constituent in apatites (a group of phosphate minerals) in amounts that range from 30 to 300 μg/g and can be enriched to as much as 500 μg/g in phosphorites that have been reworked in a marine environment. When present as U (IV), uranium can replace the calcium in an apatite’s structure, or can be absorbed as uranyl onto apatite crystal structures and is readily removed (and oxidized to U (VI)) from an apatite by weathering and oxidation (Altschuler, 1973; Cathcart, 1978).

A large variety of phosphate-rich rock deposits from around the world are used for the manufacturing of P-based fertilizers. In 2010, global phosphate rock reserves were estimated to be 65,613 million metric tons (mmt), which were predicted to contain 5.7 mmt of recoverable U (Van Kauwenbergh, 2010). During phosphate fertilizer production, apatite is dissolved by sulfuric acid (H2SO4) during the preparation of superphosphate, while uranium remains as uranyl sulfate [(UO2)SO4] and uranous sulphate [U(SO4)2], which are both soluble in water (Rothbaum, et al., 1979). The estimated U content of these reserves from the world’s major producers of phosphoric acid (an intermediary in

P-fertilizer production) in 2010 was 27g U/t for China, 97 g U/t for Morocco, and 99 g

U/t for the U.S (Ulrich, et a., 2014). The United States, Morocco, and China produced 25,

12, and 7% of the world’s phosphoric acid in 2010, respectively (Jasinski, 2013; Ulrich, et al., 2014). The U concentrations of fertilizers used in different parts of the world are shown in tables D-40 & 41.

2

1a. U in Soils; General Considerations

Soil pH, organic matter, and the presence of Fe(III) oxides and hydroxides can all play a major role in U mobility in soils. Uranium is most common in the environment in the U

(IV) and U (VI) oxidation states. These different oxidation states can affect the rate of migration through soils and other geologic media. In aqueous solution, uranium exists as

2+ the uranyl (UO2 ) ion. U becomes highly mobile in soils and groundwater when uranyl

2- 4- forms stable complexes with carbonate and bicarbonate [UO2(CO3)2 or UO2(CO3)3 ] that are poorly retained by soils (Zhou & Gu, 2005). The physical and chemical conditions found in most soils are alkaline pH and high reducing potential, allowing U to be more easily dissolved and hence transported to surface and groundwaters (Liesch, et al., 2015; Spalding & Sackett, 1972).

Cation Exchange

In soils, cation and anion exchange processes occur between ions in solution and ions in the boundary layer between the solution and charged surfaces present in a soil (Sparks,

2003b). Ion exchange is usually a rapid, diffusion-controlled, reversible, and stoichiometric process, with most cases having some selectivity of one ion over another by the surface where exchange is occurring. The stoichiometry in ion exchange processes requires ions that leave the colloidal surface in a soil to be replaced by an equivalent amount of ions (in terms of charge) in order to maintain neutrality (Sparks, 2003b). The rate at which ion exchange occurs in soils is also dependent on factors such as soil acidity

(soil pH), the variety and quantity of the soil’s organic (humus, organic matter) and inorganic (clays) surfaces and the charge and radius of the ion in question (Sparks, 3

2003b). Cation substitution can occur on surfaces such as organic matter, amorphous minerals, and within the crystal lattice of soil clay minerals like smectites, which consist of octahedrally arranged atoms between two sets of tetrahedrally arranged silicon atoms.

Cation substitution can also occur in a clay’s (such as smectite) crystal lattice within the octahedral (Mg2+ for Al3+) or tetrahedral (Al3+ for Si4+) sheets, which can result in a fixed negative charge on the clay’s surface, which can dictate the clay’s cation exchange capacity (CEC) (McKinley, et al., 1995). 1:1 clay minerals such as kaolinite have rapid rates of ion exchange, due to only external exchange sites being present, while 2:1 clay minerals, such as vermiculites and micas, have both internal and external exchange sites, resulting in slower kinetics (Sparks, 2003b). In the case of U, uranyl is sorbed onto the negative surfaces present in clay minerals, sesquioxides (ex. Fe2O3) and organic compounds (Vandenhove, et al., 2007).

Soil pH

Soil pH is a critical geochemical parameter that can have a significant impact on the leaching, precipitation, and mobilization of U, due to being able to control its chemical forms, which may behave differently in terms of mobilization (Zewainy, 2009). U can mobilize in both acidic and alkaline soils, making it amphoteric (Grassi, et al., 2005). U speciation is dominated by uranyl in acidic environments (Zewainy, 2009). In soil environments with lower pH values, U(VI) preferentially sorbs onto soil particles, and decreases in the amount of sorption to the soil particles as pH becomes more alkaline, which can be observed in a large majority of cultivated soils (aerated environments that usually have a pH range of 4-9) (McKinley, et al., 1995; Giblin, et al., 1981; Morrison, et 4 al., 1995). Environments with a pH greater than 4 allow cationic uranyl hydroxide and uranyl carbonate complexes to form, with the latter complexes being anionic at pH values above 9 (Duff & Amrhein, 1996). A soil’s CEC is directly related to pH due to the development of a greater negative charge on organic matter and clay minerals due to the deprotonation of functional groups present on these surfaces as pH increases (Sparks,

2003b).

Organic Matter

The role of organic matter in soil chemistry is significant and complex, due to its variability and interconnection with iron oxides and clay minerals in soils, as well as the chemical reactions it undergoes with organic compounds and metals present in soils

(Sparks, 2003a). U is able to sorb (any removal of a compound from solution to a solid phase), chelate (binding of metal ions to larger organic molecules), and complex

(combination of atoms, ions, or molecules to create a larger ion or molecule) with soil organic matter (Zewainy, 2009). Uranyl has been observed to combine with organic ligands in soil such as humus, acetate, oxalate, and citrate (Ganesh, et al., 1997; Huang, et al., 1998). A strong linear relationship between U(VI) and total organic carbon was observed by Zhou and Gu (2005), attributing the complexation of U(VI) with soil organic matter to the amount of U(VI) extracted by hydroxide (NaOH). Large fractions of U(VI) were determined to be retained in humic substances through complexation reactions in batch experiments (Crançon & van der Lee, 2003).

5

Iron Oxides

Fe oxides play an important role in the immobilization of uranyl in a soil’s subsurface through sorption, where U can co-precipitate with Fe oxides and become bound in the soil (Zewainy, 2009). Under neutral to slightly basic conditions, Fe-oxide minerals such as goethite (FeOOH), ferrihydrite, and hematite (Fe2O3) can strongly sorb hexavalent U

(Bargar, et al., 1999; Kohler, et al., 1992; Moyes, et al., 2000).

1b. Previous Work

The impact of uranium accumulation in soils through long-term phosphorus fertilizer application has been studied in the past, but only in a limited number of locations. A 2008 study in Japan involving over 20 years of fertilizer application on surficial agricultural soils concluded that most of the U derived from phosphate fertilizer application was incorporated into 2 pools: soil organic matter (SOM) or poorly crystalline Fe/Al minerals on the surface soils. The latter pool can experience alternating changes in redox conditions, which introduces the possibility of U associated with organic matter being redistributed to other fractions of the soil (Yamaguchi, et al., 2009).

A study comparing historical (archived, sampled 36-43 years previously) and contemporary (1992) New Zealand soils used for pastoral farming found that nearly all of the phosphate fertilizer-applied uranium remained in the soils and had not leached to groundwater or had been taken up by plants. However, the total U concentrations in the fertilizers were low compared to the concentrations found in the fertilizers associated with the historical soils. This result is possibly due to present day fertilizer manufacturing

6 processes using phosphate rock from sources containing lower U concentrations, making the current rate of U accumulation slower when compared to historical data. The study concluded that all or nearly all of the U applied in phosphate fertilizer appears to remain in the soil and did not solubilize into groundwater or get taken up by plants (Taylor,

2007).

A study of fertilized and unfertilized pastures, as well as native grasslands in central

Florida found that U has a strong affinity for soil organic matter in pasture areas (34% weight total U) at 3-6 cm of profile, and 15% weight total U at 9-12 cm is fertilizer- derived) and 5 – 42% of the total U in these soils is loosely bound and readily exchangeable. The study also measured low concentrations of U and dissolved phosphate in sampled runoff waters during rainfall events (<0.1 μg/L U and <1.0 mg/L phosphate), indicating the tendency of both to stay fixed to the soils. However, their presence in runoff waters suggest that they are slowly released over time, several years after initial fertilization (Zielinski, et al., 2006).

A 2008 study in Lower Saxony, Germany, investigated the mobility of uranium in sandy soils (sandy gleyic podzols) in both natural forests and farmland, and determined U concentrations in groundwater at both sites. U concentrations were found to be higher

(0.96 μg/g) in farmland soils at 20 cm depth and decreased at depths to 50 cm (0.58

μg/g). Forest soil samples at the same depths contained lower U concentrations (0.53 μ g/g and 0.35 μg/g, respectively) compared to farmland soils. U presence in forest soils was believed to be a result of wind-blown dispersal of fertilizers or soil particles containing U. Groundwater samples of a maximum depth of 7 meters had higher 7 concentrations of U at the farmland sites (16.4 nmol/L U) than in forest sites (<0.025-2.1 nmol/L U), and suggest that U concentrations generally decrease with depth. They explained this by stating that the U is being retained on the negatively charged surfaces present in soils (clay minerals, humic substances, etc.) (Huhle, et al., 2008).

1c. Potential Human Impacts

Toxicological and epidemiological studies on uranium exposure through drinking water collected from private wells or cisterns in areas that receive runoff near uranium processing sites, and areas with naturally elevated uranium levels have shown that chronic ingestion of uranium can induce nephrotoxic effects (affecting kidney function)

(Coyte, et al., 2018; Kurttio, et al., 2002; Pinney, et al., 2003; Zamora, et al., 1998). The

United States Environmental Protection Agency (USEPA) and the World Health

Organization (WHO) have set a provisional guideline of 126 nmol/L for uranium in drinking water (EPA, 2019; WHO, 2012), which suggests that uranium additions to natural waters through human activities need to be characterized.

A previous investigation of surface water bodies draining agricultural areas in Ohio found dissolved U concentrations ranging from 1.2 – 16.2 nmol/L, with the lowest concentrations being observed in Lake Erie. These authors, based on comparisons to concentrations of U found in global surface waters suggested that waters draining fertilized agricultural lands contain higher than global background level dissolved U concentrations (Lyons, et al., 2020).

8

1d. Rationale for Research

This investigation serves to answer the following questions:

1. How do U concentrations differ between fertilized and non-fertilized agricultural

soils?

2. Do the properties (organic matter content, pH, etc.) of Ohio agricultural soils

influence the quality and lability of U in soils?

3. What agricultural management practices introduce the least amount of uranium to

soils (i.e. manure, cover crop, fertilizer application rate, etc.)?

This work seeks to assess the uranium concentrations of select Ohio agricultural soils and compare them to U concentrations in the same type of soils without fertilizer additions.

This will be done to evaluate the role of synthetic phosphorus fertilizers as a potential source of uranium to soils, as well as extrapolate these results and apply them towards other agricultural areas. This work also seeks to compile historical geochemical, geographical, and spatial data in order to characterize uranium concentrations on a scale broader than just small managed plots where the data from this investigation is derived.

The analysis of Ohio agricultural soils will assist in determining if fertilizers act as a source of uranium to the aquatic environment. This work is one of the first studies of its kind that directly assesses the impact of chronic fertilizer use on agricultural soils in the

Midwestern United States.

9

1e. Hypothesis

This study hypothesized that:

• Agricultural soils that have been treated with P-rich fertilizer will have higher

solubilizable U concentrations than control site soils that have not been treated

with fertilizer.

• U concentrations will decrease with soil depth in the analyzed agricultural soils.

• Soils treated with higher rates of fertilizer addition will have higher solubilizable

U.

10

Chapter 2. Study Areas

2a. Waterman Farm

The Waterman Agricultural and Natural Resources Laboratory (commonly referred to as

Waterman Farm), located in Columbus, Ohio, is a 106 hectare (261 acre) Agricultural

Best Management Practices and Education site that is managed by the College of Food,

Agricultural, and Environmental Sciences (CFAES) and has been part of The Ohio State

University since its inclusion in 1923. Waterman Farm serves as a working model for students, owners, and operators around Ohio for managing farms for improved water quality and ecosystem services (Ward, et al., 2019). The two areas sampled from

Waterman Farm in this investigation included plots under different management practices, including the use of cover crops (plots 7 and 13) and manure addition (plots 9 and 11) (Figure 1. – “Morgan’s Experiment”), as well as the application (plots 3, 13, and

17), and lack of application (plots 4, 11, and 15) of fertilizer (244 kg/ha N, with 184 kg as

Urea, and 60 kg as NPK) (Figure 2. – “East Straw Mulch Experiment”).

The “East Straw Mulch Experiment” plot at Waterman Farm is part of a long-term wheat residue management experiment established in 1996, having been used to assess the impacts of surface mulching on soil organic carbon sequestration, greenhouse gas emissions, soil structural properties, and nutrient dynamics (Blanco-Canqui & Lal,

2007b; Jacinthe, et al., 2002; Saroa & Lal, 2003, 2004). The soil is a Crosby silt loam with a 1% slope in elevation and is poorly drained and developed from a loamy glacial till. Additionally, no crops are grown on the soil, and no cultural operations are performed on the field’s plots (Blanco-Canqui & Lal, 2007a). 11

The “Morgan’s Experiment” field was established in 1997, independently from the other plots at Waterman Farm, with the dominant soil series at this site being a Crosby silt loam (fine, mixed, mesic Aeric Ochraqualf, same as the East Straw Mulch Experiment plot). The four treatments at this site involve the use of compost (mixture of hardwood mulch, straw, and horse manure; applied every year), manure application (cow manure; applied every year), permanent cover crops (no amendments, grasses including 50% perennial rye, 30% annual rye, 10% red fescue, and 10% blue grass), and fallow (no amendments added to the soil, weeds). Compost was manually applied to the site’s plots at a rate of 44 Mg/ha in early April, while cow manure was applied in December at 29

Mg/ha using a manure spreader. No chemical fertilizers (N, P, K) were applied to the amended plots. Urea-N fertilizer was broadcasted manually (spread by hand) in the fallow and cover crop plots at a rate of 148 kg N/ha at the time of corn emergence in the amended plots (compost and manure plots at the same site) (Shrestha, et al., 2013).

2b. “Wooster” Sites

The Soil Fertility Group, headed by Dr. Steve Culman at the Ohio State University

Agricultural Technical Institute (ATI) in Wooster, Ohio, provided samples from 3 different plots that are part of a long-term soil fertilization experiment. This experiment involves the evaluation of three different P and K fertilizer application rates: no fertilizer

(0x), the estimated nutrient removal rate (1x), and twice the estimated nutrient removal rate (2x). In the context of this thesis, the 36 provided samples involved only the use of P fertilizer application rates (where 0x, 1x, and 2x are shown as P = 0, 100, and 300, respectively on corn (Figures 3-5) (Fulford & Culman, 2018).

12

Nutrient removal rates were determined by multiplying the 2005 Ohio state-wide average corn and soybean yield (9.1 Mg/ha and 2.7 Mg/ha, respectively) by their estimated nutrient removal rates (kg of nutrient/Mg of grain) for P2O5 and K2O (60.1 kg/ha P2O5 and 43.7 kg/ha K2O for corn; 35.9 kg/ha P2O5 and 62.9 kg/ha for soybean). The purpose of calculating the nutrient removal rate is to keep soil test P and K values within a maintenance range, where fertilizer application would replace the nutrients lost each year through crop removal (Fulford & Culman, 2018; Vitosh, et al., 1995).

Three field trials were established in 2006: one at the Western Agricultural Research

Station (Figure 4) (39°51’39” N, 83°40’45” W) in Clark County on a Kokomo silt loam

(fine, mixed, superactive, mesic Typic Argiaquoll) (Soil Survey Staff, et al., 2016), one at the East Badger Farm at the Ohio Agricultural Research and Development Center

(OARDC) (Figure 3) in Wayne County (40°46’43” N, 81°50’22” W), on a Canfield silt loam (fine-loamy, mixed, active, mesic Aquic Fragiudalf) (Soil Survey Staff, et al., 2016) and one at the Northwest Agricultural Research Station (Figure 5) in Wood County

(41°12’46” N, 83°45’50” W) on a Hoytville clay loam (fine, illitc, mesic Mollic

Epiaqualf) (Fulford & Culman, 2018; Soil Survey Staff, et al., 2016).

2c. Coshocton

Samples provided from the Fortner et al. 2012 study were taken from the former U.S.

Department of Agriculture North Appalachian Experimental Watershed (NAEW) area, near Coshocton, Ohio (40°22’ N, 81°41’ W). The Coshocton region was not glaciated during the Last Glacial Maximum [~26.5 – 19 ka (Clark, et al., 2009)], and it consists of shallow and well-drained residual soils with silt loam surfaces that are derived from 13

Pennsylvanian shales and interbedded sandstones (Kelley, et al., 1975). Soils from the

NAEW are mainly composed of aluminosilicates and contain less than 5% carbonates

(Eckstein, et al., 2007) (Fortner, et al., 2012) (Owens, et al., 2008). Analysis of the bulk mineralogy and clay fraction of the unsaturated zone in the NAEW showed that quartz

(70-90%), muscovite (5-20% as illite), and kaolinite (5-15%) are the principal minerals in both the siltstone and shale rocks present in the area, and the clay minerals in the clay fraction of the rocks include illite (30-60%), kaolinite (10-30%), and vermiculite (<10%)

(Eckstein, et al., 2007). A description of the experimental plots and their usage over time can be found in Owens et al. (2008) and Diaz, et al. (in review). In this investigation, three samples, each pertaining to a different management practice from the Coshocton site were used: “no till, no manure”, “pasture”, and “till, no manure”.

14

Chapter 3. Methods

3a. Cleaning Plasticware

Prior to use in the laboratory, 60 mL high-density polyethylene resin (HDPE) Nalgene bottles and 50 mL polyethylene Falcon tubes were rinsed three times with deionized (DI) water and left to dry for 24 hours under a laminar flow hood.

3b. Sampling and Processing

Samples from Waterman Farm were collected using a metal soil auger at depths of 5 cm and 40-50 cm. Each soil sample was removed from the auger using a metal spoon and placed in quart (0.94 L)-sized clean polyethylene bags and were closed with a slider seal or rubber bands. A total of 68 samples were collected, with a total of 24 samples collected from the manure (cow manure, 29 Mg/ha) and cover crop (no amendments, grasses) plots at the “Morgan’s Experiment” field (Figure 1) (6 samples at each plot for 4 plots) and the remaining 44 samples are from fallow (no fertilizer applied) and fertilized

(244 kg/ha N [184 kg as Urea, 60 kg as NPK]) plots at the “East Straw Mulch

Experiment” site (Figure 2) (3 plots used for each management practice, 2 of the 3 plots have 6 samples each, and the third plot of each management practice [plots 15 and 17 on

Figure 2.] has 10 samples).

