Production and Export from Serpentine Soil of the California Coast

Nov 22, 2016 - speciation (i) in four serpentine soil depth profiles derived from the California Coast. Range serpentinite belt and (ii) in local surf...
3 downloads 0 Views 1MB Size
Subscriber access provided by Olson Library | Northern Michigan University

Article

Quantifying Cr(VI) production and export from serpentine soil of the California Coast Range Cynthia N. McClain, Scott Fendorf, Samuel M. Webb, and Kate Maher Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b03484 • Publication Date (Web): 22 Nov 2016 Downloaded from http://pubs.acs.org on November 27, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 29

1 2 3 4 5 6 7 8 9 10 11 12

Environmental Science & Technology

Quantifying Cr(VI) production and export from serpentine soil of the California Coast Range Cynthia N. McClaina,*, Scott Fendorfb, Samuel M. Webbc, Kate Mahera,* a b

Department of Geological Sciences, Stanford University, Stanford, CA, 94305, USA Department of Earth System Science, Stanford University, Stanford, CA, 94305, USA c Stanford Synchrotron Radiation Lightsource, Menlo Park, CA 94025, USA * Corresponding author E-mail: [email protected], Phone: +1-650-223-5790 * Corresponding author E-mail: [email protected], Phone: +1-650-725-0927

ABSTRACT

13

Hexavalent chromium (Cr(VI)) is generated in serpentine soils and exported to

14

surface and groundwaters at levels above health-based drinking water standards.

15

Although Cr(VI) concentrations are elevated in serpentine soil pore water, few studies

16

have reported field evidence documenting Cr(VI) production rates and fluxes that govern

17

Cr(VI) transport from soil to water sources. We report Cr speciation (i) in four serpentine

18

soil depth profiles derived from the California Coast Range serpentinite belt and, (ii) in

19

local surface waters. Within soils, we detected Cr(VI) in the same horizons where Cr(III)-

20

minerals are co-located with biogenic Mn(III/IV)-oxides, suggesting Cr(VI) generation

21

through oxidation by Mn-oxides. Water-extractable Cr(VI) concentrations increase with

22

depth constituting a 7.8 to 12

23

produced at a rate of 0.3 to 4.8 kg Cr(VI)/km2/yr and subsequently flushed from soil

24

during water infiltration, exporting 0.01 to 3.9 kg Cr(VI)/km2/yr at concentrations

25

ranging from 25 to 172 µg/L. Although soil-derived Cr(VI) is leached from soil at

26

concentrations exceeding 10 µg/L, due to reduction and dilution during transport to

27

streams, Cr(VI) levels measured in local surface waters largely remain below California’s

28

drinking water limit.

kg/km2 reservoir of Cr(VI) in soil. Here, Cr(VI) is

1 ACS Paragon Plus Environment

Environmental Science & Technology

29 30

TOC

31 32 33

INTRODUCTION

34

Surface and groundwater are contaminated by Cr(VI) at levels above drinking

35

water limits (e.g. World Health Organization – 50 µg Cr/L, California – 10 µg Cr(VI)/L)

36

as a result of geogenic Cr(III) mineral oxidation.1,2 In the United States, 0.92 to 74

37

million people are exposed to hazardous Cr(VI) concentrations (> 10 µg/L) in drinking

38

water,3,4 which poses a human health threat through an increased risk for stomach and

39

intestinal cancer.5,6

40

Due to the high Cr content, ultramafic rock (2400 mg/kg),7 their metamorphic

41

derivatives (serpentinite), and weathering products (serpentine soil and sediments) are the

42

most important sources of Cr(VI) produced from geogenic sources within drinking water.

43

Serpentine soils develop during weathering of ultramafic bedrock, creating ideal

44

biogeochemical conditions for Cr(VI) generation, accumulation, and export. Chromium

45

minerals accumulate in soil, increasing Cr concentrations (29 to 80000 mg/kg),8,9 with

46

coinciding increases in Mn concentrations resulting from weathering of ultramafic 2 ACS Paragon Plus Environment

Page 2 of 29

Page 3 of 29

Environmental Science & Technology

47

rocks.10,11 Chromium(III)-bearing primary minerals within weathered ultramafic rock

48

have morphological features characteristic of dissolution, suggesting Cr(III) release.12,13

49

Chromium(VI) is thought to be generated in soil via oxidation of Cr(III) by proximal

50

Mn(III/IV)-oxide minerals,14-16 reaching concentrations up to ~ 7000 µg/kg.16-19 As a

51

consequence, Mn oxidizing bacteria can accelerate the Cr(VI) generation rate through the

52

production of the primary Cr(III) oxidant.20 Cr(VI) production21 rates vary widely

53

depending on Cr(III) source (e.g. silicates, spinels, FexCr1-x(OH)3, aqueous), oxidation

54

pathway (e.g. different Mn-oxide minerals with varying Mn(III/IV) content, O2, reactive

55

oxygen species), and pH.1,22-29 Countering production, Cr(VI) can be attenuated by

56

reduction (e.g. by Fe(II), sulfide, organic matter, microbes) and adsorption onto mineral

57

surfaces.30,31 As a result of the complexity underlying Cr release and oxidation rates, the

58

rates associated with natural systems have not yet been quantified.

59

Serpentine soils have shown Cr levels in soil water that exceed health based

60

drinking water standards, ranging from 5 – 250 µg/L. 32-35 For comparison, surface waters

61

in ultramafic catchments have Cr(VI) concentrations up to 19 µg/L.2,33,36 Due to the high

62

Cr concentration in soil water, soil-derived Cr(VI) may be an important source of Cr(VI)

63

to surface and groundwater water. However, the inventory of Cr(VI) within serpentine

64

soils, along with the rates of production and export, are largely unresolved.

65

California, USA, is a region with (i) abundant ultramafic bedrock (2860 km2) and

66

soils (Figure S1),37 and (ii) elevated Cr(VI) levels in surface and groundwater.2,33,38-42 In

67

California, 0.1 to 7.8 million individuals may be exposed to hazardous concentrations of

68

Cr(VI) and other trace elements in drinking water.3,43 Here, we studied soil derived from

69

the northern California Coast Range serpentinite belt. Our objectives were (i) to quantify

3 ACS Paragon Plus Environment

Environmental Science & Technology

70

the Cr(VI) reservoir, production rate and export fluxes from serpentine soil, and (ii) to

71

compare these fluxes to Cr(VI) fluxes in local surface water. We collected depth resolved

72

soil samples from four profiles and water samples, with concurrent discharge rates, from

73

a nearby creek. We determined physical and mineralogical properties of soil, chemical

74

composition and speciation of soil solids and water extracts (e.g. Cr(VI), Cr(III)), centi-

75

to micro-meter scale imaging of the spatial distribution and oxidation state of Cr and Mn

76

minerals, and tested for Mn oxidizing bacteria. Based on these results, we evaluate

77

approaches for determining in situ Cr(VI) production rates. Our results provide a basis for

78

understanding the biogeochemistry of Cr(VI) generation in the context of the hydrologic

79

processes governing Cr(VI) release from serpentine soils.

80 81

MATERIALS AND METHODS

82

Study Area. Serpentine soils, derived from the California Coast Range

83

serpentinite belt, are abundant in the Putah Creek/Berryessa watershed of northern

84

California, and constitute a potential source of Cr(VI) contamination to drinking water

85

(Figure S1). The major surface water reservoir in the watershed, Lake Berryessa, serves

86

as a drinking water source for more than 0.35 million people (Vallejo, Fairfield,

87

Vacaville, Suisun City) and supplies groundwater recharge (natural and irrigation return)

88

in the southwestern Sacramento Valley where groundwater is used for drinking (e.g.

89

Davis). Our field site lies near the headwaters of the Putah Creek/Berryessa watershed, in

90

upper Hunting Creek watershed (16.8 km2), within University of California’s

91

McLaughlin Natural Reserve (38.86, -122.404 [WGS 84]) (Figure S1).

4 ACS Paragon Plus Environment

Page 4 of 29

Page 5 of 29

Environmental Science & Technology

92

At our field site, Montara series Mollisols predominate, which are well-drained

93

soils derived from the Berryessa serpentinite complex.44,45 Grasses and manzanita

94

(Arctostaphylos viscida) proliferate on southwest facing slopes.

95

Precipitation in the Mediterranean climate (Xeric moisture regime, mean annual

96

temperature of 15°C, mean annual precipitation of 720 mm),46 during winter months

97

leads to through-flow and runoff, contributing to groundwater recharge and surface water

98

flow. Chromium concentrations in Hunting Creek and its tributaries range from below

99

detection (< 0.3 µg/L) to 19 µg/L.2

100

Soil Sample Collection. Field-moist soil samples were collected every 5-10 cm

101

with depth from 4 profiles (numbered 1-4) along a toposequence on the shoulder of the

102

meadow by hand digging soil pits or hand augering until refusal depth ~ 60 cm (bedrock

103

interface), and stored in airtight vials at 4°C prior to laboratory analysis to minimize

104

oxygen exposure. Bulk density samples were collected using the short core method

105

(Table S1).47 Mollisol profiles typically have brown A, and Bt horizons (0-32 cm)

106

underlain by olive colored saprolite (C horizon, 32-60 cm) (Table S2). Bioturbation from

107

rooting of vegetation, and burrowing gophers was limited to < 50 cm. Saturated hydraulic

108

conductivity was measured at multiple depths in soil profiles using a Guelph

109

Permeameter, operating under constant head conditions.48

110

Water Sample Collection. Water samples were collected during a rainstorm on

111

April 1, 2014 from Hunting Creek, tributaries draining the meadow where soil samples

112

were collected, and from overland flow on the meadow. Water samples were filtered

113

using Geotech in-line high capacity 0.45 µm cartridges, split into two subsamples

114

(unacidified and acidified to pH < 2 with nitric acid), and stored at 4°C prior to laboratory

5 ACS Paragon Plus Environment

Environmental Science & Technology

115

analysis. Volumetric flow rates were measured by the velocity-area method. Watershed

116

area was calculated using Google Earth’s polygon tool to determine contributing area

117

above each sampling location, delineated by topographic ridges. Stream-based specific

118

discharge (runoff) was calculated by normalizing the measured volumetric discharge to

119

watershed area.

120

Solid-Phase Analysis and Sample Preparation. Gravimetric water content was

121

determined by difference between field moist and oven dry soil (104°C). Particle size

122

distribution (< 2 mm) and soil texture were determined at the University of Idaho by

123

sieving and the pipette method (Table S3). X-ray diffraction (XRD) was performed on

124

powdered bulk soil samples and clay separates as described in Supporting Information

125

(Figures S2, S3, Table S4). Thin sections were prepared from oven dried soil aggregates

126

embedded in Epotek 301 2-FL epoxy resin, that were cut and polished to 30 µm using

127

heat and oxygen sensitive techniques by Spectrum Petrographics (Vancouver, WA).

128

Mineral modal abundance was determined by optical petrography and point counting of

129

thin sections on a Nikon Optiphot polarizing petrographic microscope (Table S5).

130

Regolith is composed of serpentine minerals (lizardite, antigorite, clinochrysotile), clay

131

(smectite, vermiculite, chlorite), actinolite, and oxide minerals (e.g. Fe-Cr-spinels,

132

secondary Mn-Fe-oxides) (Figure S4). Thin section imaging and determination of Fe,

133

Mn, Ni and Cr content of minerals was performed by Electron Microprobe Analysis

134

(EMPA) using a JEOL JXA-8230 SuperProbe. Oven dried soil samples were also

135

disaggregated by gentle grinding with an agate mortar and pestle and analyzed for bulk

136

major and trace element content by X-ray fluorescence spectrometry (Spectro XRF-

137

XEPOS).

6 ACS Paragon Plus Environment

Page 6 of 29

Page 7 of 29

Environmental Science & Technology

138

Extractions. Extractions of field moist soil samples were performed to quantify

139

major and trace element content of two pools: “water-exchangeable” and “dithionite-

140

extractable” (nominally from Fe- and Mn-oxides [citrate-bicarbonate-dithionite –

141

CBD]).49 Extractions for the “water-exchangeable” pool were performed for 24 h with 30

142

g of soil and 30 g of deionized water, vacuum filtered through Millipore 0.2 µm Steriflip

143

filters, split into two subsamples (unacidified and acidified to pH < 2 with nitric acid),

144

and stored at 4°C prior to chemical analysis. Concentrations are reported in mg or µg per

145

kg of dry soil. Mean normalized concentrations for profiles i = 1 – 4 are expressed as: 

1   =   4      

(1)



146

where X is concentration in mg/kg measured at each sampling depth (z) or at the deepest

147

sample from near the saprolite-bedrock interface. Chromium(VI) concentrations in soil

148

water (c [µg Cr(VI)/L of soil water]) are estimated from water-exchangeable Cr(VI)

149

concentrations (c [µg Cr(VI)/kg of dry soil]), using measured gravimetric water content

150

(w [g water/g dry soil]) and water density (ρw [g/cm3) according to: c = C × ρw / w.