The samples provided by Dr. Culman were collected from the Western Agricultural

Research Station, East Badger Farm, and the Northwest Agricultural Research Station, at the surface 20 cm of their respective plots during the fall following crop harvest, prior to broadcast (scattering of seeds, by hand or mechanically) and chisel tillage incorporation

15 of P and K fertilizer. Samples were taken in seven to 10 2.5 cm diameter soil cores in between planted rows, and were then composited, air-dried, and sieved (<2mm) before being placed into Manila envelopes for transport.

Soil samples from Coshocton, OH were taken from pits dug either by hand or with a backhoe and were sampled in profile. The three samples used in this investigation were sampled at a depth of 0-2.54 cm. Each sample was weighed into 20-30-gram aliquots on polyethylene weigh boats and dried in an oven at 25 - 28°C for 24-48 hours.

Samples were weighed again after drying to account for water loss and finely ground using a ceramic mortar and pestle before being placed into new LDPE bags prior to analysis. 50 mL polypropylene Falcon tubes were rinsed twice with deionized (DI) water and dried under a laminar flow hood for 24-48 hours before being used for extractions.

Approximately 5 grams of soil sample were then weighed and placed into clean Falcon tubes for extractions.

3c. Extractions

Forty-eight of the 68 soil samples from Waterman Farm, 27 of the 36 samples from

Wooster, and all 3 of the Coshocton samples were leached using 10% v/v trace metal grade HCl at a 7:1 ratio (35 mL of acid for 5 grams of soil) for 24 hours at room temperature (25°C). The remaining samples from all of the aforementioned sites were sequentially leached over a period of 3 days, first using DI Water (from a Millipore Milli-

Q Integral 10 Water Purification System), then a 0.1 M sodium bicarbonate (NaHCO3) solution (both at a 1:5 ratio, or 25 mL of solution per 5 grams of sample), and finally, a

16

10% v/v HCl solution (7:1 ratio). Upon extraction, all soil samples were filtered through

47 mm Whatman polycarbonate membrane filters. After filtration, all leachates were transferred to pre-cleaned 60 mL HDPE bottles. pH estimates were taken with pH strips throughout the sequential extractions, with the average pH values for DI water, bicarbonate, and 10% HCl leachates were 6, 8, and 1, respectively, after extraction. These techniques have been previously utilized for similar types of studies (Garrett, et al., 2009;

Zielinski, et al., 2006; Zielinski & Meier, 1988; Tessier, et al., 1979).

De-ionized (DI) water was used to measure the amount of water-soluble uranium in the soils (Garrett, et al., 2009) and is shown as:

2 + + UO2 + H2O ↔ UO2OH + H

(McKinley, et al., 1995)

17

The extraction of loosely bound and readily exchangeable U was performed through the use of 0.1M sodium bicarbonate (NaHCO3) solution (Zielinski, et al., 2006). As pH increases through the addition of bicarbonate solution, there is an increase in the ionization of acidic functional groups such as COOH and OH, providing negatively- charged sites for cation exchange (Zielinski & Meier, 1988), as well as the formation of very soluble complexes of U(VI) and carbonate, shown as:

2+ 2- 2- UO2 + 2CO3 ↔ UO2(CO3)2

or

2+ 3- 4- UO2 + 3CO3 ↔ UO2(CO3)3

(Duff, et al., 1998)

Acid-soluble U (and possibly U bound to organic matter under oxidizing conditions) was extracted from the soils in this investigation through leaching using a 10% HCl solution

(Tessier, et al., 1979). In principle, UO2 can react with HCl in two ways:

UO2 + 4HCl → UCl4 + 2H2O

or

2UO2 + 2HCl → 2UO2Cl + H2

(Volkovich, et al., 2007)

The effectiveness of each step in the sequential extraction process used in this investigation is determined through comparison with the concentration and quantity of U

18 extracted from the soil sample compared to the total U obtained from the sample through

XRF analysis.

3d. U-Analysis

2.8 mL of 10% v/v trace metal grade HCl was added to DI water and Bicarbonate leachates (each 25 mL) to make 10% acidified solutions of 27.8 mL. Samples were analyzed on a Perkin-Elmer SCIEX-ELAN 6000 quadrupole mass spectrometer. Each leachate was diluted using 2% HCl by a Trident Internal Standard Kit, or “T-piece”

(Figure 44), in which 900 μL of sample and 100 μL of 2% HCl are mixed in a polytetrafluoroethylene (PTFE) mixing chamber prior to introduction to the nebulizer at 1 mL/min. Element concentrations were taken as the mean of 10 replicates for each sample.

To account for drift during analysis, each leachate was spiked with a 10 μg/L (parts per billion) indium (In) solution. 10% hydrochloric acid leachates were spiked with 0.126 mL of indium (In) solution, and acidified DI water and Bicarbonate leachates were spiked with 0.1 mL of In solution.

Standards for ICP-MS (Inductively Coupled Plasma Mass Spectrometry) analysis of samples were made using a U standard solution. The U standard solution was a

PlasmaCAL 1000 μg/mL solution, in a 4% HNO3 matrix. An intermediate standard containing 1000 μg/L U was made to make the standards used for analysis on the ICPMS

(Table A-6). The analysis of the DI water extractions required making another intermediate standard (“Intermediate A”), containing 10 μg/L U) to make further diluted standards in order to detect lower element concentration (Table A-7).

19

3e. Precision and Accuracy

An internal standard (Environment Canada, TMDA-64.2, Certified Reference Water for

Trace Elements) was used to determine the accuracy of measurements performed through

ICP-MS. The standard has a certified U concentration of 142 μg/L (preserved in 0.2%

HNO3) and was measured to be 130.8 μg/L on the instrument, making measurements within 8% of their true value.

To determine the precision of analytical measurements, the relative standard deviation

(RSD) of the U concentrations for the check standards for the HCl/Sodium Bicarbonate and DI Water extraction runs were calculated (Tables A-8 & A-9). For the bicarbonate and HCl extractions, the check standard for U had an RSD of ~ 6.1 %. For the DI water extractions, the check standard for U at 0.1 μg/L had an RSD of ~ 2.4%

3f. pH Analysis

The pH of each soil sample was determined using the procedure of Kalra et al. (1995).

Two-gram aliquots of each sample were dried in an oven at 25°C for 48 hours and were subsequently ground. Ten mL of DI water was then added to each sample (for a 1:5 ratio of soil to water) and mixed and left to stand for 30 minutes and mixed again immediately before analysis using an Orion Model 520A pH electrode. The pH measurements were calibrated using a 3-point calibration using certified pH standards of 4, 7, and 10. pH readings were taken at one minute (values remained steady after approximately one minute) or when the electrode would return a final pH value (Kalra, et al., 1995). The DI water used in these tests had a pH of approximately 8.1.

20

3g. Loss on Ignition

Organic matter was analyzed as loss on ignition (LOI). 2-gram aliquots of each sample weighed in mullite crucibles and dried for an hour at 100°C to remove water from the sample and weighed again before being placed in a Fisher Scientific Isotemp muffle furnace and heated at 500 – 550°C for 4 hours, and then weighed again after cooling. The

LOI is calculated using the following equation from Heiri, et al.:

LOI550 = ((DW25 – DW550) / DW25) * 100

where LOI550 represents the percentage of LOI at 550°C, DW25 represents the dry weight of the sample prior to combustion, and DW550 represents the dry weight (grams) of the sample after it was heated to 550°C (grams). The weight loss in the sample after combustion has been interpreted as the amount of organic carbon that is present in the soil (Heiri, et al., 2001). The heating of the samples to 550°C likely results in the removal of structural water from the soils but does not lead to loss of calcium carbonate. LOI was conducted both in-house at The Ohio State University and by SES Canada for the samples sent for XRF analysis.

To confirm the validity of the results from both the in-house LOI and the LOI results from samples sent to SES Canada for XRF analysis, the organic carbon content of different aliquots of the XRF-analyzed samples (Table A-10) was determined through dichromate reduction by the Soil Chemistry lab group at The Ohio State University

School of Environment and Natural Resources. The method is a microwave-assisted digestion, involving the use of potassium dichromate (K2Cr2O7) to determine a soil

21 sample’s organic carbon content. The digestion procedure causes the soil’s organic carbon to become oxidized, and Cr(VI) is reduced to Cr(III), which is quantified through a spectrophotometer at 600nm (Soil Environmental Chemistry Program, The Ohio State

University). The comparison of these LOI data to those analyzed via the dichromate method for organic carbon is shown in Figure 45. Clearly, in these samples, the LOI overestimates the amount of organic carbon present in these soils. Because of this, this work will refer to these measurements throughout the text as “LOI” rather than organic carbon. However, because a portion of the LOI is organic matter, it will be considered as a surrogate, but not an absolute amount of organic carbon.

3h. XRF Analysis

Thirteen samples were selected (one topsoil sample from each management practice at

Waterman Farm [4 total], and one for each fertilizer application rate at each location from the provided “Wooster” samples [9 total]) and sent to SGS Canada in Lakefield, Ontario, for X-ray Fluorescence analysis of major oxides (Tables 2-4), total U (Table D-1), and loss on ignition (LOI) measurements (Table A-1). Samples were pulverized after weighing. Total U was measured by ICP-MS after multi-acid digestion (HNO3, HF,

HClO4, and HCl). Whole rock analysis (determination of major element oxides in a sample) was used to measure the major oxides (Al2O3, CaO, Cr2O3, K2O, MgO, MnO,

Na2O, P2O5, Fe2O3, SiO2, and TiO2), as well as LOI in the soil samples through Fusion-

XRF (borate fused disc). One sample (WE-202) was used as a replicate for the major oxides and for LOI, which resulted in a 2.8% overall error for the major oxides and a

22

0.47% error in LOI measurement. Another sample (NW-304) was used as a replicate for total U and resulted in a 1.7% error.

3i. GIS Maps

ArcGIS was used in conjunction with data from the 1973 National Uranium Resource

Evaluation (NURE) (Bolivar, 1980) and from a 2014 study conducted by the U.S.

Geological Survey (USGS) that mapped the geochemistry and mineralogy of soils in the conterminous United States, including Ohio (Smith, et al., 2013).

The NURE program, initiated in 1973 by the Atomic Energy Commission (now the

Department of Energy; DOE) to identify potential U resources in the US. One of the program’s components was the Hydrogeochemical and Stream Sediment Reconnaissance

(HSSR) program, involving the sampling of stream sediments, groundwater, and surface water. The data used in this evaluation were only the sediment data. The HSSR program ended in 1980, and the NURE program lost Congressional funding by 1984, having only

307 of the 625 planned quadrangles sampled and partially sampling another 86 (U.S.

Geological Survey, 2004).

In 2007, the USGS began a low-density geochemical and mineralogical survey (1 site per

1600 km2, with a total of 4857 sites) of soils in the conterminous US as part of the North

American Soil Geochemical Landscapes Project. Soils were sampled at 0 to 5 centimeters, and composites were made at their A and C horizons (in this investigation, only data from the top 5 cm of the soil profile were considered). The resulting data of this survey provides estimates for the abundance and spatial distribution of chemical elements

23 and minerals present in US soils and represents a baseline of geochemical and mineralogical data against which future changes can be recognized and quantified.

These two data sets were overlain with GIS maps showing Ohio land use/cover (USGS

Gap Analysis Program, 2016), glaciation, and geology (Nicholson, et al., 2005) in order to draw correlations between the spatial and geochemical data and to provide a broader view of U contamination in Ohio agricultural soils.

3j. Scanning Electron Microscopy

In order to identify possible U-bearing minerals from the soils used in this investigation,

6 samples were analyzed using a FEI Quanta FEG Field Emission Scanning Electron

Microscope (SEM). Spot chemical analyses on samples were conducted using a Bruker

Electron Dispersion Spectroscopy (EDS) detector.

24

Chapter 4. Results

4a. USGS and NURE Ohio GIS Map Observations

Maps showing U concentration data for both the NURE and USGS datasets, as well as the locations of the sites used in this investigation were made using ArcMap. These data were overlain with maps of counties, the glaciated/non-glaciated portions, bedrock geology, and land use/cover of Ohio.

Sample site locations from this investigation are shown in Figure 20. Figure 21 shows the

Northwest, East Badger, Western, and Waterman Farm locations in the glaciated portion of Ohio, while the Coshocton site is shown in the unglaciated area. When placed over

Ohio’s bedrock geology, the Northwest and Western sites are located over dolostone

(Salina Group; Clinton and Cataract Groups, respectively), while the East Badger

() and Coshocton (Allegheny and ) sites are located over shale, and Waterman Farm is located over limestone ()

(Figure 24) (Table A-16). Placing the sample site locations over the land use/cover map shows Waterman Farm over developed land, while the Northwest, East Badger, and

Western sites are over agricultural land, and the Coshocton site is over undeveloped land.

The NURE data set has 1000 total samples for U concentrations and the mean is 4.0 μg/g, with a standard deviation of 1.0, and a coefficient of variation of 25% (Table D-35). The

USGS data set for the top 5cm of soils in Ohio has a total of 69 samples and has an average U concentration of 4.1 μg/g, with a standard deviation of 1.2 and a CV of ± 28%

(Table D-35).

25

The extent of the NURE dataset is limited to Eastern/East Central Ohio (Figure 25), with the highest concentrations occurring in the southern portions of Eastern Ohio (Figure 26).

Placing the NURE data in the context of the previous Pleistocene glaciation events in

Ohio (Figure 27), the majority (78.2%) of the U data was collected in the unglaciated part of Ohio, averaging at 4.1 μg/g and ranging from 1.9 to 19 μg/g, with a standard deviation of 1.0 and a CV of ± 24%, (Table D-36). The remaining 21.8% of the NURE data in the northeastern glaciated part of Ohio has an average U concentration of 3.3 μg/g, ranged from 1.4 to 6.7 μg/g, a standard deviation of 0.66, and a CV of ± 20%, (Table D-36).

Placing the NURE data in the context of Ohio bedrock geology, U concentrations were averaged at 2.40, 4.54, 2.93, 3.93, and 4.35 μg/g for soils over black shale (), mudstone (Waynesville and Arnheim Formations), sandstone ( and

Bedford Shale), shale (Monongahela Group, Allegheny and Pottsville Groups, Maxville

Limestone, Rushville, Logan, and Cuyahoga Formations), and siltstone (Conemaugh

Group), respectively (no data points exist over dolostone or limestone units) (Tables A-

16 & D-38) (Figure 28) (Figure 30). Due to the high density and amount of points, as well as the spread of the data being located in only one portion of Ohio, concentrations from the NURE dataset were not compared with the land use/cover data.

The USGS dataset, while smaller in terms of sample size, is geographically spread over

Ohio (Figure 31), with the highest concentrations occurring in northwest and central Ohio

(Figure 32). Placing the USGS data in the context of past glaciations (Figure 33), 32% of sampled points are in the unglaciated region of Ohio, with an average U concentration of

3.6 μg/g, ranging from 2.7 – 5.2 μg/g, a standard deviation of 0.61, and a CV of ± 17%.

26

Sampled points from the glaciated region (68% of the Ohio dataset) range from 1.7 – 9.0

μg/g, with an average U concentration of 4.4 μg/g, a standard deviation of 1.29, and a CV of ± 29% (Table D-36). Comparing the USGS dataset to Ohio’s bedrock geology shows average U concentrations (calculated using the means of the data points over specific geological units from the same USGS data) of 4.4, 4.9, 4.1, 3.3, 3.5, 3.9, and 3.9 μg/g for the black shale (Ohio Shale), dolostone (Traverse Group, , Tymochtee and Greenfield Formations), limestone (Grant Lake and , Columbus

Limestone, ), mudstone (Waynesville and Arnheim Formations), sandstone (Berea Sandstone and ), shale (Allegheny and Pottsville Groups,

Maxville Limestone, Rushville, Logan, and Cuyahoga Formations, Monongahela Group,

Estill Shale, , Undifferentiated, Preacherville Member of the

Drakes Formation, , ), and siltstone (Conemaugh Group) bedrock units, respectively (Tables A-16 & D-38) (Figures 34 & 36). Overlaying the

USGS dataset over a simplified version (several categories were merged into 3 main categories: Agricultural Land, Undeveloped Land, Developed Land, and are shown in

Tables A-17 & 18) of the acquired land use/cover map and visually identifying each point

(USGS Gap Analysis Program, 2016) (Table D-37) shows the average soil U concentration in agricultural areas (54% of sampled locations) at 4.5 μg/g, while ranging from 1.7 to 9.0 μg/g with a CV of ± 29%. Samples taken from developed areas (10% of sampled locations) had an average U concentration of 4.0 μg/g, while ranging from 2.5 to

5.7 μg/g, with a CV of ± 25%. Samples taken from undeveloped areas in Ohio (36% of

27 sampled locations) had an average U concentration of 3.6 μg/g, while ranging from 2.5 to

4.9 μg/g, with a CV of ± 17%.

4b. Loss on Ignition for Samples

The LOI of the soils used in this investigation ranged from 2.78 – 11.0% (see: Appendix

C). Cover crop soils at Waterman Farm had an average LOI of 6.89 and 6.82 % for the topsoil (0-5cm) and bottom 40-50cm for their plots, respectively (Table C-1). Soils treated by manure application at Waterman Farm averaged at 8.17 and 6.87 % LOI at the topsoil and bottom 40-50cm depths, respectively. The notable difference between LOI averages for the manure-treated soil samples is due to sample M 1-3 having an LOI of almost 11 %, likely due to significant amounts of manure still being present in the sample, which was confirmed after processing four different aliquots of the sample

(Table C-2). Fallow soils at Waterman Farm had an average LOI of 3.69 and 3.66 % for the topsoil and bottom 40-50cm depths, respectively (Table C-3). Soils that were fertilized at Waterman Farm had an average LOI of 3.63 and 3.79 % for the topsoil and bottom 40-50cm depths, respectively (Table C-4).

Soils at the Northwest site, the soils had an average LOI of 5.71, 5.87, and 5.95 % at plots treated with 0x, 1x, and 2x the fertilizer application rate, respectively (Table C-5). At the

East Badger site had an average LOI of 4.06, 4.00, and 3.99 % for plots treated with 0x,

1x, and 2x the fertilizer application rate (60.1 kg/ha P2O5 and 43.7 kg/ha K2O for corn), respectively (Table C-6). At the Western site, soils had an average LOI of 3.80, 3.83, and

3.46 % for plots treated with 0x, 1, and 2x the fertilizer application rate, respectively

28

(Table C-7). Soils from Coshocton had an LOI of 7.07 (no till, no manure), 6.21

(pasture), and 3.97 (till, no manure) %, and averaged at 5.75 % (Table C-8).

4c. pH Values for Samples

The soils used in this investigation were found to be slightly to very acidic (pH ranging from 3.74 to 6.93) (see: Appendix B). A major cause of soil acidification in agricultural lands is the application of ammonium and urea-based fertilizers. Ammonium salts can be oxidized and produce acid through nitrification (Goulding, 2016).