151

Aqueous Chemical Analysis. Acidified water samples and extracts were

152

analyzed for major and trace element content with inductively coupled plasma optical

153

emission spectroscopy (Thermo Scientific ICAP 6300 Duo View ICP-OES) and mass

154

spectroscopy (Thermo Scientific XSERIES 2 ICP-MS), respectively. Unacidified water

155

samples and water-exchangeable extracts were measured for pH (ranging from 7.2 to

156

8.3), inorganic anions by ion chromatography (Dionex DX-500), alkalinity by titration

157

with 0.04 M sulfuric acid to a colorimetric endpoint (pH 4.5) using bromocresol green-

158

methyl red indicator, and dissolved organic carbon (DOC, measured as non-purgeable

7 ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 29

159

organic carbon) on a Shimadzu TOC-L. Chromium(VI) concentrations were determined

160

by

161

spectrophotometer.14,50 Chromium(III) concentrations were calculated by difference

162

between Cr(TOT), measured by ICP-MS, and Cr(VI).

a

modified

s-diphenyl

carbazide

method

on

a

Shimadzu

UV-1601

163

Microbiological Analysis. To enrich Mn oxidizing bacteria from field moist soils

164

(from 0-4 cm depth), serial dilutions (1/104) in artificial seawater media were plated on

165

agar-solidified K-media containing 20 mM HEPES buffer and 100 µM MnCl2 (pH 7.6).

166

Mn(III/IV)-oxides were confirmed with the Leucoberbelin Blue (LBB) colorimetric

167

assay.51

168

Synchrotron Spectroscopy and Element Mapping. Synchrotron-based X-ray

169

microprobe measurements of element distributions and oxidation state in regolith thin

170

sections from profile 1 were conducted on beam lines 10-2 and 2-3 at the Stanford

171

Synchrotron Radiation Lightsource as described in Supporting Information. Element

172

maps were generated by rastering the sample around the X-ray beam, collecting X-ray

173

fluorescence data, and using the Microanalysis Toolkit for processing.52 Each micro X-

174

ray absorption near edge structure (µXANES) spectrum was collected from 200 eV

175

below the Mn or Cr K-edge to 400 eV above the edge. All spectra were background

176

subtracted and normalized to a unit step edge with the iXAFS Athena software package.53

177

Soil Hydraulic Properties. To calculate the flux of Cr(VI) under unsaturated

178

conditions, we use a special case of Darcy’s law; the infiltration flux, is equal to the

179

unsaturated hydraulic conductivity, assuming a unit hydraulic gradient that may result

180

from either constant pressure head or local extreme of pressure head within a vertical

181

section of soil.54 Unsaturated hydraulic conductivity was calculated from measured

8 ACS Paragon Plus Environment

Page 9 of 29

Environmental Science & Technology

182

saturated hydraulic conductivity, using the Van Genuchten equation (1980) as described

183

in Supplemental Information.55

184

RESULTS

185

Trends in Major and Trace Element Chemistry of Serpentine Regolith. The

186

vertical distribution of major and trace elements systematically varies with depth in

187

relation to soil horizons and mineralogy. The saprolite is characterized by low total,

188

dithionite-extractable and water-exchangeable concentrations of major and trace elements

189

(Figure 1). In contrast, the soil, which has more abundant clay and Fe/Mn-oxides (Table

190

S5), is enriched in Al, Fe, Mn and Cr. Additionally, the concentrations of water-

191

exchangeable trace elements are highest in the soil (Figure 2).

9 ACS Paragon Plus Environment

Environmental Science & Technology

192 193 194 195 196 197 198

Figure 1. Mean normalized concentration () of major (Si, Mg, Ca, Al) and trace (Fe, Mn, Cr, Ni) elements from 4 serpentinite regolith profiles as a function of depth, normalized to the concentration in each extraction at the saprolite/bedrock interface. (a, d) Total (XRF), (b, e) dithionite-extractable (CBD), (c, f) water-extractable. Filled circles indicate sampling locations and are shown only on Si and Fe profiles for clarity.

199

Chromium Concentrations. Chromium concentrations are 1.2 to 3 times higher

200

in soil compared to saprolite, across all Cr reservoirs (Figures 1 and 2). Approximately

201

4% of total Cr resides in the dithionite-extractable fraction (decreasing with depth), and

202

0.001% is water-extractable. Water-extractable Cr is mainly Cr(VI), with maximum 10 ACS Paragon Plus Environment

Page 10 of 29

Page 11 of 29

Environmental Science & Technology

203

Cr(VI) levels in soil of 27 µg/kg, compared to 12 µg/kg in saprolite from 50-60 cm depth.

204

Water-exchangeable Cr(III) concentrations increase (up to 15 µg/kg) in near surface soils

205

(5 cm depth), coinciding with high DOC and alkalinity. However, in two profiles, water-

206

exchangeable Cr(III) levels are highest (21 and 57 µg/kg) near the soil-saprolite interface

207

(~ 40 cm). Water samples collected during a rainstorm from both overland flow in the

208

meadow, and surface water runoff have a narrow range of Cr(VI) concentrations from 5.5

209

to 6.9 µg/L, and low Cr(III) concentrations (< 0.6 µg/L).

210 211 212 213 214 215 216

Figure 2. Water-extractable concentrations of (a) Cr(VI), (b) Cr(III), (c) dissolved organic carbon (DOC) and, (d) alkalinity per kg dry soil from 4 serpentinite regolith depth profiles (filled circles). Mean concentrations shown as black squares with error bars (1 SD). Occurrence and Oxidation State of Cr and Mn in Regolith Minerals at the

217

Micrometer Scale. Chromium(III) and Mn(III/IV) minerals are juxtaposed, residing

218

within microns of each other in all soil and saprolite samples (Figures 3 and S4).

219

Chromium(III)-bearing minerals are present both as small diameter (50 µm), etched and

220

embayed, high Cr concentration minerals (e.g. Cr- and Fe-spinels such as chromite

221

[spectrum 4 in Figure 3] and magnetite), and as more diffuse sources (e.g. serpentine

11 ACS Paragon Plus Environment

Environmental Science & Technology

222

minerals [spectrum 2 in Figure 3], Fe(Cr)-oxides, and clay minerals) (Table S5).

223

Although water-extractable Cr(VI) was measured in soil, it resided at levels below

224

detection by our µXANES spectroscopic analysis (Figure 3). Manganese oxides and Mn-

225

bearing silicates in serpentine the regolith have varying oxidation state, based on the

226

position of the Mn K-edge in the energy range from 6547 to 6552 eV.

227

Manganese(III/IV)-oxides similar to birnessite are found coating mineral grains (spectra

228

7 and 12) and as < 2 µm clusters (spectrum 9), both in close proximity to Cr(III)

229

minerals. Additionally, Mn oxidizing bacteria were detected in surface soil (Figure S5).

230 231 232 233 234 235 236 237

Figure 3. X-ray fluorescence maps showing relative intensity of Cr (magenta), Mn (yellow), and Fe (cyan) in serpentinite regolith (profile 1) from (a) 4 cm depth, (b, d) 26 cm depth, and (c) 39 cm depth. Representative, normalized µXANES spectra of reference minerals and from locations numbered in panels a-d for (e) Cr K-edge and (f) Mn K-edge. Shaded gray regions in panel f indicate energies characteristic of Mn(II) (6547-6553 eV), Mn(III) (6554-6559 eV) and Mn(IV) (6560-6562 eV).