+ - + NH4 + 2O2 = NO3 + 2H + H2O

The pH of soils at the “Morgan’s Experiment” field at Waterman Farm range from 4.21 –

6.25 and are averaged at 5.47, the pH of the topsoil (0-5cm) and at the bottom 40-50cm is averaged at 5.46 and 5.48, respectively (Table B-1). Soils with cover crops grown on them, also at the “Morgan’s Experiment” field, have a pH range of 4.49 – 6.77 and is averaged at 5.59. The pH of the topsoil and bottom 40-50cm have a pH of 5.37 and 5.81, respectively (Table B-2). Soils at the “East Straw Mulch Experiment” field at Waterman

Farm that were treated with fertilizer have a pH range from 4.49 – 6.56 and are averaged at 5.12. Topsoil and bottom 40-50cm pH values these soils were averaged to be 5.20 and

5.05, respectively (Table B-3). The “fallow” soils at the “East Straw Mulch Experiment” plot, in which no amendments were added, have a pH range from 3.74 – 5.51, and average at 4.62. Topsoil and bottom 40-50cm soil pH values for this management practice were averaged to 4.57 and 4.67, respectively (Table B-4). At the Coshocton site,

29 pH values ranged from 6.36 – 6.93, and were averaged at 6.54, making it the highest average pH of all the sites in this study (Table B-5).

The soils at the East Badger site had a pH range of 4.77 – 6.48 and were averaged at a pH of 5.76. Plots with no fertilizer application, the normal calculated fertilizer application rate (60.1 kg/ha P2O5 for corn plots) (Fulford & Culman, 2018), and twice the fertilizer application rate were averaged at a pH of 5.78, 5.63, and 5.84, respectively (Table B-6).

Soils at the Western site had a pH range of 5.25 – 6.18 and were averaged at a pH of

5.75. The average pH of plots at the Western site at 0x, 1x and 2x rates of fertilizer application were 5.68, 5.97, and 5.48, respectively (Table B-7). The pH of soils at the

Northwest site ranged from 4.63 – 5.41 and were averaged at a pH of 4.94. Northwest plots with 0x, 1x, and 2x fertilizer application rates had an average pH of 4.80, 4.94, and

5.04, respectively (Table B-8).

4d. XRF Data – Impact of Weathering on U Concentrations in Soils

SiO2 concentrations (ranging from 64.3 – 76.7%) are highest at the East Badger, Western, and “East Straw Mulch” locations, while the lowest concentrations are at the “Morgan’s

Experiment” and Northwest locations. Al2O3 concentrations (ranging from 8.6 – 12.9%) are highest at the Northwest location, while the lowest concentrations are at the East

Badger location. Fe2O3 concentrations, which range from 3.34 – 5.47%, are highest at the

Northwest location and lowest at the Western location. MgO concentrations, ranging from 0.54 – 1.31%, are highest in the Northwest location and lowest at the East Badger location (Table A-2).

30

CaO concentrations ranged from 0.36 – 1.09%, with the highest concentrations being found in the Northwest location. K2O concentrations range from 1.84 – 2.81%, are highest at the Northwest location, and lowest at the East Badger location. Na2O concentrations range from 0.75 – 1.05%, and are lowest at the East Badger location, and are highest at the Western location. TiO2 concentrations range from 0.7 – 0.83%, where they are highest at the East Badger location and lowest at the cover crop field at the

“Morgan’s Experiment” location (Table A-3).

MnO concentrations ranged from 0.05 to 0.19% and were highest at the cover crop and manure plots at the “Morgan’s Experiment” location and were lowest at the Western location. P2O5 concentrations ranged from 0.09 – 0.24% and were highest at the manure plot of the “Morgan’s Experiment” location and were lowest at the Western location.

Cr2O3 concentrations were relatively similar across all locations, with a range of <0.01

(below detection limit) to 0.02%. V2O5 also remained relatively similar, ranging from below detection limit (<0.01%) to 0.02% (Table A-4).

Total U concentrations in the soils appear to be consistent within their own plots/locations. Soils from the different management practices at Waterman Farm (fallow, fertilizer addition, cover crop, and manure application) range from 4.16 – 4.99 μg/g U, with the fertilizer-applied soil having the least amount of total U and the manure-applied soil having the most. The “Wooster” locations, with the exception of the Western location, show a slight increase in total U concentration when the fertilizer application rate is doubled from 60.1 kg/ha P2O5 and 43.7 kg/ha K2O to 120.2 kg/ha P2O5 and 87.4 kg/ha K2O for corn. All three “Wooster” sites are consistent in terms of total U 31 concentrations, with the East Badger, Northwest, and Western locations having ranges of

3.05 – 3.11, 5.17 – 5.42, and 2.7 – 2.87 μg/g, respectively (Table D-1). Total U concentrations averaged at 4.62, 3.07, 5.29, and 2.78 μg/g for the Waterman Farm, East

Badger, Northwest, and Western locations, respectively. The average total U could not be determined for the Coshocton site due to none of its samples being analyzed through

XRF.

The Chemical Index of Alteration (CIA) was calculated for each soil sample as a measure of chemical weathering of the sediments. CIA is a ratio that assumes that the dominant process during chemical weathering is the degradation of feldspars and the formation of clay minerals, and is calculated using the following equation:

CIA = [Al2O3/(Al2O3 + CaO + Na2O + K2O)] * 100

Calcium (Ca), sodium (Na) and potassium (K) are removed from feldspars by chemical weathering, making the proportion of alumina to alkalis increase as the degree of chemical weathering increases (Nesbitt & Young, 1982).

The range of calculated CIA ratios for the soils range from 70.7 – 74.9 (Table A-5), placing them in the same range of average chemical weathering seen in shales, indicating that the soils have a large proportion of clay minerals (Nesbitt & Young, 1982).

4e. Uranium Concentrations in Soils

A total of 78 samples were selected from the sites in this study to determine U concentrations through leaching by 10% HCl (non-sequential). The remaining 29 samples

32 from all sites were selected to determine U concentrations through a sequential leach process (DI water, 0.1 M sodium bicarbonate, and 10% HCl leaches) (See: Appendix D).

The Cover Crop plots at Waterman Farm had an average U concentration of 0.48 and

0.51 μg/g at 0-5 and 40-50cm depths, respectively. The U concentrations for Cover Crop soils at the 0-5cm depth have a standard deviation of 0.07 and a coefficient of variation

(CV) of ± 16%. The bottom 40-50cm of the cover crop soils have a standard deviation of

0.09 and a CV of ± 17% (Table D-2). A two-tailed t-test between U concentrations at both soil depths resulted in a P value of 0.44, indicating statistical difference between U concentrations at the cover crop plots based on soil depth (Table D-3) (Figure 11).

The Manure-treated plots at Waterman Farm had a mean U concentration of 0.5 and 0.46

μg/g at 0-5 and 40-50cm, respectively. The U concentrations for manure soils at 0-5cm had a standard deviation of 0.03 and a CV of ± 7%. The bottom 40-50cm of the manure soils had a standard deviation of 0.12 and a ± 27% CV. A two-tailed T test between the soil depths resulted in a P value of 0.89, indicating no statistical difference between U concentrations at the manure plots based on soil depth (Table D-4) (Figure 11).

The Fallow (no amendments) plots at Waterman Farm had average U concentrations of

0.54 and 0.55 μg/g at 0-5cm and 40-50cm, respectively. U concentrations at the topsoil

(0.5-cm) of the fallow plots had a standard deviation of 0.08 and a CV of ± 14%. The bottom 40-50cm of the fallow soils had a standard deviation of 0.04 and a CV of ± 7%. A two-tailed T test between U concentrations at both depths resulted in a P value of 0.89,

33 suggesting no statistical difference between U concentrations in the fallow soils based on their depth (Table D-5) (Figure 11).

Plots that were fertilized at Waterman Farm (244 kg/ha N, with 184 kg as Urea, and 60 kg as NPK) ha average U concentrations of 0.57 and 0.58 μg/g at 0-5 and 40-50cm, respectively. U concentrations for the fertilized topsoil have a standard deviation of 0.08 and a CV of ± 14%. The bottom 40-50cm of the fertilized soils have a standard deviation of 0.06 and a CV of ± 10%. A two-tailed T test between the U concentrations at both depths resulted in a P value of 0.78, suggesting no statistical difference between U concentrations in the fertilized soils at Waterman Farm based on their depth (Table D-6)

(Figure 11).

Samples from the “Wooster” sites feature 3 different rates of fertilizer application (0x,

1x, and 2x of 60.1 kg/ha P2O5 and 43.7 kg/ha K2O for corn plots) (Figure 15). The East

Badger site in Wayne County had an average U concentration of 0.36, 0.38, and 0.44

μg/g, a standard deviation of 0.02, 0.052, and 0.028, and a CV of ± 5.6%, ± 14%, and ±

6.3% for the 0x, 1x, and 2x application rates, respectively (Tale 13). The Western field in

Clark County had an average U concentration of 0.36, 0.36, and 0.34 μg/g, a standard deviation of 0.014, 0.014, and 0.026, and a CV of ± 4%, ± 4%, and ± 7.5%, for the 0x,

1x, and 2x application rates, respectively (Table D-7). The Northwest field in Wood

County had an average U concentration of 0.77, 0.69, and 0.76 μg/g, a standard deviation of 0.096, 0.079. and 0.079, and a CV of ± 13%, ± 11%, and ± 10%, for the 0x, 1x, and 2x application rates, respectively (Table D-8). T-tests could not be conducted for these plots as the sample sizes were too small (n=3 per application rate). 34

Due to the samples from Coshocton each having a different management practice, and their sample size only being n=1, could not be compared to each other. The “no till, no manure”, “pasture”, and “till, with manure” samples had U concentrations of 0.32, 0.31, and 0.28 μg/g, respectively (Figure 6) (Table D-9).

U concentrations from the DI water leaches from the sequential extraction process were found to be 2-3 orders of magnitude lower than the concentrations from the 0.1 M sodium bicarbonate and 10% HCl leaches. At Waterman Farm, the Cover Crop plots had an average of 5.6 x 10-4 and 4.2 x 10-4 μg/g at the topsoil and bottom 40-50cm depths, respectively (Table D-10). The Manure plots had an average U concentration of 8.0 x 10-4 and 5.5 x 10-4 μg/g at the topsoil and bottom sampling depths, respectively. Statistical inferences could not be made for the two previously mentioned plots due to their low sample size (n=1 or 2 per soil depth) (Table D-11). The Fallow plots had average U concentrations of 2.0 x 10-4 and 2.6 x 10-4 μg/g for the topsoil and bottom depths, respectively. The topsoil fallow plots had a standard deviation of 8.95e-5 and a CV of ±

44%. The bottom 40-50cm of the fallow plots had a standard deviation of 1.1 x 10-4 and a

CV of ± 42%. A two-tailed t-test between the U concentrations at both depths resulted in a P value of 0.59, suggesting no variation between U concentrations and sampling depth in the fallow plots (Table 19). The Fertilized plots had an average U concentration of 4.9 x 10-4 and 5.1 x 10-4 μg/g at the topsoil and bottom depths, respectively. The topsoil fertilized samples had a standard deviation of 0.00015 and a CV of ± 30%. The bottom

40-50cm of the fertilized plots had a standard deviation of 0.00020 and a CV of ± 38%.

Performing a t-test between the U concentrations at both depths resulted in a P value of

35

0.89, suggesting no variation between soil depth and U concentration at the fertilized plots (Table D-13) (Figure 12).

For the DI water leaches, samples from the “Wooster” locations are only reported as concentrations, due to their sample size being insufficient to report any statistics (n=1 per fertilizer application rate, with the exception of the East Badger site, where n=2) (Figure

16). At the East Badger site, U concentrations were 1.9 x 10-4 and 2.5 x 10-4, 2.7 x 10-4 and 2.2 x 10-4, and 2.3 x 10-4 and 3.4 x 10-4 μg/g for the 0x, 1x, and 2x application rates, respectively (Table D-14). At the Western site, U concentrations were 4.0 x 10-4, 2.8 x

10-4, 3.8 x 10-4 μg/g for the 0x, 1x, and 2x application rates, respectively (Table D-15). At the Northwest site, U concentrations were 1.4 x 10-3, 8.4 x 10-4, and 1.3 x 10-3 μg/g for the 0x, 1x, and 2x application rates, respectively (Table D-16).

Bicarbonate extractions for the Cover Crop plots at Waterman Farm yielded average U concentrations of 0.32 and 0.22 μg/g at 0-5 and 40-50cm depths, respectively (Table D-

17). Averages for the Manure plots were 0.27 and 0.28 μg/g for 0-5 and 40-50cm, respectively (Table D-18). Other statistics, such as the standard deviation, CV, and t-tests were not done for the aforementioned plots due to low sample sizes (n = 2 or 3 per soil depth). The Fallow plots had a U concentration average of 0.22 and 0.24 for topsoil and the bottom 40-50cm, respectively. U concentrations for the topsoil of the fallow plot had a standard deviation of 0.018, and a CV of ± 8.4%, and the bottom 40-50cm of the plots had a standard deviation of 0.028, and a CV of ± 12%. Performing a two-tailed t-test between fallow U concentrations at both soil depths resulted in a P value of 0.35, suggesting there is no variance between U concentrations at the two depths (Table D-19). 36

U concentrations for the fertilized plots at Waterman Farm averaged at 0.22 and 0.25

μg/g, had standard deviations of 0.04 and 0.03, and had a CV of ± 17 and ± 11% for the topsoil and bottom 40-50cm depths, respectively. A two-tailed t-test resulted in a P value of 0.28, suggesting no variation between U concentrations between the two sampled depths at the fertilized plots (Table D-20) (Figure 13).

Due to having a small sample size (n=1 or 2), samples from the “Wooster” locations are only reported as concentrations (Figure 17). At the East Badger site, U concentrations from 0.1 M sodium bicarbonate extractions were 0.13 and 0.09, 0.18 and 0.09, and 0.15 and 0.11 μg/g for the 0x, 1x, and 2x fertilizer application rates, respectively (Table D-21).

For the Western site, U concentrations were 0.20, 0.14, and 0.23 μg/g for the 0x, 1x, and

2x application rates, respectively (Table D-22). At the Northwest site, U concentrations were 0.58, 0.39, and 0.54 μg/g for the 0x, 1x, and 2x application rates, respectively

(Table D-23).

For the 10% HCl leach step of the sequential extractions, the cover crop plots at

Waterman Farm had an average U concentration of 0.46 and 0.41 μg/g for the topsoil and bottom 40-50 cm depths, respectively (Table D-24). The plots treated with manure had an average U concentration of 0.59 and 0.47 μg/g for the topsoil and bottom 40-50cm depths, respectively (Table D-25). The fallow plots had an average U concentration of

0.33 and 0.23 μg/g for the 0-5 and 40-50 cm depths, respectively. At 0-5 cm, the U concentrations for the fallow soils had a standard deviation of 0.07 and a CV of ± 22%.

The bottom 40-50 cm of the fallow soils had a standard deviation of 0.02 and a CV of ±

8%. A t-test between the U concentrations at both depths resulted in a P value of 0.084, 37 putting it within a 92% confidence interval, implying there could be some variation between U concentrations by depth at this plot (in terms of the U that is able to be extracted by 10% HCl) (Table D-26). For the fertilized plots at Waterman Farm, the average U concentrations were 0.36 for both the 0-5 cm and 40-50 cm depths. U concentrations at the 0-5 cm depth had a standard deviation of 0.08 and a CV of ± 22%.

At the 40-50 cm depth, U concentrations had a standard deviation of 0.12, and a CV of ±

33%. Performing a t-test between the two depths at the fertilizer plots resulted in a P value of 0.99, suggesting there is no variation between U concentrations at the two sampling depths (Table D-27) (Figure 14).

The East Badger site had U concentrations of 0.34 and 0.27, 0.42 and 0.29, and 0.42 and

0.36 μg/g for the 0x, 1x, and 2x fertilizer application rates, respectively, in terms of the U that was extracted from the 10% HCl leach step of the sequential extractions (Table D-

28). At the Western site, U concentrations were 0.25, 0.27, and 0.29 μg/g for the 0x, 1x, and 2x application rates, respectively (Table D-29). At the Northwest site, U concentrations were 0.67, 0.71, and 0.65 μg/g for the 0x, 1x, and 2x application rates, respectively (Table D-30). Statistics could not be performed on the aforementioned samples from the “Wooster” sites due to the sample size being n=1 (with the exception of

East Badger, where n=2) (Figure 18).

Sequential extractions were also used to evaluate the fraction of mobile U in the samples that were analyzed for total U through XRF. The overall amount of mobile U leached from the soils was found to be up to 3 orders of magnitude less than the U leached through the bicarbonate and acid extractions. At Waterman Farm, U concentrations from 38

DI water leaching ranged from 3.0 – 6.38 x 10-4 μg/g, which constitutes 0.007 – 0.015 % of the total U in the soil (Figure 7). At the East Badger site, U concentrations ranged from

2.22 – 3.43 x 10-4 μg/g (0.007 – 0.011 % total U) (Figure 8). At the Northwest site, U concentrations ranged from 8.4 x 10-4 - 1.25 x 10-3 μg/g (0.016 – 0.026 % total U) (Figure

9). At the Western site, U concentrations ranged from 2.75 – 3.99 x 10-4 μg/g (0.01 –

0.015 % total U) (Tables D-31, D-32, & D-34) (Figure 10).

For the 0.1M sodium bicarbonate sequential extractions, samples from Waterman Farm ranged from 0.20 – 0.32 μg/g U (4.6 – 6.7 % total U) (Figure 7). At the East Badger site,

U concentrations ranged from 0.09 – 0.11 μg/g (3.1 – 3.4 % total U) (Figure 8). At the

Northwest site, U concentrations ranged from 0.39 – 0.58 μg/g (7.6 – 10.9 % total U)

(Figure 9). At the Western site, U concentrations ranged from 0.14 – 0.23 μg/g (5.0 – 8.4

% total U) (Figure 10) (Table D-34).

For the 10% HCl sequential extractions, samples from Waterman Farm ranged from 0.44

– 0.47 μg/g U (9.0 – 15.3 % total U) (Figure 7). At the East Badger site, U concentrations ranged from 0.27 – 0.36 μg/g (8.8 – 11.5 % total U) (Figure 8). At the Northwest site, U concentrations ranged from 0.65 – 0.71 μg/g (12.6 – 13.2 % total U) (Figure 9). At the

Western site, U concentrations ranged from 0.32 – 0.36 μg/g (11.4 – 12.5 % total U)

(Figure 10) (Table D-34).

4f. Uranium-Bearing Minerals in Soils

Samples from the “Wooster” sites (East Badger, Western, and Northwest sites), as well as the fallow and fertilized plots at Waterman Farm were analyzed using SEM. Minerals

39 identified in soil samples through SEM analysis include zircons, monazite, ilmenite, biotite mica, rutile, and potassium feldspar (K-spar) (Table A-11). Minerals identified in soils from the East Badger site include monazite (Figure 37), zircon, and biotite mica. In the Western site, a zircon (Figure 38) was visually identified, while a cerianite (Figure

39) was identified through spot chemical analysis. In samples from the Northwest site, as well as from a fallow and fertilized plot at Waterman Farm, identified minerals included zircon (Figures 40 & 41) and monazite (Figures 42 & 43). The average size of the aforementioned minerals is about 14 μm. Mineral to grain size estimates were made by counting the pixels from the SEM images (Table A-15).