12 ACS Paragon Plus Environment

Page 12 of 29

Page 13 of 29

238

Environmental Science & Technology

DISCUSSION

239

Cr(VI) Generation in Serpentine Soil. Chromium(VI) is produced in serpentine

240

soil as a result of both long-term (100’s to 1000’s of years) soil formation that

241

concentrates and redistributes Cr and Mn, and contemporary biogeochemical reactions

242

that produce Cr(VI). Vertical and lateral transport of solutes further redistribute reaction

243

products, with a dependence on landscape position, soil hydraulic properties, and

244

precipitation (rainfall) intensity. Figure 4 shows our conceptual model of geochemical

245

reactions and movement of water through serpentine soil that results in Cr(VI)

246

production, accumulation, and export.

247 248 249

Figure 4. Conceptual cycle of Cr(VI) production and export from serpentinite regolith. 13 ACS Paragon Plus Environment

Environmental Science & Technology

250 251

During pedogenesis Cr-bearing primary minerals (e.g. spinels and serpentine)

252

dissolve (Figure S4), repartitioning ~ 4% of Cr into secondary minerals resulting in Cr

253

accumulation in clay and oxyhydroxides (Tables S1c and S6), as also noted by Oze et al.

254

(2004) and Morrison et al. (2015).13,56 The patterns of Cr and Mn partitioning we

255

observed are comparable to other studies in temperate climates where ~ 10 % of Cr was

256

associated with Fe-oxides,57,58 and Cr repartitioned to oxyhydroxides in the B

257

horizon.10,11 Some Cr will also be lost to the fluid phase during weathering.9,59 However,

258

the majority of Cr (~ 96%) remains in low solubility primary minerals that preferentially

259

accumulate as the surrounding, more easily weathered, soil matrix collapses. In contrast

260

to Cr, more Mn (> 10%) in soil is hosted in secondary minerals (e.g. oxides), having been

261

redistributed from primary minerals during soil formation. Given the enrichment of Mn

262

oxidizing bacteria from surface soil (0-4 cm depth, [Figure S5]), neoformed Mn-oxides

263

may be biologically promoted, giving rise to highly reactive Mn-oxide phases.60 Mixed

264

Mn oxidation state (III/IV) provides evidence of Mn redox cycling,61 whereby Mn-oxide

265

surfaces can be regenerated with Mn2+, a byproduct of Cr(III) oxidation. Neoformed

266

Mn(III/IV) minerals are co-located (within microns) with etched and pitted Cr(III)-

267

bearing spinel and silicate minerals, and at various locations, they surround Cr(III)-

268

bearing mineral surfaces (Figures 3 and S4). Within soil, water-extractable Cr(VI)

269

concentrations are also high (Figure 2). Together, these observations support the potential

270

for continual dissolution of Cr(III) from oxides and silicates and subsequent Cr(III)

271

oxidation by biogenic Mn(III/IV)-oxides in serpentine soils developed in temperate

272

climates, similar to previous studies in tropical climates,15,16 and in the laboratory.20,24-26

14 ACS Paragon Plus Environment

Page 14 of 29

Page 15 of 29

Environmental Science & Technology

273

As these soils have low iron oxide content and alkaline pH values, little Cr(VI) will

274

adsorb to soil matrix, and Cr(VI) will be readily transported vertically to saprolite and

275

groundwater, and laterally to surface water.

276

Coupled biogeochemical and hydrologic processes generate depth gradients in

277

water-exchangable Cr. Water-exchangeable Cr concentrations are highest in the B

278

horizon, where Cr(VI) and Cr(III) concentrations reach up to 25 and 15 µg/kg,

279

respectively (Figure 2). Here, total Cr and Mn, and dithionite extractable Cr levels are

280

highest. In this horizon, dissolved organic acids (e.g. acetate, citrate) and increased

281

carbon dioxide fugacity may enhance Cr(III) mineral dissolution,62,63 stabilize Cr(III) in

282

solution,64,65 or enhance Cr(VI) mobility due to decreased Cr(VI) sorption in the presence

283

of carbonate.66,67 Near the soil-saprolite interface (40 cm), anomalously high Cr(III)

284

concentrations may result from heterogeneity or localized reducing conditions near

285

manzanita roots. In the saprolite (from 50 to 60 cm), Cr(VI) concentrations may be lower

286

due to lower total Cr concentrations, Cr(VI) reduction in areas with limited oxygen

287

supply, or from flushing of Cr(VI) due to preferential lateral flow above the saprolite-

288

bedrock interface.

289

Calculated Cr(VI) concentrations in soil pore water under unsaturated conditions

290

(c [µg Cr(VI)/L of soil water]) all exceed California’s drinking water limit of 10 µg/L,

291

ranging from 25 to 172 µg/L (Table S7). These calculations, based on water extractions

292

performed in the laboratory, may overestimate field measurements due to lower

293

soil/water ratios, different water chemistry, or release of exchangeable Cr(VI) from pores

294

and surfaces that may not otherwise be accessible. Nevertheless, calculated Cr(VI)

295

concentrations in soil water are similar to measurements of dissolved plus fine colloids

15 ACS Paragon Plus Environment

Environmental Science & Technology

296

(filtered to < 0.45 µm) in soil water collected by tension-free lysimetry by Gasser et al.

297

(1994) (9-104 µg/L).32 This suggests that calculating soil water Cr(VI) concentrations

298

from measured water content and Cr(VI) in water extractions and may be a good proxy

299

for in-situ soil water Cr(VI) levels when field measurements are not available.

300

Importantly, Cr may also be present as colloids,32 however in this study we focus on

301

dissolved Cr(VI) and use calculated soil water Cr(VI) concentrations to estimate Cr(VI)

302

fluxes and export rates from serpentine soils.

303

Cr(VI) Accumulation in Serpentine Soil. Over the top 40 cm of the regolith, we

304

integrate the profile of water-extractable Cr(VI) concentrations per mass of soil (C

305

[µg/kg]) with depth (z [m]) (Figure 2), estimating the inventory of water-extractable

306

Cr(VI) to be 7.8 to 12 kg/km2 (I = C × z × ρb [kg Cr(VI)/km2], where ρb is bulk density

307

[g/cm3]) (Figure 4, Tables S1 and S7). This Cr(VI) reservoir is ~ 16 times smaller than in

308

unsaturated ultramafic-bearing sediments of southwest Sacramento Valley, California

309

because serpentine soils are shallower than the valley’s unsaturated zone.68 In these

310

California Coast Range serpentine soils, Cr(VI) only makes up ~ 0.001% of the total Cr

311

reservoir in soil (1.1 × 106 kg/km2), yet contributes to geogenic Cr(VI) contamination of

312

surface water and groundwater during infiltration, runoff and recharge.

313

Cr(VI) Flux through Serpentine Soil. During infiltration through unsaturated

314

soil Cr(VI) will be transported downward according to the water infiltration flux (q

315

[mm/yr], here assumed to equal unsaturated hydraulic conductivity), at Cr(VI)

316

concentrations in soil water (c [µg Cr(VI)/L of soil water]) (Tables S1 and S7).