40

Chapter 5. Discussion

5a. Relationship of LOI and pH to U Concentrations in Soils

Uranium concentration correlations to LOI and pH were made using the Pearson product- moment correlation coefficient method (Pearson’s r) for soils from Waterman Farm only.

The Coshocton, East Badger, Northwest, and Western sites had sample sizes that were too small to apply these tests, and therefore, to draw significant conclusions.

Soil acidity generally has a negative correlation with U concentrations in soils analyzed.

Comparing pH to [U] from non-sequentially extracted samples (only leached by 10%

HCl) from Waterman Farm shows a weakly negative correlation (-0.43, p < 0.01) (Table

A-12). pH does not have a statistically significant relationship with U concentrations from sequentially leached samples from Waterman Farm (Table A-13). The generally acidic conditions of the soils have the potential to influence the sorption of U to cation exchange surfaces, resulting in more mobile U, which can lead to lower concentrations of extractable U (Giblin, et al., 1981; Johnson, et al., 2004; McKinley, et al., 1995;

Morrison, et al., 1995).

The LOI of the soils analyzed in this investigation has been found to have a generally positive correlation with U concentrations from soils that have been sequentially leached.

Sequential extraction U concentrations from soils at Waterman farm show a moderately positive relationship with LOI (r = 0.56, 0.57, 0.68 and p < 0.01, 0.01, 0.001 for DI

Water, bicarbonate, and HCl extractions, respectively) (Table A-13). These correlations suggest that soils with higher LOI will have higher U concentrations. This is probably

41 due to the fact that increased amounts of organic matter provide more negatively charged surface area for uranyl ions to bind (Vandenhove, et al., 2007).

5b. Extractions – Water vs. Base vs. Acid

A sequential extraction procedure was used to quantify the extent and strength of U binding to different surfaces on soils, involving the use of gradually more aggressive chemicals to remove the U bound to the soil (Johnson, et al., 2004). The use of continuously harsher or more severe leaching solutions allows for the quantification of the potentially environmentally available U fraction from the most soluble (water) to the least soluble (acid).

Soil Mineralogy

Monazite and zircon are two principal host minerals for U and Th in the Earth’s crust and occur as accessory minerals in crystalline rocks such as granites (Hurley & Fairbarn,

1957). These minerals were observed in all 6 soil samples, but in small abundances. The small amount of these minerals observed throughout the samples analyzed using SEM suggests that there are not many observable trace U solid phases present in these soils.

Monazite and zircon are insoluble minerals and would not be greatly affected by our relatively benign leaching solutions. Therefore, it is unlikely that they are the source of U in these soils, even in our acid leached samples. The concentrations of U that have been quantified through these chemical leaches suggest that the U that is solubilized is present from other sources, such as cation-exchange sites (oxides/hydroxides (Fe, Mn) and organic matter) (Table A-15).

42

Comparison of Sequential Extractions

The DI water extractions yielded U concentrations that were two to three orders of magnitude less than concentrations obtained from the bicarbonate and acid extraction methods (Figures 7 – 10), averaging 5.6 x 10-4 μg/g (Table D-31). The 0.1M bicarbonate leaches yielded an average U concentration of 0.26 μg/g (Table D-31). The 10% HCl leaches proved to be the step in the sequential extraction process that yielded the highest

U concentrations, averaging at 0.47 μg/g (Table D-31).

Compared to the total U concentrations from samples analyzed by XRF (ranging from 2.7

– 5.4 μg/g and averaged at 4.0 μg/g U), water-soluble U accounts for only ~ 0.013 % of the total U in the soils analyzed (Table D-32). The bicarbonate extractions yielded ~

6.1% of the total U observed (Figures 7 -10) (Table D-32). The sodium bicarbonate leaches increase the pH of the soils to ~ 8. The addition of carbonate and bicarbonate ions also leads to the formation of U(VI) complexes (Zhou & Gu, 2005). Because of this, the solubility of U should be increased during this sequential extraction. The 10% HCl leaches account for ~ 11.6% of the total U by XRF (Figures 7 – 10) (Table D-32). Not surprisingly, the aforementioned comparison indicates that most of the U bound to the soils that was able to be chemically extracted was acid soluble. This is potentially the fraction that is bound to surfaces of such constituents as oxides/hydroxides (Fe, Mn) and organic matter.

43

5c. U Concentrations in Ohio Soils: Based on a Datamining and GIS Approach

As part of this study, it was decided to use the very large datasets on U concentrations in

Ohio soils to investigate if the variation in soil U is related to the past glacial history, underlying bedrock, and/or land use. The majority of the NURE dataset (1000 sampled sites total) is located in the last glacial maximum (LGM) unglaciated portion of Ohio, and is therefore higher in sample size (78% of sampled locations) compared to the U concentrations taken from locations that were sampled in the LGM glaciated portion of

Ohio (22% of sampled locations) (Figure 27) (Table D-36). The USGS topsoil dataset (69 sampled sites total), which spans throughout all of Ohio, has greater sample size in the

LGM glaciated portion of Ohio, which contains 68% of the sampled locations, compared to the LGM unglaciated portion which only contains 32% of the sampled locations. The total U concentrations from this work can be compared to both the NURE and USGS data

(Smith, et al., 2013; U.S. Geological Survey, 2004) obtained for both agricultural and non-agricultural lands. Correlations made between sequential U extractions and total U data obtained for agricultural and non-agricultural soils, and in the case of the NURE data, stream sediments as well, also indicate that total U data can be compared to leached

U data in this investigation (Table A-14). This work compares the following data:

1) LGM unglaciated locations of NURE to USGS

2) LGM glaciated USGS to LGM unglaciated USGS

3) NURE and USGS data to underlying bedrock type

4) USGS data to Ohio land use data from the USGS Gap Analysis Program

44

Impact of Glacial History

The geographic distribution of the NURE dataset is concentrated in the East/Southeast portion of Ohio, so holistic interpretations cannot be made for the entire state. However, due to the greater density of sampling locations in the NURE dataset compared to the

USGS topsoil dataset, more informed interpretations can be made in regard to U concentrations in Ohio’s LGM unglaciated area (Figure 27).

When comparing the NURE and USGS dataset soil U concentrations in the LGM unglaciated portion of Ohio, concentrations from the NURE dataset range from 1.9 – 19.0

μg/g and has an average of 4.1 μg/g with a CV of ± 24%. The USGS U concentrations range from 2.7 – 5.2 μg/g, and average at 3.6 μg/g with a CV of ± 17% (Table D-36). The

LGM glaciated portion of Ohio has U concentrations ranging from 1.7 – 9.0 μg/g and has an average concentration (4.4 μg/g with a CV of ± 29%) while as noted above, the concentration of the LGM unglaciated portion is 3.6 μg/g with a CV of 17% (Table D-

36). The higher average U concentrations in Ohio’s LGM glaciated portion from the

USGS dataset indicates that higher U concentrations can be expected in soils that have been glaciated when compared to the non-glaciated soils. This may be due to the incorporation of more U-rich materials from either far field or short field glacial transport and deposition. This is an important observation in the understanding of U distribution in

Ohio soils.

45

The Impact of Underlying Bedrock

Geologic formations that are associated with their specific bedrock type are identified in

Table A-16. The highest average soil U concentrations for the NURE data are samples overlying mudstone (4.5 μg/g) and siltstone (4.4 μg/g) units, which are only present in the

East/Southeast portion of Ohio (Figure 28) (Tables A-16 & D-38). The highest average U concentrations in the USGS dataset occur in soils overlying dolostone and black shale units (4.9 and 4.4 μg/g, respectively) (Figure 34) (Table A-16 & D-38). Placing the samples in this investigation in the context of Ohio’s geology, soil samples that overlie a limestone bedrock have the highest concentrations (within their dataset) of total U at an average of 4.6 μg/g, followed by dolostone (4.0 μg/g) and shale (3.0 μg/g). This does not necessarily imply that the soils are derived from the underlying bedrock, but it does suggest a further evaluation of the relationship of U in soils to U concentrations in Ohio bedrock.

The four available datasets (NURE, USGS, Barnes et al., 2020, and this work) can be best compared when considering U concentrations of soils that overlie shales, due to the fact that a portion of all four datasets overlie shale. The highest average U concentration in the soils that overlie shales belongs to the NURE dataset at 3.9 μg/g, followed by the

USGS, the average concentration observed in this investigation, and a glacial till in southwest Ohio sampled by Barnes, et. al, 2020 (3.8, 3.1, and 1.2 μg/g respectively)

(Table A-16, D-38, & D-39). In the portions of the datasets that overlie shales, the variations within the NURE and USGS dataset U concentrations are not statistically different from each other (p > 0.05) but comparing the NURE and USGS datasets to both 46 the U concentrations observed in this work and Barnes, et al., 2020, as well as comparing the values in this investigation with the Barnes 2020, et al. dataset show statistical difference in their U concentration variations.

The Impact of Land Use

The USGS topsoil dataset was the only source of data able to be used as a comparison to the land use/cover data (Figure 35) (Table D-37), due to its sampled locations representing the entire state of Ohio. The NURE dataset could not be considered in regard to comparing land use and U concentrations due to the incompatibility between data set formats and the high density of sampled points that are concentrated in a smaller portion of the state of Ohio preventing visual counting from being done properly.

The highest average U concentration was observed in agricultural areas (4.5 μg/g), followed by developed (4.0 μg/g) and undeveloped areas (3.6 μg/g). These differences in

U concentrations in regard to land use may not be directly attributed to a single factor such as anthropogenic activities (such as fertilizer addition) . As noted above, the U in

Ohio soils can naturally vary in concentration due to factors like soil type, the composition and origin of the till in glaciated areas, the underlying geology, etc.

5d. Comparison of Agricultural Management Practices in Relation to U Concentrations in Agricultural Soils

Management Practices at Waterman Farm

Comparison of U concentrations for soils at Waterman Farm with their different management practices do not show a straightforward relationship, due to the low

47 variability in U concentrations throughout this soil type. Samples that were leached using only 10% HCl (non-sequential) show the highest median concentrations of U in plots that were treated with fertilizer (244 kg/ha N, with 184 kg as Urea, and 60 kg as NPK) at 0.55

± 0.006 μg/g. The fallow and manure plots have the same median U concentration, 0.52 ±

0.006 and ± 0.007 μg/g, respectively, and the cover crop plots have the lowest concentrations at 0.48 ± 0.008 μg/g (Figure 11). Given the variability of these measurements, it is not possible to statistically tell the difference between these values.

At Waterman Farm, samples leached using DI water as part of the sequential leach process have U concentrations three orders of magnitude lower than concentrations observed in the bicarbonate and acid leaches. Among the four management practices, the manure plots show the highest median water-leachable U concentrations at 6.5 x 10-4 ±

2.0 x10-5 μg/g, while the fertilizer, cover crop, and fallow plots decrease to 5.0 x 10-4 ±

2.0 x 10-5, 4.3 x 10-4 ± 2.0 x 10-5, and 2.5 x 10-4 ± 2.0 x 10-5 μg/g, respectively (Figure

12). Samples from Waterman Farm leached with 0.1 M sodium bicarbonate solution show the least amount of variation between management practices, with both the cover crop and manure plots having median U concentrations of 0.27 ± 0.005 μg/g, followed by the fertilizer and fallow plots having median concentrations of 0.23 ± 0.005 and 0.22 ±

0.004 μg/g, respectively (Figure 13). Samples leached with 10% HCl as the last step of the sequential extraction process show the manure plots having the highest median U concentrations at 0.52 ± 0.007 μg/g, followed by the median concentration for the cover crop plots at 0.44 ± 0.002 μg/g. The fertilizer and fallow plots have the lower median U concentrations at 0.38 ± 0.007 and 0.27 ± 0.003 μg/g, respectively (Figure 14). When

48 comparing all four management practices at Waterman Farm, the average and median U concentrations are not very different (Table D-33) and could be attributed to natural variations of U in this location/soil type.

U and Soil Depth

One of the hypotheses in this investigation was to test if U concentrations would decrease with soil depth. The non-sequential HCl extractions at Waterman Farm showed average

U concentrations 0.48 ± 0.009 and 0.51 ± 8.0 x 10-5 μg/g, 0.50 ± 0.006 and 0.46 ± 6.0 x

10-5 μg/g, 0.54 ± 0.006 and 0.55 ± 6.0 x 10-5 μg/g, and 0.57 ± 0.007 and 0.58 ± 6 x 10-5

μg/g for the for the 0-5 and 40-50cm depths of the cover crop, manure, fallow, and fertilized plots, respectively (Tables D-2 to D-5). The only Waterman Farm plots that could be compared in terms of U concentrations from sequential extractions were the fallow and fertilized plots, due to the plots for other two management practices (cover crop and manure) having too small of a sample size. The average water-soluble U concentrations for the fallow and fertilized plots were 2.0 x 10-4 ± 9.9 x 10-6 and 2.6 x 10-

4 ± 1 x 10-7 μg/g, and 4.9 x 10-4 ± 9 x 10-6 and 5.1 x 10-4 ± 7 x 10-8 μg/g for the 0-5 and

40-50cm depths, respectively (Tables D-12 & D-13). The average readily leachable U concentrations (through 0.1M bicarbonate extraction) for the fallow and fertilized plots were 0.22 ± 0.004 and 0.24 ± 4.0 x 10-5 μg/g, and 0.22 ± 0.003 and 0.25 ± 4 x 10-5 μg/g for the 0-5 and 40-50cm depths, respectively (Tabled D-19 & D-20). The average acid- soluble U concentrations for the fallow and fertilized plots were 0.33 ± 0.005 and 0.23

±2 x 10-5 μg/g, and 0.36 ± 0.004 and 0.36 ± 4 x 10-5 μg/g for the 0-5 and 40-50cm depths, respectively (Tables D-26 & D-27). The U concentrations between the 0-5cm and 49

40-50cm depths of the sampled soils have been shown to not be statistically different.

The inability to detect differences between 0-5 and 40-50cm depths may be due to in part to the limited sample size (n = 2 up to 8 for each depth) used in this portion of the study.

Fertilizer Application Rate at Wooster Sites

The management practices at the “Wooster” sites (East Badger Farm, Northwest

Agricultural Research Station, and Western Agricultural Research Station) involved different rates of fertilizer application (0x, 1x, and 2x the calculated nutrient removal rate for corn). While grouped together in terms of management practices, samples from the different locations cannot be compared to one another due to their different soil types.

The Northwest site had the highest overall U concentrations among the three locations, with U concentrations for plots treated with 0x, 1x and 2x fertilizer application rates averaging at 0.76, 0.77, and 0.69 μg/g, respectively. Average U concentrations at the East

Badger site for plots treated with 0x, 1x, and 2x the fertilizer application rate were 0.36,

0.38, and 0.44 μg/g, respectively. At the Western site, the average U concentrations were

0.36, 0.36, and 0.34 μg/g for plots treated with 0x, 1x, and 2x the fertilizer application rate, respectively. Figures 16-18 show U concentrations for the sequential extractions at the Wooster sites. Although the sample size is low, the HCl-only (Figure 15) and sequential extractions (n = 3 and n = 1 or 2, respectively) (Figures 16-18) indicate that the application of fertilizer has not increased the U concentrations in these soils.

50

Coshocton Soils

In the Coshocton area, the “No Till No Manure” sample had the highest concentration of

U at 0.32 μg/g, while, while the Pasture and “Till, with Manure” samples had lower concentrations, 0.25 and 0.24 μg/g, respectively (Table D-9). Due to small sample size

(n=1 for each management practice in Coshocton), rigorous conclusions cannot not be drawn based on U concentrations in regard to how the soil is treated.

Comparison of Results to Previous Work

Yamaguchi, et al., 2009 observed U concentrations in Japanese agricultural fields after long-term P-fertilizer application to range from 1.11 – 1.76 μg/g and concluded that most of the fertilizer-derived U was incorporated into the soil organic matter or on poorly crystalline Fe/Al minerals in the surface soil. Huhle, et al., 2008 observed U concentrations of 0.96 and 0.58 μg/g at the 20 and 50cm depths in German farmland soils

(pH of 5.8-5.9), respectively, while forest soils (pH of 3.7-3.9) had lower U concentrations (0.53 and 0.35 μg/g for the 20 and 50cm depths, respectively). The study speculated that the pH value of the soils has an influence on the binding of U to organic matter in both of the sites. Zielinski, et al., 2006 observed U concentrations ranging from

0.3 – 1.4 μg/g in two pasture and one grassland site Central Florida. Among the three sites, the amount of U extracted using 0.1M sodium bicarbonate solution ranged from 5 –

42% of the total U, and averaging at about 16%, higher than any of the sequential extraction U concentrations observed in this investigation (range of 0.007 – 15.3 % total

U), and higher than the average of 11.6 % of the total U being leached using 10 % HCl

51

(Table D-32). These authors concluded that the reason for these elevated fertilizer- derived and natural U concentrations is due to the U being efficiently sorbed onto soil organic matter (which ranged from 0.2- 52.3 μg/g and averaged at 11.7 μg/g). Taylor,

2007 observed an increase in U concentrations when comparing historical (0.62 – 2.34

μg/g U) to present-day (1.69 – 3.54 μg/g U) soil samples in New Zealand pastoral farming areas. The largest increase in total U concentrations among the soils in this study was observed in a soil containing peat (from 0.79 – 2.48 μg/g), which reflected the high inputs of superphosphate fertilizer applied to the soil (600 kg/ha per year), and indicated that U is strongly sorbed by organic surfaces such as peat under reducing conditions.

In comparing the observations from the aforementioned studies with the results of this investigation, a common theme is recognized, that soil organic matter acts as an important reservoir for U in agricultural soils. The data indicate no excess solubilizable U has accumulated, even from 2x fertilizer additions at the Wooster locations. A possible explanation is that any U that is added to agricultural soil via fertilizer addition is rapidly solubilized and lost.

5e. Comparison of Ohio U Soil Chemistry to US and Global Inventories

The comparison between the inventories of soil U in the United States and Ohio were from the Smith, et al., 2013 study for the USGS, which gave an average topsoil U concentration of 2.1 μg/g for the contiguous US (Smith, et al., 2013). From the same dataset, the average U concentration for soils in Ohio was 4.1 μg/g (Figure 19). The LGM unglaciated portion of Ohio has a lower U concentration (3.6 μg/g) compared to the LGM glaciated portion of Ohio (4.4 μg/g). The total U concentrations observed in this 52 investigation from agricultural locations throughout Ohio ranged from 2.7 – 5.4 μg/g

(Table D-1). Placing these soil U data in context with United States and global inventories, Ohio tends to have higher soil U concentrations compared to both the contiguous US and the upper continental crust (2.7 μg/g) (Rudnick & Gao, 2003).

53

Chapter 6. Conclusions

6a. Summary of Research

1. The variation in solubilizable U concentrations in agricultural soils and their non-

agricultural equivalents are similar with no apparent buildup of U in the

agricultural soils.

2. No statistical difference was found in U concentrations between the topsoil (0-

5cm) and the bottom 40-50cm of soils sampled at Waterman Farm.

3. Most of the solubilizable U was found in the acid-soluble fraction (~ 11.6%),

followed by the base-soluble fraction (~ 6.1%) and the water-soluble fraction (~

0.013%) of soils that were analyzed for total U (ranging from 2.7 – 5.4 μg/g).

4. U concentrations from the NURE dataset are higher (4.1 μg/g) than

concentrations from the USGS dataset (3.6 μg/g) in the LGM unglaciated portion

of Ohio. In the USGS dataset, U concentrations in the LGM glaciated portion of

the state are 4.4 μg/g, clearly higher than the U concentrations in the LGM

unglaciated portion. Based on these data, soils in the LGM glaciated portion of

Ohio may be expected to have higher U concentrations than soils in the LGM

unglaciated portion. These values are within the range of values we observed for

total U (2.7 – 5.4 μg/g) in this investigation. The difference between the LGM

unglaciated and LGM glaciated values also suggest an influx of U-rich materials

by previous glacial transport.

54

5. The highest U concentrations in the NURE data were observed for soils that

overlie a mudstone (4.5 μg/g) and siltstone (4.4 μg/g) bedrock. The USGS dataset

had the highest U concentrations in soils that overlie dolostone (4.9 μg/g) and

black shale (4.4 μg/g) units. Samples from this investigation had the highest U

concentrations in soils that overlie limestone (4.6 μg/g).

6. The highest soil U concentrations in the USGS dataset were observed in

agricultural areas (4.5 μg/g), followed by developed (4.0 μg/g) and undeveloped

areas (3.6 μg/g) for the land use data. The differences in U concentrations in terms

of land use may not necessarily directly attributed to a single factor, such as

fertilizer addition.

6b. Future Work

The work in this investigation can be expanded upon through modifications of the aforementioned methods of chemical analyses. These modifications can include:

1. Determining the U concentrations and its solubility in fertilizers used in

agricultural soils in Ohio.

2. Analysis of other elements and compounds that could contribute to the mobility of

- 2- U (Fe, Ca, P, HCO3 , CO3 , etc.) and of elements belonging to U’s decay series

(Th, Pa, Ra, Rn). The analysis of U isotopes in soils would also aid in determining

the U’s source.

3. Use of different and/or stronger (in the case of acid leaches) chemicals for U

extractions. 55

4. Obtaining larger amounts of samples from the “Wooster” study sites so that better

statistics can be done to determine if fertilizer application rates affect U

concentrations in Ohio soils.

56

Works Cited

Alloway, B. J. (1995). Heavy metals in soils. Glasgow: Blackie Academic & Professional.

Altschuler, Z. S. (1973). The weathering of phosphate deposits - gechemical and environmental aspects. In E. J. Griffith, A. Beeton, J. M. Spencer, & D. T. Mitchell, Environmental phosphorus handbook (pp. 33-96). New York: Wiley-Interscience.

Bargar, J. R., Reitmeyer, R., & Davis, J. A. (1999). Spectroscopic confirmation of uranium(VI)- carbonate adsorption complexes on hematite. Environmental Science & Technology, 14, 33, 2381-2484.

Barnes, M., McLeod, C. L., Chappell, C., Faraci, O., Gibson, B., & Krekeler, M. P. (2020). Characterizing the geogenic background of the Midwest: a detailed mineralogical and geochemical investigation of a glacial till in southwestern Ohio. Environmental Earth Sciences(79), 159.

Blanco-Canqui, H., & Lal, R. (2007a). Impacts of Long-Term Wheat Straw Management on Soil Hydraulic Properties under No-Tillage. Soil Science Society of America Journal, 1166- 1173.

Blanco-Canqui, H., & Lal, R. (2007b). Soil structure and organic carbon relationships following 10 years of wheat straw management in no-till. Soil & Tillage Research, 240-254.

Bolivar, S. L. (1980). An Overview of the National Uranium Resource Evaluation Hydrogeochemical and Stream Sediment Reconnaissance Program. Los Alamos, NM; Grand Junction, CO: Los Alamos Scientific Laboratory, U.S. Department of Energy.

Burnett, W. C., & Kim, K. H. (1985). 226Ra in phosphate nodules from the Peru/Chile seafloor. Geochimica et Cosmochimica, 1073-1081.

Cathcart, J. B. (1978). Uranium in Phosphate Rock. Washington, D.C.: U.S. Govt. Print. Off.

Clark, P. U., Dyke, A. S., Shakun, J. D., Carlson, A. E., Clark, J., Wohlfarth, B., . . . McCabe, A. M. (2009). The Last Glacial Maximum. Science, 710-714.

Clark, S. B., Johnson, W. H., Malek, M. A., Serkiz, S. M., & Hinton, T. G. (1996). A comparison of sequential extraction techniques to estimate geochemical controls on the mobility of fission product, actinide, and heavy metal contaminants in soils. Radiochimica Acta, 173- 179.

57

Coyte, R. M., Jain, R. C., Srivastava, S. K., Sharma, K. C., Khalil, A., Ma, L., & Vengosh, A. (2018). Large-Scale Uranium Contamination of Groundwater Resources in India. Environmental Science and Technology Letters, 341-347. doi:10.1021/acs.estlett.8b00215

Crançon, P., & van der Lee, J. (2003). Speciation and mobility of uranium(VI) in humic- containing soils. Radiochimica Acta, 673-679.

Da Conceição, F. T., & Bonotto, D. M. (2006). Radionuclides, heavy metals and fluorite incidence at Tapira phosphate rocks, Brazil, and their industrial (by) products. Environmental Pollution, 232-243.

Diaz, M. A., Fortner, S. K., & Lyons, W. B. (in review). High resolution concentration-discharge relationships in managed watersheds: a 30+ year analysis. Journal of Hydrogeology.

Duff, M. C., & Amrhein, C. (1996). Uranium(VI) Adsorption on Goethite and Soil in Carbonate Solutions. Soil Science Society of America Journal, 1393-1400.

Duff, M. C., Mason, C. F., & Hunter, D. B. (1998). Comparison of acid and base leach for the removal of uranium from contaminated soil and catch-box media. Canadian Journal of Soil Science, 675-683.

Eckstein, Y., Lewis, V. E., & Bonta, J. V. (2007). Chemical evolution of acid precipitation in the unsaturated zone of the siltstones and shale of central Ohio. Hydrogeology Journal, 1489-1505.

El Ghawi, U., Patzay, G., Vajda, N., & Bodizs, D. (1999). Analysis of selected fertilizers imported to Libya for major, minor, trace and toxic elements using ICP-OES and INAA. Journal of Radioanalytical Nuclear Chemistry, 693-701.

EPA. (2019, March 30). Radionuclides Rule: A quick Reference Guide. Retrieved from nepis.epa.gov: https://nepis.epa.gov/Exe/ZyPDF.cgi?Dockey=30006644.txt

Fortner, S. K., Lyons, W. B., Carey, A. E., Shiptalo, M. J., Welch, S. A., & Welch, K. A. (2012). Silicate weathering and CO2 consumption within agricultural landscapes, the Ohio- Tennessee River Basin, USA. Biogeosciences, 941-955.

Fulford, A. M., & Culman, S. W. (2018). Over-Fertilization Does Not Build Soil Test Phosphorus and Potassium in Ohio. Agronomy Journal, 56-65.

Ganesh, R., Robinson, K. G., Reed, G. D., & Sayler, G. S. (1997). Reduction of Hexavalent Uranium from Organic Complexes by Sulfate- and Iron-Reducing Bacteria. Applied and Environmental Microbiology, 4385-4391.

58

Garrett, R. G., Hall, G. E., Vaive, J. E., & Pelchat, P. (2009). A water-leach procedure for estimating bioaccessibility of elements in soils from transects across the United States and Canada. Applied Geochemistry, 1438-1453.

Giblin, A. M., Batts, B. D., & Swaine, D. J. (1981). Laboratory simulation studies of uranium mobility in natural waters. Geochimica Cosmochimica Acta, 699-709.

Godinez, M., Iturbe, J. L., Ordoñez, E., & Solache-Rios, M. (1995). Determination of radium- 226 in phosphate fertilizers and gypsum by gamma-ray spectrometry. International Journal of Environment and Pollution, 195-200.

Goulding, K. T. (2016). Soil acidification and the importance of liming agricultural soils with particular reference to the United Kingdom. Soil Use and Management, 390-399.

Goyette, J. -O., Bennett, E. M., & Maranger, R. (2018). Low buffering capacity and slow recovery of anthropogenic phosphorus pollution in watersheds. Nature Geoscience, 921- 925. doi:10.1038/s41561-018-0238-x

Grassi, V., Toran, L., Stevens, G., Bednar, A., & Young, C. (2005). A Case Study on Uranium Mobility in Groundwater at the Dupont Chambers Work Site. WM Symposia. Tuscon: Waste Management.

Guzmán, E. R., Solache Rios, M., Iturbe Garcia, J. L., & Ordoñez Regil, E. (1995). Uranium in phosphate rock and derivatives. Journal of Radioanalytical Nuclear Chemistry, 301-306.

Hamamo, H., Landsberger, S., Harbottle, G., & Panno, S. (1995). Studies of radioactivity and heavy metals in phosphate fertilizer. Journal of Radioanalytical Nuclear Chemistry, 331- 336.

Heiland, K. (1986). Analysis of trace elements in phosphate rock of various deposits and industrially processed phosphates by INAA. Universität Hamburg: Fachbereich Chemie.

Heiri, O., Lotter, A. F., & Lemcke, G. (2001). Loss on ignition as a method for estimating organic and carbonate content in sediments: reproducibility and comparability of results. Journal of Paleolimnology, 101-110.

Huang, F. Y., Brady, P. V., Lindgren, E. R., & Guerra, P. (1998). Biodegradation of Uranium - Citrate Complexes: Implications for Extraction of Uranium from Soils. Environmental Science & Technology, 379-382.

Huhle, B., Kummer, S., Stadler, S., & Merkel, B. (2008). Mobility of Uranium from Phosphate Fertilizers in Sandy Soils. In L. J. De Kok, & E. Schnug, Loads and Fate of Fertilizer- Derived Uranium (pp. 47-55). Leiden: Backhuys Publishers.

59

Hurley, P. M., & Fairbarn, H. W. (1957). Abundance and Distribution of Uranium and THorium in Zircon, Sphene, Apatite, Epidote, and Monazite in Granitic Rocks. Transactions, American Geophysical Union , 939-944.

Jacinthe, P. A., Lal, R., & Kimble, J. M. (2002). Carbon budget and seasonal carbon dioxide emission from a central Ohio Luvisol as influenced by wheat residue amendment. Soil Tillage Res., 147-157.

Jasinski, S. M. (2013). 2011 Minerals Yearbook. U.S. Department of the Interior; U.S. Geological Survey.

Johnson, W. H., Buck, B. J., Brogonia, H., & Brock, A. L. (2004). Variations in Depleted Sorption and Solubilty with Depth in Arid Soils. Soil & Sediment Contamination, 533- 544.

Kalra, Y. P., Agrawal, H. P., Allen, E., Ashworth, J., Audesse, P., Case, V. W., . . . Yeung, P. (1995). Determination of pH of Soils by Different Methods: Collaborative Study. Journal of AOAC International, 310-324.

Kelley, G. E., Edwards, W. M., Harrold, L. L., & McGuinness, J. L. (1975). Soils of the North Appalachian Experimental Watershed, Coshocton, Ohio. Washington D.C.: USDA Misc. Publ. 1296.

Kling, C. L., Panagopoulos, Y., Rabotyagov, S. S., Valcu, A. M., Gassman, P. W., Campbell, T., . . . Rabalais, N. N. (2014). LUMINATE: linking agricultural land use, local water quality and Gulf of Mexico hypoxia.

Kohler, M., Weiland, E., & Leckie, J. O. (1992). Metal ligand interactions during sorption of uranyl and neptunyl on oxides and silicates. In Y. K. Kharaka, & A. S. Maest (Ed.), Proceedings of the 7th International Symposium on Water-Rock Interaction (pp. 51-54). Park City, Utah: A.A. Balkema.

Kratz, S., Knappe, F., Rogasik, J., & Schnug, E. (2008). Uranium Balances in Agroecosystems. In L. J. De Kok, & E. Schnug (Eds.), Loads and Fate of Fertilizer-derived Uranium (pp. 179-189). Leiden, The Netherlands: Backhuys Publishers.

Kurttio, P., Auvinen, A., Salonen, L., Saha, H., Pekkanen, J., Mäkeläinen, I., . . . Komulainen, H. (2002). Renal Effects of Uranium in Drinking Water. Environmental Health Perspectives, 337-342.

Liesch, T., Hinrichsen, S., & Goldscheider, N. (2015). Uranium in groundwater - Fertilizers versus geogenic sources. Science of the Total Environment, 981-995.

Lyons, W. B., Gardner, C. B., Welch, S. A., & Israel, S. (2020). Uranium in Ohio, USA Surface Waters: Implications for a Fertilizer Source in Waters Draining Agricultural lands. Scientific Reports.

60

MacDonald, G. K., Bennett, E. M., Potter, P. A., & Ramankutty, N. (2011). Agronomic phosphorus imbalances across the world's croplands. Proceedings of the National Academy of Sciences of the United States of America (PNAS), 3086-3091.

Makweba, M. M., & Holm, E. (1993). The natural radioactivity of the rock phosphates, phosphatic products and their environmental implications. Science of the Total Environment, 99-110.

McBride, M. B., & Spears, G. (2001). Trace element content of selected fertilizers and dairy manures as determined by ICP-MS. Comm. Soil Sci. Plant Anal., 139-156.

McKinley, J. P., Zachara, J. M., Smith, S. C., & Turner, G. D. (1995). The Influence of Uranyl Hydrolysis and Multiple Site-Binding Reactions on Adsoprtion of U(VI) to Montmorillonite. Clay and Clay Minerals, 586-598.

Menzel, R. G. (1968). Uranium, radium and thorium content in phosphate rocks and their possible radiation hazard. J. Agr. Food Chem., 231 - 234.

Michalak, A. M., ANderson, E. J., Beletsky, D., Boland, S., Bosch, N. S., Bridgeman, T. B., . . . Zagorski, M. A. (2013). Record-setting algal bloom in Lake Erie caused by agricultural and meteorological trends consistent with expected future conditions. Proceedings of the National Academy of Sciences of the United States of America (PNAS), 6448-6452.

Morrison, S. J., Spangler, R., & Tripathi, V. S. (1995). Adsorption of uranium(VI) on amorphous ferric oxyhydroxide at high concentrations of dissolved carbon(IV) and sulfur(VI). Journal of Contaminant Hydrology, 333-346.

Moyes, L. N., Parkman, R. H., Charnock, J. M., Vaughan, D. J., Livens, F. R., Hughes, C. R., & Braithwaite, A. (2000). Uranium Uptake from Aqueous Solution by Interaction with Goethite, Lepidocrocite, Muscovite, and Mackinawite: An X-ray Adsorption Spectroscopy Study. Environmental Science & Technology, 1062-1068.

Nesbitt, H. W., & Young, G. M. (1982). Early climates and plate motions inferred from major element chemistry of lutites. Nature, 715-717.

Nicholson, S. W., Dicken, C. L., Horton, J. D., Labay, K. A., Foose, M. P., & Mueller, J. A. (2005). Preliminary integrated Geologic Map Databases for the United States: Kentucky, Ohio, Tennessee, and West Virginia. Reston: U.S. Geological Survey. Retrieved from https://mrdata.usgs.gov/geology/state/metadata/oh.html

Otero, N., Vitòria, L., Soler, A., & Canals, A. (2005). Fertiliser characterisation: Major, trace, and rare earth elements. Applied Geochemistry, 1473-1488.

Owens, L. B., Shipitalo, M. J., & Bonta, J. B. (2008). Water quality response times to pasture management changes in small and large watersheds. Journal of Soil and Water Conservation, 292-299.

61

Pantelica, A. I., Salagean, M. N., Georgescu, I. I., & Pincovschi, E. T. (1997). INAA of some phpsphates used in fertilier industries. Journal of Radioanalytical Nuclear Chemistry, 261-264.

Pelley, J. (2016). Taming Toxic Algae Blooms. ACS Central Science, 270-273.

Pinney, S. M., Freyburg, R. W., Levine, G. H., Brannen, D. E., Mark, L. S., Nasuta, J. M., . . . Wones, R. (2003). Health Effects in Community Residents Near a Uranium Plant at Fernald, Ohio, USA. International Journal of Occupational Medicine and Environmental Health, 139-153.

Robarge, W. P., Boos, D., & Proctor, C. (2004). Determination of trace metal content of fertilizer source materials produced in North America. Environmental Impact of Fertilizer on Soil and Water. ACS Symposium Series 872 (pp. 75-89). Washington, D.C.: American Chemical Society.

Rothbaum, H. P., McGaveston, D. A., Wall, T., Johnston, A. E., & Mattingly, G. G. (1979). Uranium Accumulation in Soils from Long-Continued Applications of Superphosphate. Journal of Soil Science, 147-153.

Rudnick, R. L., & Gao, S. (2003). Composition of the Continental Crust. In H. D. Holland, & K. K. Turekian (Eds.), Treatise on Geochemistry (Vol. 3, pp. 1-64). Elsevier Science.

Sam, A. K., Ahamed, M. O., El Khangi, F. A., El Nigumi, Y. O., & Holm, E. (1999). Radiological and chemical assessment of Uro and Kurun rock phosphates. J. Environ. Radioact., 65-75.

Saroa, G. S., & Lal, R. (2003). Soil restorative effects of mulching on aggregation and carbon sequestration in a Miamian soil in central Ohio. Land Degradation & Development, 481- 493.

Saroa, G. S., & Lal, R. (2004). Mulching effects on phosphorus and sulfur concentrations in a Miamian soil in central Ohio, USA. Land Degradation & Development, 351-365.

Saueia, C. R., & Mazilli, B. P. (2006). Distribution of natural radionuclides in the production and use of phosphate fertilizers in Brazil. J. Envrion. Radioact., 229-239.

Schultz, M. K., Burnett, W., Inn, K. W., & Smith, G. (1998). Geochemical partitioning of actinides using sequential chemical extractions: Comparison to stable elements. Journal of Radioanalytical and Nuclear Chemistry, 234, 251-256.

Shrestha, R. K., Lal, R., & Rimal, B. (2013). Soil carbon fluxes and balances and soil properties of organically amended no-till corn production systems. Geoderma, 177-185.

62

Smith, D. B., Cannon, W. F., Woodruff, L. G., Solano, F., & Ellefsen, K. J. (2014). Geochemical and Mineralogical Data for Soils of the Conterminous United States Geological Survey Open-File Report. Reston: U.S. Geological Survey. Retrieved from http://dx.doi.org/10.3133/ofr20141082

Smith, D. B., Cannon, W. F., Woodruff, L. G., Solano, F., Kilburn, J. E., & Fey, D. L. (2013). Geochemical and Mineralogical Data for Soils of the Coterminous United States: U.S. Geological Survey Data Series 801. Reston, VA: U.S. Geological Survey.