317

Infiltration fluxes calculated here for unsaturated serpentine soil are spatially variable,

318

ranging from 0.03 to 65 mm/yr, with a mean travel time (τ ) of 7.3 years through the top 16 ACS Paragon Plus Environment

Page 16 of 29

Page 17 of 29

Environmental Science & Technology

319

40 cm of soil (τ = θ × z ⁄ q, [yr], volumetric water content is θ = w × ρb / ρw, where ρw is

320

water density (g/cm3)) (Figure 4, Table S7). These water fluxes have large uncertainties,

321

primarily due to the variability in saturated hydraulic conductivity measurements, which

322

typically have uncertainties of up to two orders of magnitude.69 However, because

323

hydraulic conductivity can vary by orders of magnitude with depth and along the

324

hillslopes (where we measured),54,70 these calculated infiltration fluxes are thought to be

325

representative of natural variability in this meadow. We estimate that the resulting flux, J,

326

of Cr(VI) (J = q × c [kg/km2/yr]) through meadow soil to surface or groundwater is

327

spatially variable, ranging from 0.01 to 3.9 kg/km2/yr, with a mean of 1.1 kg/km2/yr

328

(Figure 4, Table S7). This flux provides a minimum constraint on the long-term average

329

Cr(VI) production rate.

330

Cr(VI) Production Rate. To our knowledge, no field-estimates of Cr(VI)

331

production in soils have been developed. Such estimates are important for evaluating the

332

role of soils and unsaturated zones in contributing to Cr(VI) contamination of

333

groundwaters and surface waters. Here, we calculate the net rate of Cr(VI) generation in

334

three different ways: we estimate it using two different mass balance approaches that

335

require knowledge of the infiltration flux, and then compare these results to a laboratory-

336

determined rate law for Fe-Cr-spinel oxidation by birnessite that does not depend on the

337

infiltration flux.1 First, we simulate the top 40 cm of each profile as a completely mixed

338

flow reactor (CMFR), assuming the system has spatially uniform properties and the only

339

source of Cr(VI) to the system is in situ generation, due to the low Cr(VI) concentrations

340

in rainwater.71 The net in situ Cr(VI) production rate (R [µg/L/yr]) can be expressed as:

17 ACS Paragon Plus Environment

Environmental Science & Technology

=

1   + "# 

Page 18 of 29

(2)

341

where m is the water to soil ratio (w × ρb × z /ρw), I (kg/km2) is the amount of Cr(VI) that

342

accumulates in soil during the time (τ [yr]) it takes water to travel through soil, J is the

343

Cr(VI) export flux (kg/km2/yr) (Tables S1 and S7). We estimate the Cr(VI) production

344

rate in serpentine soil to range from 1.2 to 17.3 µg/L/yr, which is equivalent to 0.19 to

345

2.35 kg/km2/yr using the conversion factor m (Table S7). This CMFR model is a

346

reasonable representation of the relatively constant Cr(VI) concentration with depth for

347

profiles 3 and 4 (located farther downslope than profiles 1 and 2), where the assumption

348

of spatially uniform properties may be valid (Figure S6). This homogeneity may originate

349

from the lower infiltration fluxes and/or greater mixing due to complex, two-dimensional,

350

water and soil movement (e.g. eluviation, creep, deposition) down the hillslope.

351

Second, we simulate profiles 1 and 2 as plug flow reactors (PFRs), which is

352

consistent with Cr(VI) concentrations that increase with depth. The change in Cr(VI)

353

concentration in soil water with depth (c(z) [µg/L]) during steady-state vertical advective

354

transport, with Cr(VI) generation described by zeroth-order kinetics (typical for surface

355

catalyzed reactions) can be expressed as: $%&' = $ +

(& )

(3)

356

where co is Cr(VI) concentration at the ground surface (25 µg/L) (equation development

357

and assumptions described in Supporting Information). Using equation (2) to fit our

358

Cr(VI) concentration data with depth by adjusting R, the best fits are Cr(VI) production

359

rates of 21 and 30 µg/L/yr (2.92 and 2.96 kg/km2/yr) for profiles 1 and 2, respectively,

360

assuming mean θ and q for each profile (Figure S6a, Table S7). The PFR model provides 18 ACS Paragon Plus Environment

Page 19 of 29

Environmental Science & Technology

361

a better fit than the CMFR for profiles 1 and 2 (located closer to the summit) suggesting

362

that depth gradients in concentration arise from the balance between net Cr(VI)

363

production and net downward transport due to infiltration, as per equation (3) above. The

364

fit of two different mass balance models for Cr(VI) concentration in soil water (CMFR,

365

PFR) reveals how coupled hydrologic and geochemical processes influence the depth

366

distribution of Cr(VI) at different landscape positions along a hillslope.

367

Third, the Cr(VI) production rate from oxidation of Fe-spinel (the mineral that

368

hosts the most Cr by mass [Table S6]) by the Mn(IV)-oxide birnessite, parameterized by

369

data measured at our field site, can be described after Oze et al. (2007) by:1 :. 99% of Cr is immobile, mineral-bound Cr(III). Soil Cr(III)-minerals, while

436

limited in solubility, constitute a seemingly infinite source of Cr that can be oxidized to

437

Cr(VI) and transmitted to the hydrosphere, given the proper conditions. We show that

438

Cr(III)-minerals reside proximal to Mn(III/IV)-oxide minerals, providing a pathway for

439

generating a reservoir of Cr(VI). Although it constitutes < 0.001% of the Cr reservoir in

440

serpentine soils, Cr(VI) is flushed to surface and groundwater at concentrations above 10

441

µg/L during infiltration. However, soil-derived Cr(VI) is attenuated by dilution and/or

442

reduction before reaching surface water ways, leading to Cr(VI) concentrations largely 22 ACS Paragon Plus Environment

Page 22 of 29

Page 23 of 29

Environmental Science & Technology

443

below California’s drinking water limit in surface water of the Berryessa/Putah Creek

444

watershed.2 Nevertheless, Cr removed by reduction in streambed sediments may be

445

transported downstream, loading depositional environments such as alluvial plains with

446

Cr(III)-phases susceptible to re-oxidation. This study summarizes the size of water-

447

extractable Cr(VI) reservoir in serpentine soil (~10 kg/km2), provides the first estimate of

448

in situ Cr(VI) production rate (~2.5 kg/km2/yr), and constrains the amount of Cr(VI)

449

released from soil during infiltration (~1.1 kg/km2/yr), relative to Cr(VI) levels measured

450

in surface water used for drinking.

451 452

ASSOCIATED CONTENT

453

Supporting Information. Additional method and equation details; list of abbreviations,

454

symbols, equations, and parameters; tables; figures. This information is available free of

455

charge via the Internet at http://pubs.acs.org/.