Soil Environmental Chemistry Program, The Ohio State University. (n.d.). Organic Carbon Determination by Reduction of Dichromate, Version 8.

Soil Survey Staff, NRCS, & USDA. (2016). Web soil survey. Retrieved from Natural Resources Conservation Service: https://websoilsurvey.sc.egov.usda.gov/

Spalding, R. F., & Sackett, W. M. (1972). Uranium in Runoff from the Gulf of Mexico Distributive Province: Anomalous. Science, 629-631.

Sparks, D. L. (2003a). Chemistry of Soil Organic Matter. In Environmental Soil Chemistry, 2nd Edition (pp. 75-113). San Diego: Academic Press.

Sparks, D. L. (2003b). Ion Exchange Processes. In Environmental Soil Chemistry, 2nd Edtion (pp. 187-205). San Diego: Academic Press.

Suh, S., & Yee, S. (2011). Phosphorus use-efficiency of agriculture and food system in the US. Chemosphere, 806-813. doi:10.1016/j.chemosphere.2011.01.051

Taylor, M. D. (2007). Accumulation of uranium in soils from impurities in phosphate fertilisers. Landbauforschnug Völkenrode 2, 133-139.

Tessier, A., Campbell, P. G., & Bisson, M. (1979). Sequential Extraction Procedure for the Speciation of Particulate Trace Metals. Analytical Chemistry, 844-851.

Tyler, G., & Olsson, T. (2001). lant uptake of major and minor mineral elements as influenced by soil acidity and liming. Plant and Soil, 307-321.

U.S. Geological Survey. (2004). National Uranium Resource Evaluation (NURE) Hydrogeochemical and Stream Sediment Reconnaissance data. Denver, CO: U.S. Geological Survey.

Ulrich, A. E., Schnug, E., Prasser, H.-M., & Frossard, E. (2014). Uranium endowments in phosphate rock. Science of the Total Environment, 226-234.

UNSCEAR. (1993). United Nations Scientific Committee on the Effects of Atomic Radiation. New York: United Nations.

63

USGS Gap Analysis Program. (2016). GAP/LANDFIRE National Terrestrial Ecosystems 2011. Boise: U.S. Geological Survey. Retrieved from https://www.sciencebase.gov/catalog/file/get/573cc51be4b0dae0d5e4b0c5?f=__disk__5d /11/f4/5d11f4366a3402f7e0d23ffa77258a4e12f04809&transform=1&allowOpen=true

Van Kauwenbergh, S. J. (2010). World Phosphate Rock Reserves and Resources. Muscle Shoals: IFDC.

Vandenhove, H., Van Hees, M., Wouters, K., & Wannijn, J. (2007). Can we predict uranium bioavailability ased on soil parameters? Part 1: Effect of soil parameters on soil solution uranium concentration. Environmental Pollution, 587-595.

Vitosh, M. L., Johnson, J. W., & Mengel, D. B. (1995). Tri-State Fertilizer recommendations for corn, soybeans, wheat and alfalfa Ext. Bull. E-2567. East Lansing: Michigan State University.

Volkovich, V. A., Polovov, I. B., Vasin, B. D., Griffiths, T. R., Sharrad, C. A., May, I., & Charnock, J. M. (2007). Effect of Melt Composition on the Reaction of Uranium Dioxide with Hydrogen Chloride in Molten Alkali Chlorides. Zeitschrift für Naturforschung, 671- 676.

Ward, A., D'Ambrosio, J., & Witter, J. (2019, September 5th). Extension/Outreach - Waterman Farm. Retrieved from FABE - Department of Food, Agricultural and Biological Engineering: https://fabe.osu.edu/extension-outreach/waterman-farm

WHO. (2012). Uranium in Drinking-water. Geneva: World Health Organization.

Yamaguchi, N., Kawasaki, A., & Iiyama, I. (2009). Distribution of uranium in soil components of agricultural fields after long-term application of phosphate fertilizers. Science of the Total Environment, 1383-1390.

Yamazaki, I. M., & Geraldo, L. P. (2003). Uranium content in phosphate fertilizers commercially produced in Brazil. Appl. Rad. Isotopes, 133-136.

Zamora, M. L., Tracy, B. L., Zielinski, J. M., Meyerhof, D. P., & Moss, M. A. (1998). Chronic Ingestion of Uranium in Drinking Water: A Study of Kidney Bioaffects in Humans. Toxicological Sciences, 68-77.

Zewainy, R. M. (2009). Evaluation of Soil Extraction Methods for Uranium. Quedlinburg, Germany: Julius Kühn Institute Federal Research Center for Cultivated Plants.

Zhou, P., & Gu, B. (2005). Extraction of Oxidized and Reduced Forms of Uranium from Contaminated Soils: Effects of Carbonate Concentration and pH. Environmental Science and Technology, 4435-4440.

64

Zielinski, R. A., & Meier, A. L. (1988). The association of uranium with organic matter in Holocene peat: an experimental leaching study. Applied Geochemistry, 631-643.

Zielinski, R. A., Orem, W. H., Simmons, K. R., & Bohlen, P. J. (2006). Fertilizer-Derived Uranium and Sulfur in Rangeland Soil and Runoff: A Case Study in Central Florida. Water, Air, and Soil Pollution, 163-183.

65

Appendix A: General Tables and Figures

Figure 1. Field at Waterman Farm showing plots for Fallow, Compost, Manure, and Cover Crop management practices.

66

Figure 2 Field at Waterman Farm showing plots with different rates of Fertilizer and Mulch application.

67

Figure 3 East Badger Field in Wayne County, OH (Fulford & Culman, 2018)

68

Figure 4 Western Agricultural Research Station Field in Clark County, OH (Fulford & Culman, 2018)

69

Figure 5 Northwest Agricultural Research Station in Wood County, OH (Fulford & Culman, 2018)

70

Figure 6 U concentrations from Coshocton samples.

71

Figure 7 Pie charts showing the distribution of total U in samples from Waterman Farm.

72

Figure 8 Pie charts showing the distribution of total U in samples from the East Badger location in Wayne County, OH.

73

Figure 9 Pie charts showing the distribution of total U in samples from the Northwest location in Wood County, OH.

74

Figure 10 Pie charts showing the distribution of total U in samples from the Western location in Clark County, OH.

75

Figure 11 Box plot showing U concentrations at Waterman Farm from samples that were leached using only 10% HCl.

76

Figure 12 Box plot showing U concentrations at Waterman Farm for samples that were leached with DI Water as part of the sequential extraction process.

77

Figure 13 Box plot showing U concentrations at Waterman Farm for samples that were leached with 0.1 M sodium bicarbonate solution as part of the sequential extraction process.

78

Figure 14 Box plot showing U concentrations at Waterman Farm for samples that were leached with 10% HCl as part of the sequential extraction process.

79

Figure 15 U concentrations from non-sequential HCl extractions for “Wooster” soils.

80

Figure 16 U concentrations from DI Water extractions for “Wooster” soils.

81

Figure 17 U concentrations from 0.1M sodium bicarbonate extractions for “Wooster” soils

82

Figure 18 U concentrations from the sequential HCl extractions for “Wooster” soils

83

Figure 19 U concentrations at topsoil depth (0-5cm) throughout the contiguous US (Smith, et al., 2014).

84

Figure 20 Location of sample sites used in this investigation over their respective counties in Ohio.

85

Figure 21 Locations of sampled sites over the glaciated and unglaciated portions of Ohio.

86

Figure 22 Locations of sampled sites over Ohio’s bedrock geology.

87

Figure 23 Locations of sampled sites over Ohio’s general land use/cover.

88

Figure 24 Locations of sampled sites over Ohio’s geological units

89

Figure 25 Map depicting the location and concentration range of samples taken from the NURE dataset.

90

Figure 26 “Heat map” showing the ranges of U concentrations throughout the extent of the NURE dataset using Nearest Neighbor interpolation.

91

Figure 27 Map showing the amount of NURE data that is contained within both the glaciated and unglaciated areas of Ohio.

92

Figure 28 Map showing the NURE dataset over Ohio’s bedrock geology.

93

Figure 29 Map showing the NURE dataset over Ohio’s general land use/cover.

94

Figure 30 Map showing the NURE dataset over Ohio’s geologic units.

95

Figure 31 Map showing the extent of the USGS topsoil dataset, including the ranges in U concentrations, over Ohio.

96

Figure 32 “Heat map” of USGS topsoil U concentrations over Ohio using Nearest Neighbor interpolation.

97

Figure 33 Map showing USGS topsoil U concentrations over Ohio’s glaciated and unglaciated areas.

98

Figure 34 Map showing USGS topsoil U concentrations over Ohio’s bedrock geology.

99

Figure 35 Map showing USGS topsoil U concentrations over Ohio’s general land use/cover.

100

Figure 36 Map showing USGS topsoil U concentrations over Ohio’s geologic units.

101

Figure 37 Monazite from sample EB-311 (East Badger Site)

102

Figure 38 Zircon from sample WE-202 (Western Site).

103

Figure 39 Cerianite identified through spot chemical analysis from sample WE-202 (Western Site).

104

Figure 40 Zircon from sample NW-303 (Northwest Site).

105

Figure 41 Zircon from sample FA 1-2 0-5cm (Fallow plot at Waterman Farm).

106

Figure 42 Monazite from sample FERT 5-5 0-5cm (Fertilized plot at Waterman Farm).

107

Figure 43 Monazite from sample FA 1-2 0-5cm (Fallow plot at Waterman Farm).

108

Figure 44 Trident Internal Standard Kit “T-Piece”

109

Figure 45. Graph comparing LOI and dichromate organic carbon values.

110

Analyte Weight (g) LOI (%) FA 1-3 0-5CM 7.1 5.64 FERT 5-4 0-5CM 7.4 5.14 CC 2-2 0-5CM 7 10 M 1-1 0-5CM 9 11 EB-311 7.5 5.82 EB-313 7.1 5.7 EB-315 7.5 5.51 NW-303 7.4 9.19 NW-304 9.1 9.07 NW-305 8.9 8.98 WE-202 8.7 6.41 WE-207 7.9 6.05 WE-209 9.9 6.3 Table A- 1 Loss on Ignition (LOI) data provided by SGS Canada

111

Analyte Weight (g) SiO2 (%) Al2O3 (%) Fe2O3 (%) MgO (%) FA 1-3 0-5CM 7.1 73.6 9.94 4.75 0.75 FERT 5-4 0-5CM 7.4 74.9 9.36 4.71 0.64 CC 2-2 0-5CM 7 66.8 10.9 5.31 0.92 M 1-1 0-5CM 9 66.5 10.4 5.29 0.84 EB-311 7.5 76 8.86 4.31 0.59 EB-313 7.1 75.4 9.1 4.97 0.61 EB-315 7.5 76.7 8.6 4.03 0.54 NW-303 7.4 64.5 12.9 5.47 1.29 NW-304 9.1 64.4 12.9 5.19 1.31 NW-305 8.9 64.3 12.8 5.41 1.28 WE-202 8.7 74.2 9.41 3.58 0.82 WE-207 7.9 74.9 9.23 3.61 0.7 WE-209 9.9 75.3 9.15 3.34 0.7 Table A- 2 Si, Al, Fe, and Mg Oxide data provided by SGS Canada

112

Analyte Weight (g) CaO (%) K2O (%) Na2O (%) TiO2 (%) FA 1-3 0-5CM 7.1 0.53 2.16 0.83 0.77 FERT 5-4 0-5CM 7.4 0.47 2.05 0.85 0.78 CC 2-2 0-5CM 7 0.86 2.21 0.82 0.7 M 1-1 0-5CM 9 0.82 2.1 0.81 0.73 EB-311 7.5 0.4 1.88 0.75 0.82 EB-313 7.1 0.36 1.92 0.77 0.83 EB-315 7.5 0.38 1.84 0.79 0.83 NW-303 7.4 1.08 2.8 0.99 0.72 NW-304 9.1 1.07 2.79 1 0.72 NW-305 8.9 1.09 2.81 1.04 0.71 WE-202 8.7 0.87 2 1 0.74 WE-207 7.9 0.75 2.02 1.05 0.71 WE-209 9.9 0.74 2.01 1.02 0.72 Table A- 3 Ca, K, Na, and Ti Oxide data provided by SGS Canada

113

Analyte Weight (g) MnO (%) P2O5 (%) Cr2O3 (%) V2O5 (%) FA 1-3 0-5CM 7.1 0.11 0.13 0.01 0.02 FERT 5-4 0-5CM 7.4 0.1 0.16 <0.01 <0.01 CC 2-2 0-5CM 7 0.19 0.17 0.01 0.02 M 1-1 0-5CM 9 0.19 0.24 0.01 0.02 EB-311 7.5 0.13 0.12 0.01 0.01 EB-313 7.1 0.13 0.13 0.01 <0.01 EB-315 7.5 0.17 0.15 0.01 0.01 NW-303 7.4 0.06 0.18 0.02 0.02 NW-304 9.1 0.08 0.2 0.01 0.02 NW-305 8.9 0.07 0.19 0.01 0.02 WE-202 8.7 0.05 0.09 <0.01 <0.01 WE-207 7.9 0.06 0.1 0.01 <0.01 WE-209 9.9 0.05 0.11 0.01 0.02 Table A- 4 Mn, P, Cr, and V Oxide data provided by SGS Canada

114

Analyte Al2O3 (%) CaO (%) K2O (%) Na2O (%) CIA FA 1-3 0-5CM 9.94 0.53 2.16 0.83 73.8 FERT 5-4 0-5CM 9.36 0.47 2.05 0.85 73.5 CC 2-2 0-5CM 10.9 0.86 2.21 0.82 73.7 M 1-1 0-5CM 10.4 0.82 2.1 0.81 73.6 EB-311 8.86 0.4 1.88 0.75 74.5 EB-313 9.1 0.36 1.92 0.77 74.9 EB-315 8.6 0.38 1.84 0.79 74.1 NW-303 12.9 1.08 2.8 0.99 72.6 NW-304 12.9 1.07 2.79 1 72.6 NW-305 12.8 1.09 2.81 1.04 72.2 WE-202 9.41 0.87 2 1 70.9 WE-207 9.23 0.75 2.02 1.05 70.7 WE-209 9.15 0.74 2.01 1.02 70.8 Table A- 5 Calculated Chemical Index of Alteration (CIA) values for samples provided to SGS Canada

115

Standard Intermediate 1 2 3 4 5 U 1000 μg/L 1 μg/L 5 μg/L 10 μg/L 30 μg/L 60 μg/L Table A- 6 Concentrations of standards used for ICP-MS analysis of 10% HCl and 0.1M sodium bicarbonate leaches.

DI Water Intermediate 1A 2A 3A 4A Standard A U 10 μg/L 1 ppt 10 ppt 100 ppt 500ppt Table A- 7 Concentrations of standards used for ICP-MS analysis of DI water leaches.

116

HCl and Bicarbonate Check Standards Name [U] μg/L Standard 3 10 Check Standard 3-A 9.179 Check Standard 3-B 10.145 Check Standard 3-C 9.356 Check Standard 3-D 8.962 Check Standard 3-E 10.269 Check Standard 3-F 10.124 Check Standard 3-G 10.361 Check Standard 3-H 9.313 Check Standard 3-I 10.389 Check Standard 3-J 8.967 Check Standard 3-K 9.146 Check Standard 3-L 10.451 Check Standard 3-M 10.506 Check Standard 3-N 10.5 Check Standard 3-A1 9.96 Check Standard 3-B1 10.43 Check Standard 3-C1 9.783 Check Standard 3-D1 9.983 Mean 9.879 StDev 0.606 RSD (%) 6.137 Table A- 8 Check Standard concentrations and RSD calculations for HCl and Bicarbonate extraction ICP- MS measurement precision.

117

DI Water Check Standards Name [U] μg/L Standard 3A 0.1 Check Std 3A-1 0.108 Check Std 3A-2 0.104 Check Std 3A-3 0.102 Check Std 3A- 1A 0.108 Check Std 3A- 2A 0.108 Mean 0.11 StDev 0.00 RSD (%) 2.39 Table A- 9 Check Standard concentrations and RSD calculations for DI water extraction ICP-MS measurement precision.

118

Organic Matter Determination Method Comparison (% OM) Sample XRF - LOI Dichromate In-house LOI FA 1-3 0-5CM 5.64 0.85 3.40 FERT 5-4 0-5CM 5.14 0.97 3.66 CC 2-2 0-5CM 10 2.16 6.94 M 1-1 0-5CM 11 2.76 7.76 EB-311 5.82 0.88 4.17 EB-313 5.7 0.81 3.97 EB-315 5.51 0.83 3.43 NW-303 9.19 1.73 6.28 NW-304 9.07 1.71 6.58 NW-305 8.98 1.72 6.05 WE-202 6.41 1.00 4.08 WE-207 6.05 1.11 4.12 WE-209 6.3 1.15 4.00 Table A- 10 Comparison of values from three different instances of organic matter determination.

119

Samples Analyzed using SEM Sample Minerals Identified EB-311 Monazite, Zircon EB-315 Zircon, Biotite, Monazite WE-202 Zircon, Cerianite NW-303 Monazite, Zircon FA 1- 0-5CM Monazite, Zircon FERT 5-5 0-5CM Monazite, Zircon Table A- 11 List of minerals identified through SEM analyses.

Variable [U] HCl Only pH LOI LOI -0.14 0.23 1 pH -0.43** 1 [U] HCl Only 1 *p < 0.05, **p < 0.01, ***p < 0.001 Table A- 12 Correlations between U concentrations for non-sequentially leached samples from Waterman Farm with LOI and pH.

[U] Seq DI [U] Seq Variable [U] Seq Acid pH LOI Water Bicarb LOI 0.56** 0.57** 0.68*** 0.41 1 pH 0.24 0.22 -8.40E-03 1 [U] Seq Acid 0.43 0.27 1 [U] Seq Bicarb 0.41 1 [U] Seq DI Water 1 *p < 0.05, **p < 0.01, ***p < 0.001 Table A- 13 Correlations between U concentrations for sequentially leached samples from Waterman Farm with LOI and pH.

120

[U] Seq [U] Seq [U] Seq Total U Variable CIA pH LOI DI Water Bicarb Acid (XRF) LOI 0.59* 0.67* 0.78** 0.76** 0.0061 -0.39 1 pH -0.56* -0.45 -0.77** -0.61* -0.044 1 CIA -0.17 -0.25 0.11 0.19 1 Total U (XRF) 0.75** 0.80** 0.93*** 1 [U] Seq Acid 0.81 0.75 1 [U] Seq Bicarb 0.92 1 [U] Seq DI Water 1 *p < 0.05, **p < 0.01, ***p < 0.001 Table A- 14 Correlations between U concentrations from sequential extractions to total U (obtained from XRF) and calculated CIA values.