456 457

ACKNOWLEDGEMENTS

458

This work was supported by the U.S. National Science Foundation (Graduate Research

459

Fellowship DGE-114747 to CNM, and EAR-1254156 to KM), and Stanford University’s

460

McGee Grant, and conducted at the University of California’s Donald and Sylvia

461

McLaughlin Natural Reserve. We thank Cathy Koehler, Paul Aigner, Scott Moore, Dr.

462

Guangchao Li, Doug Turner, Dr. Juan Lezama Pacheco, Dr. Adam Jew, Anita Falen and

463

numerous students from Stanford University for field and laboratory assistance. Use of

464

the Stanford Synchrotron Radiation Lightsource, SLAC National Accelerator Laboratory,

465

is supported by the U.S. Department of Energy, Office of Science, Office of Basic

23 ACS Paragon Plus Environment

Environmental Science & Technology

466

Energy Sciences under Contract No. DE-AC02-76SF00515. The SSRL Structural

467

Molecular Biology Program is supported by the DOE Office of Biological and

468

Environmental Research, and by the National Institutes of Health, National Institute of

469

General Medical Sciences (including P41GM103393). The contents of this publication

470

are solely the responsibility of the authors and do not necessarily represent the official

471

views of NIGMS or NIH.

472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505

REFERENCES (1) Oze, C.; Bird, D. K.; Fendorf, S. Genesis of hexavalent chromium from natural sources in soil and groundwater. Proc. Natl. Acad. Sci. 2007, 104 (16), 6544– 6549. (2) McClain, C. N.; Maher, K. Chromium fluxes and speciation in ultramafic catchments and global rivers. Chem. Geol. 2016, 426, 135–157. (3) Occurrence estimation methodology and occurrence findings report for the sixyear review of existing national primary drinking water regulations; EPA-815R-03-006; United States Environmental Protection Agency Office of Water: Washington, DC, 2003; www.epa.gov/sites/production/files/201412/documents/815r03006.pdf. (4) Chromium-6 in U.S. tap water; Environmental Working Group: Washington, DC, 2010; www.ewg.org/chromium6-in-tap-water. (5) Beaumont, J. J.; Sedman, R. M.; Reynolds, S. D.; Sherman, C. D.; Li, L.-H.; Howd, R. A.; Sandy, M. S.; Zeise, L.; Alexeeff, G. V. Cancer mortality in a Chinese population exposed to hexavalent chromium in drinking water. Epidemiology 2008, 19 (1), 12–23. (6) Toxicology and carcinogenesis studies of sodium dichromate dehydrate in F344/N rats and B6C3F1 mice; NTP TR 546; National Toxicology Program: Research Triangle Park, NC, 2008; ntp.niehs.nih.gov/ntp/htdocs/lt_rpts/tr546.pdf. (7) Nriagu, J. O.; Nieboer, E. Chromium in the natural and human environments; Wiley-Interscience: New York, 1988. (8) Oze, C.; Fendorf, S.; Bird, D. K.; Coleman, R. G. Chromium geochemistry of serpentine soils. Int. Geol. Rev. 2004, 46 (2), 97–126. (9) Kierczak, J.; Neel, C.; Bril, H.; Puziewicz, J. Effect of mineralogy and pedoclimatic variations on Ni and Cr distribution in serpentine soils under temperate climate. Geoderma 2007, 142 (1-2), 165–177. (10) Lee, B. D.; Graham, R. C.; Laurent, T. E.; Amrhein, C. Pedogenesis in a wetland meadow and surrounding serpentinitic landslide terrain, northern California, USA. Geoderma 2004, 118 (3), 303–320. (11) Caillaud, J.; Proust, D.; Philippe, S.; Fontaine, C.; Fialin, M. Trace metals distribution from a serpentinite weathering at the scales of the weathering profile 24 ACS Paragon Plus Environment

Page 24 of 29

Page 25 of 29

506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551

Environmental Science & Technology

(12) (13)

(14) (15)

(16)

(17) (18) (19)

(20) (21)

(22)

(23)

(24)

(25) (26)

(27)

and its related weathering microsystems and clay minerals. Geoderma 2009, 149 (3-4), 199–208. Garnier, J.; Quantin, C.; Guimarães, E.; Becquer, T. Can chromite weathering be a source of Cr in soils? Mineral. Mag. 2008, 72 (1), 49–53. Morrison, J. M.; Goldhaber, M. B.; Mills, C. T.; Breit, G. N.; Hooper, R. L.; Holloway, J. M.; Diehl, S. F.; Ranville, J. F. Weathering and transport of chromium and nickel from serpentinite in the Coast Range ophiolite to the Sacramento Valley, California, USA. Appl. Geochem. 2015, 61, 72–86. Bartlett, R. Behavior of chromium in soils. III. Oxidation. J. Environ. Qual. 1979, 8 (1), 31–34. Garnier, J.; Quantin, C.; Guimarães, E. M.; Vantelon, D.; Montargès-Pelletier, E.; Becquer, T. Cr(VI) genesis and dynamics in Ferralsols developed from ultramafic rocks: The case of Niquelândia, Brazil. Geoderma 2013, 193-194, 256–264. Fandeur, D.; Juillot, F.; Morin, G.; Olivi, L.; Cognigni, A.; Webb, S. M.; Ambrosi, J.-P.; Fritsch, E.; Guyot, F.; Brown, G. E., Jr. XANES evidence for oxidation of Cr(III) to Cr(VI) by Mn-oxides in a lateritic regolith developed on serpentinized ultramafic rocks of New Caledonia. Env. Sci. Technol. 2009, 43 (19), 7384–7390. Becquer, T. Chromium availability in ultramafic soils from New Caledonia. Sci. Total Environ. 2003, 301 (1-3), 251–261. Cooper, G. Oxidation and toxicity of chromium in ultramafic soils in Zimbabwe. Appl. Geochem. 2002, 17 (8), 981–986. Garnier, J.; Quantin, C.; Martins, E. S.; Becquer, T. Solid speciation and availability of chromium in ultramafic soils from Niquelândia, Brazil. J. Geochem. Explor. 2006, 88 (1-3), 206–209. Murray, K. J.; Tebo, B. M. Cr(III) Is Indirectly Oxidized by the Mn(II)-oxidizing bacterium Bacillus sp. strain SG-1. Env. Sci. Technol. 2007, 41 (2), 528–533. Landrot, G.; Ginder-Vogel, M.; Sparks, D. L. Kinetics of chromium (III) oxidation by manganese (IV) oxides using quick scanning X-ray absorption fine structure spectroscopy (Q-XAFS). Env. Sci. Technol. 2010, 44 (1), 143–149. Rajapaksha, A. U.; Vithanage, M.; Ok, Y. S.; Oze, C. Cr(VI) Formation related to Cr(III)-muscovite and birnessite interactions in ultramafic environments. Env. Sci. Technol. 2013, 47 (17), 9722–9729. Namgung, S.; Kwon, M. J.; Qafoku, N. P.; Lee, G. Cr(OH)3(s) Oxidation induced by surface catalyzed Mn(II) oxidation. Env. Sci. Technol. 2014, 48 (18), 10760–10768. Weaver, R. M.; Hochella, M. F. J. The reactivity of seven Mn-oxides with Cr3+aq: A comparative analysis of a complex, environmentally important redox reaction. Am. Mineral. 2003, 88 (11-12), 2016. Nico, P. S.; Zasoski, R. J. Importance of Mn(III) Availability on the rate of Cr(III) oxidation on δ-MnO2. Env. Sci. Technol. 2000, 34 (16), 3363–3367. Tang, Y.; Webb, S. M.; Estes, E. R.; Hansel, C. M. Chromium(III) oxidation by biogenic manganese oxides with varying structural ripening. Env. Sci. : Processes Impacts 2014, 16, 2127–2136. Eary, L. E.; Rai, D. Kinetics of chromium(III) oxidation to chromium(VI) by 25 ACS Paragon Plus Environment

Environmental Science & Technology

552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597

(28) (29)

(30)

(31) (32)

(33)

(34)

(35)

(36)

(37) (38)

(39)

(40)

(41)

(42)

reaction with manganese dioxide. Env. Sci. Technol. 1987, 21 (12), 1187–1193. Oze, C.; Sleep, N. H.; Coleman, R. G.; Fendorf, S. Anoxic oxidation of chromium. Geology 2016, G37844.1. Rock, M. L.; James, B. R.; Helz, G. R. Hydrogen peroxide effects on chromium oxidation state and solubility in four diverse, chromium-enriched soils. Env. Sci. Technol. 2001, 35 (20), 4054–4059. Fendorf, S.; Wielinga, B. W.; Hansel, C. M. Chromium transformations in natural environments: The role of biological and abiological processes in chromium(VI) reduction. Int. Geol. Rev. 2000, 42 (8), 691–701. Fendorf, S. E. Surface reactions of chromium in soils and waters. Geoderma 1995, 67 (1-2), 55–71. Gasser, U. G.; Juchler, S. J.; Sticher, H. Chemistry and speciation of soil water from serpentinitic soils: Importance of colloids in the transport of Cr, Fe, Mg, and Ni. Soil Sci. 1994, 158 (5), 314. Gough, L.; Meadows, G.; Jackson, L.; Dudka, S. Biogeochemistry of a highly serpentinized, chromite-rich ultramafic area, Tehama County, California; United States Geological Survey Bulletin 1901: Denver, CO; 1989; pubs.usgs.gov/bul/1901/report.pdf. Anderson, A. J.; Meyer, D. R.; Mayer, F. K. Heavy metal toxicities: levels of nickel, cobalt and chromium in the soil and plants associated with visual symptoms and variation in growth of an oat crop. Aust. J. Agric. Res. 1973, 24 (4), 557–571. Becquer, T.; Quantin, C.; Boudot, J. P. Toxic levels of metals in Ferralsols under natural vegetation and crops in New Caledonia. Eur. J. Soil Sci. 2010, 61 (6), 994–1004. Novák, M.; Chrastný, V.; Čadkova, E.; Farkaš, J.; Bullen, T. D.; Tylcer, J.; Szurmanova, Z.; Cron, M.; Prechova, E.; Curik, J.; et al. Common occurrence of a positive δ53Cr shift in central European waters contaminated by geogenic/industrial chromium relative to source values. Env. Sci. Technol. 2014, 48 (11), 6089–6096. Kruckeberg, A. R. California serpentines: flora, vegetation, geology, soils, and management problems; University of California Press, Ltd.: Berkeley, CA, 1984. Izbicki, J. A.; Wright, M. T.; Seymour, W. A.; McCleskey, R. B.; Fram, M. S.; Belitz, K.; Esser, B. K. Cr(VI) occurrence and geochemistry in water from public-supply wells in California. Appl. Geochem. 2015, 63, 203-217. Manning, A. H.; Mills, C. T.; Morrison, J. M.; Ball, L. B. Insights into controls on hexavalent chromium in groundwater provided by environmental tracers, Sacramento Valley, California, USA. Appl. Geochem. 2015, 62, 186–199. Gonzalez, A. R.; Ndung'u, K.; Flegal, A. Natural occurrence of hexavalent chromium in the Aromas Red Sands aquifer, California. Env. Sci. Technol. 2005, 39 (15), 5505–5511. Abu-Saba, K. E.; Flegal, A. R. Temporally variable freshwater sources of dissolved chromium to the San Francisco Bay Estuary. Env. Sci. Technol. 1997, 31 (12), 3455–3460. Morrison, J. M.; Goldhaber, M. B.; Lee, L.; Holloway, J. A. M.; Wanty, R. B.; Wolf, R. E.; Ranville, J. F. A regional-scale study of chromium and nickel in 26 ACS Paragon Plus Environment

Page 26 of 29

Page 27 of 29

598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643

Environmental Science & Technology

(43)

(44)

(45)

(46) (47)

(48) (49) (50)

(51)

(52)

(53)

(54)

(55) (56)

(57) (58)

soils of northern California, USA. Appl. Geochem. 2009, 24 (8), 1500–1511. Belitz, K.; Fram, M. S.; Johnson, T. D. Metrics for assessing the quality of groundwater used for public supply, CA, USA: Equivalent-population and area. Env. Sci. Technol. 2015, 49 (14) 8330-8338. SoilWeb; United States Department of Agriculture Natural Resources Conservation Service SSURGO Database; casoilresource.lawr.ucdavis.edu/gmap/. Hopson, C. A.; Mattinson, J. M.; Pessagno, E. A.; Luyendyk, B. P. California Coast Range ophiolite: Composite middle and late Jurassic oceanic lithosphere. In Ophiolites, Arcs, and Batholiths: A Tribute to Cliff Hopson; Wright, J. E., Shervais, J., Eds.; The Geological Society of America: Boulder, CO; 2008. Knoxville Creek California; www.wrcc.dri.edu/cgi-bin/rawMAIN.pl?caCKNO. Blake, G. R.; Hartge, K. H. Bulk Density. In Methods of Soil Analysis Part 1 Physical and Mineralogical Methods; Klute, A., Ed.; Soil Science Society of America: Madison, WI, 1986. Guelph Permeameter Kit; www.soilmoisture.com/GUELPH-PERMEAMETERKIT/. Loppert, R. L.; Inskeep, W. P. Iron. In Methods of Soil Analysis; Bigham, J. M., Ed.; Soil Science Society of America: Madison, WI, 1996; pp 639–664. Arar, E. J.; Pfaff, J. D.; Martin, T. D. Determination of dissolved hexavalent chromium in drinking water, groundwater and industrial wastewater effluents by ion chromatography; Method 218.6, Revision 3.3.; U.S. Environmental Protection Agency: Cincinnati, OH, 1994; https://www.epa.gov/sites/production/files/2015-08/documents/method_2186_rev_3-3_1994.pdf Krumbein, W. E.; Altmann, H. J. A new method for the detection and enumeration of manganese oxidizing and reducing microorganisms. Helgol. Wiss. Meeresunters. 1973, 25, 347–356. Webb, S. The MicroAnalysis Toolkit: X-ray fluorescence image processing software. AIP Conference Proceedings. 2011, 1365 (1), 196-199; www.samsxrays.com/#!smak/n44a5. Ravel, B.; Newville, M. ATHENA, ARTEMIS, HEPHAESTUS: Data analysis for X-ray absorption spectroscopy using IFEFFIT. J. Synchrotron Radiat. 2005, 12, 537–541. Stonestrom, D. A.; White, A. F.; C Akstin, K. Determining rates of chemical weathering in soils—solute transport versus profile evolution. J. Hydrol. 1998, 209 (1-4), 331–345. Van Genuchten, M. T. A closed-form equation for predicting the hydraulic conductivity of unsaturated soils. Soil Sci. Soc. Am. J. 1980, 44 (5), 892–898. Oze, C. Chromium geochemistry in serpentinized ultramafic rocks and serpentine soils from the Franciscan complex of California. Am, J. Sci. 2004, 304 (1), 67–101. Gasser, U. G..; Juchler, S.; Sticher, H. The fate of chromium and nickel in subalpine soils derived from serpentinite. Can. J. Soil Sci. 1995, 75 (2), 187-195. Gasser, U. G.; Dahlgren, R. A. Solid-Phase Speciation and Surface Association of Metals in Serpentinitic Soils. Soil Sci. 1994, 158 (6), 409-420. 27 ACS Paragon Plus Environment