121

Grain Size Mineral Size Sample Name Mineral Estimate (%) (pixels) (pixels) Monazite 277576 2597 0.94 Monazite 72917 1453 1.99 EB-311 Monazite 382087 758 0.20 Zircon 37442 2062 5.51 Zircon 3736 2360 63.17 EB-315 Biotite 4484 149 3.32 Zircon 66801 1139 1.71 Zircon 5982 481 8.04 WE-202 Cerianite/Zircon 24370 488 2.00 Monazite/Zircon 106988 629 0.59 Zircon 5260 1997 37.97 NW-303 Zircon 17502 11803 67.44 Zircon 580478 455 0.08 Zircon 14802 8357 56.46 ZIrcon 7965 223 2.80 FA 1-2 0-5CM Monazite 113363 2878 2.54 Zircon 6077 5256 86.49 Zircon 26107 14866 56.94 Monazite 2659 461 17.34 FERT 5-5 0- Zircon 76462 10073 13.17 5CM Monazite 2432 1757 72.25 Monazite 118051 2367 2.01 Table A- 15 Approximations of mineral to grain sizes from SEM images.

122

Bedrock Dataset Associated Formation(s) Type Dolostone Salina Group; Clinton and Cataract Groups Sample Site Limestone Maxville Limestone; Columbus Limestone Locations Shale Allegheny and Pottsville Groups Black Ohio Shale Shale Mudstone Waynesville and Arnheim Formations Sandstone Berea Sandstone and Bedford Shale NURE Monongahela Group; Allegheny and Pottsville Shale Groups; Maxville Limestone, Rushville, Logan, and Cuyahoga Formations Siltstone Conemaugh Group Black Ohio Shale Shale Traverse Group; Lockport Dolomite; Tymochtee and Dolostone Greenfield Formations; Salina Group Grant Lake and Fairview Formation; Columbus Limestone Limestone; Dundee Limestone Mudstone Waynesville and Arnheim Formations Sandstone Berea Sandstone and Bedford Shale USGS Allegheny and Pottsville Groups; Maxville Limestone; Rushville, Logan, and Cuyahoga Formations; Monongahela Group; ; Shale Olentangy Shale; Ordovician Undifferentiated; Preacherville Member of the , Waynesville and Arnheim Formations; Cincinnati Group; Antrim Shale Siltstone Conemaugh Group Barnes, et al., Cincinnati Shale Shale 2020 Table A- 16 Ohio geologic units associated with the NURE and USGS datasets.

123

Land Use Type Classification Orchards Vineyards and Other High Structure Agriculture Agricultural Cultivated Cropland Pasture/Hay Developed, High Intensity Developed, Low Intensity Developed, Medium Intensity Developed, Open Space Quarries, Mines, Gravel Pits and Oil Wells Disturbed, Non-specific Disturbed/Successional - Grass/Forb Developed Regeneration Disturbed/Successional - Shrub Regeneration Harvested Forest - Grass/Forb Regeneration Harvested Forest-Shrub Regeneration Recently burned grassland Water Open Water (Fresh) Table A- 17 The grouping of original land use classifications into their simplified categories for the agricultural, developed and water land use categories.

124

Undeveloped Land - Classifications Evergreen Plantation or Managed Pine North-Central Interior Dry Oak Forest and Woodland North-Central Interior Dry-Mesic Oak North-Central Interior Oak Savanna Forest and Woodland North-Central Oak Barrens Managed Tree Plantation Ruderal forest Laurentian-Acadian Northern Hardwoods Forest Northern Atlantic Coastal Plain Dry Southern Interior Low Plateau Dry-Mesic Oak Hardwood Forest Forest Laurentian Pine-Oak Barrens Allegheny-Cumberland Dry Oak Forest and Woodland - Hardwood Allegheny-Cumberland Dry Oak Forest Central Appalachian Pine-Oak Rocky Woodland and Woodland - Pine Modifier Northeastern Interior Dry-Mesic Oak Southern and Central Appalachian Oak Forest Forest Southern Ridge and Valley Dry Southern Ridge and Valley Dry Calcareous Calcareous Forest Forest - Pine modifier North-Central Interior Beech-Maple North-Central Interior Maple-Basswood Forest Forest Appalachian Hemlock-Hardwood Forest South-Central Interior Mesophytic Forest Central Interior and Appalachian Central Interior and Appalachian Riparian Floodplain Systems Systems North-Central Interior and Appalachian South-Central Interior Large Floodplain - Forest Rich Swamp Modifier South-Central Interior Small Stream and North-Central Interior Wet Flatwoods Riparian Laurentian-Acadian Floodplain Systems Laurentian-Acadian Swamp Systems Central Tallgrass Prairie North-Central Interior Sand and Gravel Tallgrass Prairie Central Interior Highlands Calcareous Central Interior and Appalachian Shrub- Glade and Barrens Herbaceous Wetland Systems Great Lakes Coastal Marsh Systems Great Lakes Wet-Mesic Lakeplain Prairie Great Plains Prairie Pothole Central Interior Acidic Cliff and Talus Central Interior Calcareous Cliff and Southern Interior Acid Cliff Talus Introduced Upland Vegetation - Perennial Introduced Upland Vegetation - Treed Grassland and Forbland Unconsolidated Shore Undifferentiated Barren Land Table A- 18 The grouping of original land use classifications into their simplified categories for the undeveloped land use category.

125

Appendix B: pH values for samples

Waterman Farm, Manure 0-5cm 40-50cm Sample pH [H+] pH [H+] M 1-1 4.21 6.17E-05 4.84 1.45E-05 M 1-2 5.34 4.57E-06 5.28 5.25E-06 M 1-3 6.14 7.24E-07 5.36 4.37E-06 M 2-1 5.43 3.72E-06 5.74 1.82E-06 M 2-2 6.25 5.62E-07 6.18 6.61E-07 M 2-3 5.68 2.09E-06 5.66 2.19E-06 Mean 5.51 1.22E-05 5.51 4.79E-06 Table B- 1 pH values for Manure plots at Waterman Farm.

Waterman Farm, Cover Crop 0-5cm 40-50cm Sample pH [H+] pH [H+] CC 1-1 5.37 4.27E-06 5.58 2.63E-06 CC 1-2 5.13 7.41E-06 5.39 4.07E-06 CC 1-3 4.49 3.24E-05 5.07 8.51E-06 CC 2-1 5.4 3.98E-06 5.31 4.9E-06 CC 2-2 5.72 1.91E-06 6.76 1.74E-07 CC 2-3 6.12 7.59E-07 6.77 1.7E-07 Mean 5.37 8.45E-06 5.81 3.41E-06 Table B- 2 pH values for Cover Crop plots at Waterman Farm.

126

Waterman Farm, Fertilizer 0-5cm 40-50cm Sample pH [H+] pH [H+] FERT 1-1 4.9 1.26E-05 5.07 8.51E-06 FERT 1-2 5.17 6.76E-06 4.86 1.38E-05 FERT 1-3 5.1 7.94E-06 4.98 1.05E-05 FERT 2-1 5.02 9.55E-06 4.52 3.02E-05 FERT 2-2 4.49 3.24E-05 5.55 2.82E-06 FERT 2-3 4.93 1.17E-05 4.67 2.14E-05 FERT 5-1 6.03 9.33E-07 5.03 9.33E-06 FERT 5-2 5.01 9.77E-06 5.42 3.8E-06 FERT 5-3 5.06 8.71E-06 5.53 2.95E-06 FERT 5-4 4.88 1.32E-05 4.82 1.51E-05 FERT 5-5 6.56 2.75E-07 5.1 7.94E-06 Mean 5.20 1.03E-05 5.05 1.15E-05 Table B- 3 pH values for Fertilized plots at Waterman Farm.

Waterman Farm, Fallow 0-5cm 40-50cm Sample pH [H+] pH [H+] FA 1-1 3.74 1.82E-04 4.99 1.02E-05 FA 1-2 4.55 2.82E-05 4.8 1.58E-05 FA 1-3 4.83 1.48E-05 4.92 1.2E-05 FA 2-1 4.73 1.86E-05 4.7 2E-05 FA 2-2 4.71 1.95E-05 4.99 1.02E-05 FA 2-3 4.48 3.31E-05 4.85 1.41E-05 FA 5-1 4.23 5.89E-05 4.36 4.37E-05 FA 5-2 4.65 2.24E-05 4.62 2.4E-05 FA 5-3 4.64 2.29E-05 4.92 1.2E-05 FA 5-4 5.51 3.09E-06 4.83 1.48E-05 FA 5-5 4.32 4.79E-05 4.07 8.51E-05 Mean 4.58 4.1E-05 4.73 2.38E-05 Table B- 4 pH values for Fallow plots at Waterman Farm.

127

Coshocton Samples Sample Management Practice pH [H+] Coshocton NTNM- No Till No Manure 0- 6.36 4.37E-07 01 2.5cm) Till with Manure (0- Coshocton T-01 6.93 1.17E-07 2.5cm) Grassy Pasture (0- Coshocton P-01 6.37 4.27E-07 2.5cm) Mean 6.55 3.27E-07 Table B- 5 pH values for the Coshocton site.

Wooster, East Badger Fertilizer Application Rate 0x 1x 2x Sample pH [H+] pH [H+] pH [H+] EB-111 6.48 3.31E-07 EB-207 6.34 4.57E-07 EB-311 5.86 1.38E-06 EB-404 5.48 3.31E-06 EB-115 6.08 8.32E-07 EB-202 6.21 6.17E-07 EB-313 5.45 3.55E-06 EB-409 4.77 1.70E-05 EB-114 6 1.00E-06 EB-209 6.11 7.76E-07 EB-315 5.44 3.63E-06 EB-405 5.81 1.55E-06 Mean 6.04 1.37E-06 5.63 5.49E-06 5.84 1.74E-06 Table B- 6 pH values for the East Badger site.

128

Wooster, Western Fertilizer Application Rate 0x 1x 2x Sample pH [H+] pH [H+] pH [H+] WE-111 6.18 6.61E-07 WE-207 5.25 5.62E-06 WE-311 5.66 2.19E-06 WE-404 5.64 2.29E-06 WE-115 6.13 7.41E-07 WE-202 5.83 1.48E-06 WE-313 5.58 2.63E-06 WE-409 6.08 8.32E-07 WE-114 5.56 2.75E-06 WE-209 5.25 5.62E-06 WE-315 5.32 4.79E-06 WE-405 5.77 1.7E-06 Mean 5.68 2.69E-06 5.91 1.42E-06 5.48 3.72E-06 Table B- 7 pH values for the Western site.

Wooster, Northwest Fertilizer Application Rate 0x 1x 2x Sample pH [H+] pH [H+] pH [H+] NW-101 4.74 1.82E-05 NW-212 4.63 2.34E-05 NW-303 4.7 2E-05 NW-417 5.12 7.59E-06 NW-107 4.77 1.7E-05 NW-214 4.93 1.17E-05 NW-305 5.19 6.46E-06 NW-414 4.88 1.32E-05 NW-105 4.74 1.82E-05 NW-216 4.77 1.7E-05 NW-304 4.82 1.51E-05 NW-416 5.17 6.76E-06 Mean 4.80 1.73E-05 4.94 1.21E-05 4.88 1.43E-05 Table B- 8 pH values for the Northwest site.

129

Appendix C: Loss On Ignition Values

Cover Crop % LOI Sample Name 0-5cm 40-50cm CC 1-1 7.35 8.14 CC 1-2 7.57 6.75 CC 1-3 7.37 9.12 CC 2-1 6.99 5.61 CC 2-2 6.94 6.68 CC 2-3 5.11 4.63 Mean 6.89 6.82 Table C- 1 LOI values for Cover Crop plots at Waterman Farm.

Manure % LOI Sample Name 0-5cm 40-50cm M 1-1 7.76 9.35 M 1-2 9.67 8.91 M 1-3 11.0 7.82 M 2-1 6.27 4.80 M 2-2 8.22 5.95 M 2-3 6.14 4.42 Mean 8.17 6.87 Table C- 2 LOI values for Manure plots at Waterman Farm.

130

Fallow % LOI Sample Name 0-5cm 40-50cm FA 1-1 5.28 4.19 FA 1-2 3.34 3.36 FA 1-3 3.40 3.26 FA 2-1 3.79 3.54 FA 2-2 3.28 3.62 FA 2-3 2.78 3.42 FA 5-1 3.61 3.58 FA 5-2 3.22 3.74 FA 5-3 3.54 3.67 FA 5-4 3.48 4.09 FA 5-5 4.92 3.80 Mean 3.69 3.66 Table C- 3 LOI values for Fallow plots at Waterman Farm.

Fertilizer % LOI Sample Name 0-5cm 40-50cm FERT 1-1 3.22 3.71 FERT 1-2 3.48 3.80 FERT 1-3 3.24 4.21 FERT 2-1 4.00 3.92 FERT 2-2 3.97 3.62 FERT 2-3 3.76 3.87 FERT 5-1 3.22 3.80 FERT 5-2 3.39 4.15 Fert 5-3 3.30 3.18 FERT 5-4 3.66 3.56 FERT 5-5 4.63 3.86 Mean 3.63 3.79 Table C- 4 LOI values for Fertilizer plots at Waterman Farm.

131

Northwest % LOI 0x Fertilizer 1x Fertilizer 2x Fertilizer Sample Name Application Application Application NW 101 5.27 NW-212 5.98 NW-303 6.28 NW-417 5.33 NW 107 5.07 NW-214 5.78 NW-305 6.05 NW-414 6.60 NW 105 5.26 NW-216 5.55 NW-304 6.58 NW-416 6.40 Mean 5.71 5.87 5.95 Table C- 5 LOI values for the Northwest site.

132

East Badger % LOI 0x Fertilizer 1x Fertilizer 2x Fertilizer Sample Name Application Application Application EB-111 4.18 EB 207 3.81 EB-311 4.17 EB-404 4.08 EB-115 4.18 EB 202 4.00 EB-313 3.97 EB-409 3.85 EB-114 4.56 EB 209 4.04 EB-315 3.43 EB-405 3.93 Mean 4.06 4.00 3.99 Table C- 6 LOI values for the East Badger site.

Western % LOI 0x Fertilizer 1x Fertilizer 2x Fertilizer Sample Name Application Application Application WE-111 3.81 WE-207 4.12 WE-311 3.19 WE 404 4.08 WE-115 3.79 WE-202 4.08 WE-313 3.55 WE-409 3.91 WE-114 3.64 WE-209 4.00 WE-315 3.35 WE-405 2.85 Mean 3.80 3.83 3.46 Table C- 7 LOI values for the Western site.

133

Coshocton Sample Name Management Practice % LOI COSHOCTON NTNM-1 No Till, No Manure 7.07 COSHOCTON P-01 Pasture 6.21 CHOSHOCTON T-01 Till, No Manure 3.97 Mean 5.75 Table C- 8 LOI values for the Coshocton site.

134

Appendix D: U Concentrations

Averages Analyte Location U (μg/g) (μg/g) FA 1-3 0-5CM 4.49 FERT 5-4 0- Waterman 5CM 4.16 4.62 Farm CC 2-2 0-5CM 4.83 M 1-1 0-5CM 4.99 EB-311 3.06 EB-313 East Badger 3.05 3.07 EB-315 3.11 NW-303 5.28 NW-304 Northwest 5.42 5.29 NW-305 5.17 WE-202 2.87 WE-207 Western 2.7 2.78 WE-209 2.78 Table D- 1 Total U data provided by SGS Canada

135

Waterman Farm, Cover Crop Plots, 10% HCl [U], μg/g Extractions Cover Crop Plot Replicate 0-5cm 40-50cm Sample CC 1-1 1 1 0.38 0.44 CC 1-3 1 3 0.57 0.62 CC 2-1 2 1 0.52 0.58 CC 2-3 2 3 0.43 0.42 Mean 0.48 0.51 Std. Dev. 0.07 0.09 Coeff. Var. 16 % 17 % P value 0.85 Table D- 2 U concentrations and statistics for 10% HCl extractions for Cover Crop plots, Waterman Farm.

136

Waterman Farm, Manure Plots, 10% HCl Extractions [U], μg/g Manure Sample Location Replicate 0-5cm 40-50cm M 1-1 1 1 0.52 0.56 M 1-3 1 3 0.44 0.59 M 2-1 2 1 0.52 0.31 M 2-3 2 3 0.52 0.36 Mean 0.50 0.46 Std. Dev. 0.03 0.12 Coeff. Var. 0.07 0.27 P value 0.44 Table D- 3 U concentrations and statistics for 10% HCl extractions for Manure plots, Waterman Farm.

Waterman Farm, Fallow Plots, 10% HCl Extractions [U], μg/g Fallow Sample Location Replicate 0-5cm 40-50cm FA 1-1 1 1 0.74 0.50 FA 1-3 1 3 0.52 0.57 FA 2-1 2 1 0.49 0.58 FA 2-3 2 3 0.57 0.51 FA 5-1 5 1 0.51 0.57 FA 5-2 5 2 0.51 0.57 FA 5-4 5 4 0.51 0.61 FA 5-5 5 5 0.51 0.52 Mean 0.54 0.55 Std. Dev. 0.08 0.04 Coeff. Var. 0.14 0.07 P Value 0.89 Table D- 4 U concentrations and statistics for 10% HCl extractions for Fallow plots, Waterman Farm.

137

Waterman Farm, Fertilizer Plots, 10% HCl Extractions [U], μg/g Fertilizer Location Replicate 0-5cm 40-50cm Samples FERT 1-1 1 1 0.45 0.53 FERT 1-3 1 3 0.48 0.54 FERT 2-1 2 1 0.68 0.65 FERT 2-3 2 3 0.68 0.68 FERT 5-1 5 1 0.57 0.62 FERT 5-2 5 2 0.56 0.53 FERT 5-4 5 4 0.52 0.53 FERT 5-5 5 5 0.61 0.54 Mean 0.57 0.58 Std. Dev. 0.08 0.06 Coeff. Var. 0.14 0.10 P Value 0.78 Table D- 5 U concentrations and statistics for 10% HCl extractions for Fertilizer plots, Waterman Farm.

East Badger, 10% [U], μg/g HCl Extractions EB Samples 0x Application 1x Application 2x Application EB-207 0.38 EB-311 0.44 EB-404 0.33 EB-202 0.45 EB-313 0.33 EB-409 0.36 EB-209 0.44 EB-315 0.40 EB-405 0.47 Mean 0.36 0.38 0.44 Std. Dev. 0.020 0.052 0.028 Coeff. Var. 0.056 0.14 0.063 Table D- 6 U concentrations and statistics for 10% HCl extractions for the East Badger site.

138

Western, 10% HCl [U], μg/g Extractions WE Samples 0x Application 1x Application 2x Application WE-111 0.34 WE-311 0.36 WE-404 0.38 WE-115 0.37 WE-313 0.34 WE-409 0.36 WE-114 0.37 WE-315 0.32 WE-405 0.33 Mean 0.36 0.36 0.34 Std. Dev. 0.014 0.014 0.026 Coeff. Var. 0.04 0.04 0.075 Table D- 7 U concentrations and statistics for 10% HCl extractions for the Western location.