Environmental Science & Technology

644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689

(59)

(60)

(61)

(62)

(63)

(64)

(65) (66)

(67)

(68)

(69)

(70) (71)

(72)

(73)

Traore, D.; Beauvais, A.; Chabaux, F.; Peiffert, C.; Parisot, J. C.; Ambrosi, J. P.; Colin, F. Chemical and physical transfers in an ultramafic rock weathering profile: Part 1. Supergene dissolution of Pt-bearing chromite. Am. Mineral. 2008, 93 (1), 22–30. Tebo, B. M.; Bargar, J. R.; Clement, B. G.; Dick, G. J.; Murray, K. J.; Parker, D.; Verity, R.; Webb, S. M. Biogenic manganese oxides: Properties and mechanisms of formation. Annu. Rev. Earth Planet. Sci. 2004, 32 (1), 287–328. Bargar, J. R.; Tebo, B. M.; Bergmann, U.; Webb, S. M.; Glatzel, P.; Chiu, V. Q.; Villalobos, M. Biotic and abiotic products of Mn(II) oxidation by spores of the marine Bacillus sp.strain SG-1. American Mineralogist 2005, 90 (1), 143–154. Duckworth, O. W.; Akafia, M. M.; Andrews, M. Y.; Bargar, J. R. Siderophorepromoted dissolution of chromium from hydroxide minerals. Env. Sci. : Processes Impacts 2014, 16 (6), 1348–1359. Rai, D.; Moore, D. A.; Hess, N. J.; Rosso, K. M.; Rao, L.; Heald, S. M. Chromium(III) hydroxide solubility in the aqueous K+-H+-OH−-CO2-HCO3− CO32−-H2O system: A thermodynamic model. J. Solution Chem. 2007, 36 (10), 1261–1285. Gabriel, C.; Raptopoulou, C. P.; Terzis, A.; Tangoulis, V.; Mateescu, C.; Salifoglou, A. pH-Specific synthesis and spectroscopic, structural, and magnetic studies of a Chromium(III)−citrate species. Aqueous solution speciation of the binary Chromium(III)−citrate system. Inorg. Chem. 2007, 46 (8), 2998–3009. Hamada, Y. Z.; Holyfield, H.; Rosli, K.; Burkey, T. Equilibrium models of Cr3+ and Cu2+ with glutamate. J Coord. Chem. 2009, 62 (5), 721–733. Zachara, J. M.; Girvin, D. C.; Schmidt, R. L.; Resch, C. T. Chromate adsorption on amorphous iron oxyhydroxide in the presence of major groundwater ions. Env. Sci. Technol. 1987, 21 (6), 589–594. Villalobos, M.; Trotz, M. A.; Leckie, J. O. Surface complexation modeling of carbonate effects on the adsorption of Cr(VI), Pb(II), and U(VI) on goethite. Env. Sci. Technol. 2001, 35 (19), 3849–3856. Mills, C.; Morrison, J. M.; Goldhaber, M. B.; Ellefsen, K. J. Chromium (VI) generation in vadose zone soils and alluvial sediments of the southwestern Sacramento Valley, California: A potential source of geogenic Cr (VI) to groundwater. Appl. Geochem. 2011, 26 (8), 1488–1501. Sherlock, M. D.; Chappell, N. A.; McDonnell, J. J. Effects of experimental uncertainty on the calculation of hillslope flow paths. Hydrol. Processes 2000, 14, 2457–2471. Chappell, N.; Ternan, L. Flow path dimensionality and hydrological modelling. Hydrol. Processes 1992, 6, 327–345. Kieber, R. J.; Willey, J. D.; Zvalaren, S. D. Chromium speciation in rainwater: Temporal variability and atmospheric deposition. Env. Sci. Technol. 2002, 36 (24), 5321–5327. Malmström, M. E.; Destouni, G.; Banwart, S. A.; Strömberg, B. H. E. Resolving the Scale-Dependence of Mineral Weathering Rates. Env. Sci. Technol. 2000, 34 (7), 1375–1378. White, A. F.; Brantley, S. L. The effect of time on the weathering of silicate minerals: why do weathering rates differ in the laboratory and field? Chem. Geol. 28 ACS Paragon Plus Environment

Page 28 of 29

Page 29 of 29

690 691 692 693 694 695 696 697 698 699 700 701

Environmental Science & Technology

(74)

(75) (76) (77)

2003, 202 (3-4), 479–506. Madison, A. S.; Tebo, B. M.; Mucci, A.; Sundby, B.; Luther, G. W. Abundant porewater Mn(III) is a major component of the sedimentary redox system. Science 2013, 341 (6148), 875–878. Di Baldassare, G.; Montanari, A. Uncertainty in river discharge observations: a quantitative analysis. Hydrol. Earth Syst. Sci. 2009, 13, 913–921. Maher, K.; Chamberlain, C. P. Hydrologic Regulation of Chemical Weathering and the Geologic Carbon Cycle. Science 2014, 343 (6178), 1502–1504. Unruh, J. R.; Moores, E. M. Quaternary blind thrusting in the southwestern Sacramento Valley, California. Tectonics 1992, 11 (2), 192–203.

29 ACS Paragon Plus Environment