Northwest, 10% HCl [U], μg/g Extractions NW Samples 0x Application 1x Application 2x Application NW-101 0.71 NW-212 0.69 NW-417 0.90 NW-107 0.75 NW-214 0.58 NW-414 0.74 NW-105 0.82 NW-216 0.65 NW-416 0.82 Mean 0.77 0.69 0.76 Std. Dev. 0.096 0.079 0.079 Coeff. Var. 0.13 0.11 0.10 Table D- 8 U concentrations and statistics for 10% HCl extractions for the Northwest location.

Coshocton, 10% HCl Extractions Sample Name Sample Description [U], μg/g COSHOCTON NTNM-1 No till, no manure 0.32 COSHOCTON P-01 Pasture (grassy) 0.31 COSHOCTON T-01 Till, with manure 0.28 Table D- 9 U concentrations and statistics for 10% HCl extractions for Coshocton.

139

Waterman Farm, Cover Crop plots, DI Water Extractions [U], μg/g Cover Crop Location Replicate 0-5cm 40-50cm Sample CC 1-2 1 2 0.00077 0.00050 CC 2-2 2 2 0.00035 0.00034 Mean 0.00056 0.00042 Table D- 10 U concentrations for DI Water extractions, Cover Crop plots, Waterman Farm.

Waterman Farm, Manure plots, DI Water Extractions [U], μg/g Manure Sample Location Replicate 0-5cm 40-50cm M 1-1 1 1 0.00064 ------M 1-2 1 2 0.0011 0.00043 M 2-2 2 2 0.00065 0.00066 Mean 0.00080 0.00055 Table D- 11 U concentrations for DI Water extractions, Manure plots, Waterman Farm.

Waterman Farm, Fallow plots, DI Water Extractions [U], μg/g Fallow Sample Location Replicate 0-5cm 40-50cm FA 1-2 1 2 0.00025 0.00040 FA 1-3 1 3 0.00030 ------FA 2-2 2 2 0.00021 0.00025 FA 5-3 5 3 0.00006 0.00013 Mean 0.00020 0.00026 Std. Dev. 8.95E-05 0.00011 Coeff. Var. 0.44 0.42 P value 0.59 Table D- 12 U concentrations and statistics for DI Water extractions, Fallow plots, Waterman Farm.

140

Waterman Farm, Fertilizer plots, DI Water Extractions [U], μg/g Fertilizer Location Replicate 0-5cm 40-50cm Samples FERT 1-2 1 2 0.00025 0.00036 FERT 2-2 2 2 0.0005 0.00039 FERT 5-3 5 3 0.00057 0.00079 FERT 5-4 5 4 0.00063 ------Mean 0.00049 0.00051 Std. Dev. 0.00015 0.00020 Coeff. Var. 0.30 0.38 P value 0.89 Table D- 13 U concentrations and statistics for DI Water extractions, Fertilizer plots, Waterman Farm.

East Badger, DI Water Extractions EB Samples Fertilizer Application Rate [U], μg/g EB-111 0x Application 0.00019 EB-115 1x Application 0.00027 EB-114 2x Application 0.00023 EB-311 0x Application 0.00025 EB-313 1x Application 0.00022 EB-315 2x Application 0.00034 Table D- 14 U concentrations for DI Water extractions, East Badger site.

Western, DI Water Extractions WE Samples Fertilizer Application Rate [U], μg/g WE-207 0x Application 0.00040 WE-202 1x Application 0.00028 WE-209 2x Application 0.00038 Table D- 15 U concentrations for DI Water extractions, Western site.

Northwest, DI Water Extractions NW Samples Fertilizer Application Rate [U], μg/g NW-303 0x Application 0.0014 NW-305 1x Application 0.00084 NW-304 2x Application 0.0013 Table D- 16 U concentrations for DI Water extractions, Northwest site.

141

Waterman Farm, Cover Crop plots, 0.1M Bicarbonate Extractions [U], μg/g Cover Crop Location Replicate 0-5cm 40-50cm Sample CC 1-2 1 2 0.31 0.22 CC 2-2 2 2 0.32 0.22 Mean 0.32 0.22 Table D- 17 U concentrations for 0.1 M sodium bicarbonate extractions, Cover Crop plots, Waterman Farm.

Waterman Farm, Manure plots, 0.1M Bicarbonate [U], μg/g Extractions Manure Sample Location Replicate 0-5cm 40-50cm M 1-1 1 1 0.23 M 1-2 1 2 0.34 0.29 M 2-2 2 2 0.23 0.27 Mean 0.27 0.28 Table D- 18 U concentrations for 0.1 M sodium bicarbonate extractions, Manure plots, Waterman Farm.

Waterman Farm, Fallow plots, 0.1M Bicarbonate Extractions [U], μg/g Fallow Sample Location Replicate 0-5cm 40-50cm FA 1-2 1 2 0.22 0.26 FA 1-3 1 3 0.21 FA 2-2 2 2 0.24 0.26 FA 5-3 5 3 0.19 0.2 Mean 0.22 0.24 Std. Dev. 0.018 0.028 Coeff. Var. 0.084 0.12 P value 0.35 Table D- 19 U concentrations and statistics for 0.1 M sodium bicarbonate extractions, Fallow plots, Waterman Farm.

142

Waterman Farm, Fertilizer plots, 0.1M Bicarbonate Extractions [U], μg/g Fertilizer Location Replicate 0-5cm 40-50cm Samples FERT 1-2 1 2 0.22 0.23 FERT 2-2 2 2 0.27 0.29 FERT 5-3 5 3 0.17 0.23 FERT 5-4 5 4 0.2 Mean 0.22 0.25 Std. Dev. 0.04 0.028 Coeff. Var. 0.17 0.11 P value 0.28 Table D- 20 U concentrations and statistics for 0.1 M sodium bicarbonate extractions, Fertilizer plots, Waterman Farm.

143

East Badger, 0.1M Bicarbonate Extractions EB Samples Fertilizer Application Rate [U], μg/g EB-111 0x Application 0.13 EB-115 1x Application 0.18 EB-114 2x Application 0.15 EB-311 0x Application 0.09 EB-313 1x Application 0.09 EB-315 2x Application 0.11 Table D- 21 U concentrations for 0.1 M sodium bicarbonate extractions, East Badger site.

Western, 0.1M Bicarbonate Extractions WE Samples Fertilizer Application Rate [U], μg/g WE-207 0x Application 0.20 WE-202 1x Application 0.14 WE-209 2x Application 0.23 Table D- 22 U concentrations for 0.1 M sodium bicarbonate extractions, Western site.

Northwest, 0.1M Bicarbonate Extractions NW Samples Fertilizer Application Rate [U], μg/g NW-303 0x Application 0.58 NW-305 1x Application 0.39 NW-304 2x Application 0.54 Table D- 23 U concentrations for 0.1 M sodium bicarbonate extractions, Northwest site.

Waterman Farm, Cover Crop Plots, Sequential HCl [U], μg/g Extractions Cover Crop Location Replicate 0-5cm 40-50cm Sample CC 1-2 1 2 0.48 0.45 CC 2-2 2 2 0.44 0.37 Mean 0.46 0.41 Table D- 24 U concentrations for 10% HCl sequential extractions, Cover Crop plots, Waterman Farm.

144

Waterman Farm, Manure Plots, Sequential HCl [U], μg/g Extractions Manure Sample Location Replicate 0-5cm 40-50cm M 1-1 1 1 0.76 M 1-2 0-5CM 1 2 0.49 0.53 M 2-2 0-5CM 2 2 0.52 0.41 Mean 0.59 0.47 Table D- 25 U concentrations for 10% HCl sequential extractions, Manure plots, Waterman Farm.

Waterman Farm, Fallow Plots, Sequential HCl [U], μg/g Extractions Fallow Sample Location Replicate 0-5cm 40-50cm FA 1-2 1 2 0.27 0.22 FA 1-3 1 3 0.45 ------FA 2-2 2 2 0.33 0.26 FA 5-3 5 3 0.28 0.22 Mean 0.33 0.23 Std. Dev. 0.07 0.02 Coeff. Var. 0.22 0.08 P Value 0.084 Table D- 26 U concentrations for 10% HCl sequential extractions, Fallow plots, Waterman Farm.

Waterman Farm, Fertilizer Plots, Sequential HCl [U], μg/g Extractions Fertilizer Location Replicate 0-5cm 40-50cm Sample FERT 1-2 1 2 0.29 0.48 FERT 2-2 2 2 0.38 0.39 FERT 5-3 5 3 0.28 0.2 FERT 5-4 5 4 0.47 ------Mean 0.36 0.36 Std. Dev. 0.08 0.12 Coeff. Var. 0.22 0.33 P Value 0.99 Table D- 27 U concentrations for 10% HCl sequential extractions, Fertilized plots, Waterman Farm.

145

East Badger, Sequential HCl Extractions EB Samples Fertilizer Application Rate [U], μg/g EB-111 0x Application 0.34 EB-115 1x Application 0.42 EB-114 2x Application 0.42 EB-311 0x Application 0.27 EB-313 1x Application 0.29 EB-315 2x Application 0.36 Table D- 28 U concentrations for 10% HCl sequential extractions, East Badger site.

Western, Sequential HCl Extractions WE Samples Fertilizer Application Rate [U], μg/g WE-207 0x Application 0.25 WE-202 1x Application 0.27 WE-209 2x Application 0.29 Table D- 29 U concentrations for 10% HCl sequential extractions, Western site.

Northwest, DI Water Extractions NW Samples Fertilizer Application Rate [U], μg/g NW-303 0x Application 0.67 NW-305 1x Application 0.71 NW-304 2x Application 0.65 Table D- 30 U concentrations for 10% HCl sequential extractions, Northwest site.

146

[U] (μg/g) Management Sample DI Water Bicarbonate Practice Leach Leach Acid Leach Total FA 1-3 0-5CM Fallow 3.00E-04 0.21 0.45 4.49 FERT 5-4 0-5CM Fertilizer 6.33E-04 0.20 0.47 4.16 CC 2-2 0-5CM Cover Crop 3.52E-04 0.32 0.44 4.83 M 1-1 0-5CM Manure 6.38E-04 0.23 0.76 4.99 EB-311 0x App Rate 2.51E-04 0.09 0.27 3.06 EB-313 1x App Rate 2.22E-04 0.09 0.29 3.05 EB-315 2x App Rate 3.43E-04 0.11 0.36 3.11 NW-303 0x App Rate 1.37E-03 0.58 0.67 5.28 NW-304 1x App Rate 1.25E-03 0.54 0.71 5.42 NW-305 2x App Rate 8.40E-04 0.39 0.65 5.17 WE-202 0x App Rate 2.75E-04 0.14 0.36 2.87 WE-207 1x App Rate 3.99E-04 0.20 0.34 2.7 WE-209 2x App Rate 3.85E-04 0.23 0.32 2.78 Mean 5.59E-04 0.26 0.47 3.99 Table D- 31 Comparison of U concentrations from sequential extractions to total U concentrations from XRF analyses.

Amount of total U extracted from leaching (%) Sample DI Water Leach Bicarb Leach Acid Leach FA 1-3 0-5CM 0.007 4.6 10.1 FERT 5-4 0-5CM 0.015 4.9 11.4 CC 2-2 0-5CM 0.007 6.7 9.0 M 1-1 0-5CM 0.013 4.7 15.3 EB-311 0.008 3.1 8.8 EB-313 0.007 3.1 9.7 EB-315 0.011 3.4 11.5 NW-303 0.026 10.9 12.7 NW-305 0.016 7.6 12.6 NW-304 0.023 9.9 13.2 WE-207 0.015 7.5 12.4 WE-202 0.010 5.0 12.5 WE-209 0.014 8.4 11.4 Mean 0.013 6.1 11.6 Table D- 32 Percentage of total U leached from each sample through sequential extractions.

147

Waterman Farm – Mean & Median Comparison for Box Plots Extraction Value (μg/g) Cover Crop Manure Fallow Fertilizer median 0.48 0.52 0.52 0.55 HCl Only mean 0.49 0.48 0.55 0.57 DI Water median 4.3E-04 6.5E-04 2.5E-04 5.0E-04 Sequential mean 4.9E-04 6.9E-04 2.3E-04 5.0E-04 Bicarbonate median 0.27 0.27 0.22 0.23 Sequential mean 0.27 0.27 0.23 0.23 HCl median 0.44 0.52 0.27 0.38 Sequential mean 0.43 0.54 0.29 0.39 Table D- 33 Mean and Median comparisons for Waterman Farm extraction values

Fraction of Mobile U in samples analyzed through XRF - Extractions Location, Extraction Min (μg/g U) Max (μg/g U) % of Total U Waterman, DI Water 3.0 x 10-4 6.38 x 10-4 0.007 – 0.015 East Badger, DI Water 2.22 x 10-4 3.43 x 10-4 0.007 – 0.011 Northwest, DI Water 8.4 x 10-4 1.25 x 10-4 0.016 – 0.026 Western, DI Water 2.75 x 10-4 3.99 x 10-4 0.01 – 0.015 Waterman, Bicarbonate 0.20 0.32 4.6 – 6.7 East Badger, 0.09 0.11 3.1 – 3.4 Bicarbonate Northwest, Bicarbonate 0.39 0.58 7.6 – 10.9 Western, Bicarbonate 0.14 0.23 5.0 – 8.4 Waterman, HCl 0.44 0.47 9.0 – 15.3 East Badger, HCl 0.27 0.36 8.8 – 11.5 Northwest, HCl 0.65 0.71 12.6 – 13.2 Western, HCl 0.32 0.36 11.4 – 12.5 Table D- 34 Range of each sequential extraction at Waterman and Wooster sites, compared to the amount of total U extracted.

Dataset Avg Stdev CV Min Max Count NURE 4.0 1.00 25% 1.4 19.0 1000 USGS Top 5cm 4.1 1.2 28% 1.7 9.0 69 Table D- 35 Comparison of totals for U concentration data for the NURE and USGS datasets in Ohio.

148

Unglaciated Glaciated Dataset Average StDev CV Min Max Count Average StDev CV Min Max Count NURE 4.1 1.0 24% 1.9 19.0 782 3.3 0.66 20% 1.4 6.7 218 USGS Top 5cm 3.6 0.6 17% 2.7 5.2 22 4.4 1.29 29% 1.7 9.0 47 Barnes, 2020, et al. 1.2 0.18 15% 1.0 1.7 20 Table D- 36 Comparison of U concentration data from the NURE and USGS datasets, as well as Barnes, 2020, et al. in the glaciated and unglaciated areas of Ohio.

USGS Top 5cm Land Cover Average % of Area (μg/g) StDev CV Min Max Count data Agricultural 4.5 1.3 29% 1.7 9 37 54% Developed 4.0 1.0 25% 2.5 5.7 7 10% Undeveloped 3.6 0.62 17% 2.5 4.9 25 36% Table D- 37 Comparison of data from visually identified points from the USGS Top 5cm data set in regard to their land cover.

149

NURE (μg/g) USGS Top 5cm (μg/g) This Work (μg/g) Lithology Average StDev CV Min Max Count Average StDev CV Min Max Count Average StDev CV Min Max Count Black Shale 2.4 0.0 0% 2.4 2.4 1 4.4 1.4 32% 1.7 5.8 5 Dolostone 4.9 1.1 22% 2.5 6.2 18 4.0 1.3 31% 2.7 5.4 6 Limestone 4.1 0.7 17% 3.4 5.0 3 4.6 0.3 7% 4.2 5.0 4 Mudstone 4.5 1.1 25% 2.6 9.6 107 3.3 0.6 18% 2.7 3.9 2 Sandstone 2.9 0.3 11% 2.5 3.2 3 3.5 0.0 0% 3.5 3.5 2 Shale 3.9 0.8 21% 1.4 8.2 613 3.8 1.2 30% 2.5 9.0 32 3.1 0.0 1% 3.1 3.1 3 Siltstone 4.4 1.2 28% 1.9 19.0 275 3.9 0.5 12% 2.8 4.4 7 Table D- 38 Comparison of U concentrations per geologic unit and dataset in Ohio.

Barnes, et al., 2020 (μg/g) Lithology Average StDev CV Min Max Count Shale 1.2 0.18 15% 1.0 1.7 20 Table D- 39 Southwestern Ohio glacial till U data (Barnes, et al., 2020). 1

150

2 Fertilizer Type and Mean U Concentration Source Origin (µg/g) Superphosphate Hamamo, et al. (1995); USA 104 McBride & Spears (2001) Godinez, et al. (1997); Mexico 90.5 Guzmán, et al. (1995) Da Conceição & Bonotto Brazil 65.2 (2006); Saueia & Mazilli (2006) Tanzania 325 Makweba & Holm (1993) Hungary (sourced from FAL sample collection, 2.0 Kola Peninsula, Russia) 2003-2007 Germany (source FAL sample collection, 91 unknown) 2003-2007 Triple Superphosphate Hamamo, et al. (1995); USA 178 McBride & Spears (2001); Robarge, et al. (2004) Godinez, et al. (1997); Mexico 197 Guzmán, et al. (1995) Da Conceição 7 Bonotto Brazil 50.2 (2006); Saueia & Mazilli (2006) Tanzania 362 Makweba & Holm (1993) Germany (source FAL sample collection, 106 unknown) 2003-2007 Soft/Ground Rock Phosphate USA 42.9 Hamamo, et al. (1995) Heiland (1986); Sam, et al. North Africa 205 (1999) Germany (source FAL sample collection, 64.8 unknown) 2003-2007 Table D- 40 U concentrations in different types of P-mineral fertilizers from their respective countries, adapted from Kratz, et al. (2008). Fertilizers are sourced from their own country unless specified otherwise.

151

Fertilizer Type and Mean U Concentration Source Origin (µg/g) NP Fertilizers USA 163 Robarge, et al. (2004) Da Conceição & Bonotto Brazil 71.7 (2006) Morocco 133 El Ghawi, et al. (1999) Romania (source Pantelica, et al. (1997) 59.4 unknown) Hungary (sourced from FAL sample collection, 2.9 Kola Peninsula, Russia) 2003-2007 Germany (source FAL sample collection, 27 unknown) 2003-2007 PK Fertilizers USA 89.4 McBride & Spears (2001) Belgium (source unknown) 98.6 El Ghawi, et al. (1999) Germany (source FAL sample collection, 82.1 unknown) 2003-2007 NPK Fertilizers USA 65.5 McBride & Spears (2001) Yamazaki & Geraldo Brazil 27.1 (2003) Romania (source Pantelica, et al. (1997) 42.7 unknown) Hungary (sourced from FAL sample collection, 0.5 Kola Peninsula, Russia) 2003-2007 Belgium (source unknown) 46 El Ghawi, et al. (1999) Germany (source FAL sample collection, 9.9 unknown) 2003-2007 Table D- 41 U concentrations in different types of P-containing mineral compound fertilizers from their respective countries, adapted from Kratz, et al. (2008). Fertilizers are sourced from their own country unless specified otherwise.

152