Stabilization of Natural Organic Matter by Short-Range-Order Iron

Oct 26, 2016 - Division of Environmental Studies, Graduate School of Frontier Sciences, The University of Tokyo, 332 Building of Environmental Studies...
0 downloads 0 Views 644KB Size
Subscriber access provided by University of Idaho Library

Article

Stabilization of Natural Organic Matter by Short-Range-Order Fe Hydroxides Kai-Yue Chen, Tsan-Yao Chen, Ya-Ting Chan, ChingYun Cheng, Yu-Min Tzou, Yu-Ting Liu, and Heng-Yi Teah Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b02793 • Publication Date (Web): 26 Oct 2016 Downloaded from http://pubs.acs.org on November 1, 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 36

Environmental Science & Technology

1 2 3

Stabilization of Natural Organic Matter by Short-Range-Order Fe Hydroxides

4 5 6

Kai-Yue Chen,a Tsan-Yao Chen,b Ya-Ting Chan,a Ching-Yun Cheng,a Yu-Min Tzou,*,a Yu-Ting Liu,*,a and Heng-Yi Teahc

7 8 9

a

Department of Soil and Environmental Sciences, National Chung Hsing University, Taichung 40227, Taiwan, R.O.C.

10 11

b

12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27

c

Department of Engineering and System Sciences, National Tsing Hua University, Hsinchu 30043, Taiwan, R.O.C.

Division of Environmental Studies, Graduate School of Frontier Sciences, The University of Tokyo, 332 Building of Environmental Studies, 5-1-5 Kashiwanoha, Kashiwa City, Chiba 277-8563, Japan

Corresponding author's information: Yu-Ting Liu Department of Soil and Environmental Sciences, National Chung Hsing University, 145 Xingda Rd., Taichung 40227, Taiwan, R.O.C. Tel: +886-4-22840373 Ext. 3402 Fax: +886-4-22855167 E-mail address: [email protected] Yu-Min Tzou Department of Soil and Environmental Sciences,

28 29 30 31

National Chung Hsing University, 145 Xingda Rd., Taichung 40227, Taiwan, R.O.C. Tel: +886-4-22840373 Ext. 4206 Fax: +886-4-22855167

32

E-mail address: [email protected]

1

ACS Paragon Plus Environment

Environmental Science & Technology

33

Abstract

34

Dissolved organic matter (DOM) is capable of modifying the surfaces of soil

35

minerals (e.g., Fe hydroxides) or even forming stable coprecipitates with Fe(III) in a

36

neutral environment. The DOM/Fe coprecipitation may alter biogeochemical carbon

37

cycling in soils if the relatively mobile DOM is sorbed by soil minerals against

38

leaching, runoff, and biodegradation. In this study, we aimed to determine the

39

structural development of DOM/Fe coprecipitates in relation to changes in pH and

40

C/(C+Fe) ratios using XRD, XPS, Fe K-edge XAS, FTIR, and C-NEXAFS

41

techniques. The results showed that in the system with bulk C/(C+Fe) molar ratios ≤

42

0.65, the ferrihydrite-like Fe domains were precipitated as the core and covered by the

43

C shells. When the C/(C+Fe) molar ratio ranged between 0.71 and 0.89, the emerging

44

Fe-C bonding suggested a more substantial association between Fe domains including

45

edge- and corner-sharing FeO6 octahedra and DOM. With C/(C+Fe) bulk molar ratios

46

≥ 0.92, only corner-sharing FeO6 octahedra along with Fe-C bonding was found. The

47

homogeneously distributed C and Fe domains caused the enhancement of Fe and C

48

solubilisation from coprecipitates. The C/(C+Fe) ratios dominated structural

49

compositions and stabilities of C/Fe coprecipitates and may directly affect the Fe and

50

C cycles in soils.

51 2

ACS Paragon Plus Environment

Page 2 of 36

Page 3 of 36

Environmental Science & Technology

52

Introduction

53

The total carbon (C) present in terrestrial ecosystems is approximately 3170 GT,

54

and it is estimated that nearly 80 % of the inventory, including 1550 GT of organic

55

and 950 GT of inorganic carbon, is preserved in soils.

56

major source of soil organic carbon, consists of non-humic substances such as

57

biochemical materials3 and dissolved organic matter (DOM), and humic substances

58

such as humic acid, fulvic acid, and humin.4 Generally, non-humic substances are

59

more biodegradable than humic substances. Any changes in the turnover rates of soil

60

carbon pools potentially alters the atmospheric CO2 concentration as a result of the

61

significant correlation between the global carbon cycle and the soil carbon pool.

62

Therefore, stabilizing SOM and preventing its physicochemical and/or biological

63

mineralization may partially alleviate global warming caused by an elevation of the

64

carbon content in the atmosphere. Soil minerals play a relatively important and key

65

role in preserving SOM.5

1,2

Soil organic matter (SOM), a

66

The SOM could be stabilized by forming humic substances, black carbon-like

67

materials, and aggregates with soil minerals.6-16 By means of cation bridges, SOM/

68

DOM could complex/precipitate with metal cations, particularly the polyvalent metals,

69

and subsequently stabilize C against bio-degradations and transportations.5, 15, 17-20

70

Complexation or precipitation between trivalent Fe/Al cations and SOM/DOM has 3

ACS Paragon Plus Environment

Environmental Science & Technology

71

been commonly found in the leached soils, mine drainage waters, and aquifers.21-29

72

Along with the fluctuation of redox conditions, organic-metal precipitates could also

73

form via a strong covalent bond between DOM and Fe(III),27 which was derived from

74

the oxidation of soluble Fe(II) as a result of an intrusion of dissolved oxygen.30

75

Although partition/adsorption of organic molecules may occur on the Fe hydroxides,

76

carboxyl groups that dominate DOM structures provide abundant binding sites to

77

coprecipitate with Fe(III).31-33 Because coprecipitation showed a relatively higher C

78

stabilization than partition/adsorption,31,

79

coprecipitation between DOM and Fe was an important factor contributing to soil C

80

storages.34

33

previous studies evidenced that

81

The pH values and C/(C+Fe) molar ratios affected both the structures and the

82

electro-affinity of DOM and Fe-containing minerals during the coprecipitation

83

processes of Fe and DOM.29, 35-37 For instance, the positive charge density on the

84

surfaces of Fe hydroxides varied along with changes in the solution pH and

85

essentially affected bonding abilities of Fe cations with DOM.35 While being

86

dissociated, O-containing functional groups including carboxyl and phenolic groups

87

of DOM provide abundant binding sites for Fe.38,

88

precipitation between DOM and Fe was favorable due to an inhibition of the

89

formation of Fe hydroxides.27 In terms of C/(C+Fe) ratios, previous studies indicated

39

4

ACS Paragon Plus Environment

Under acidic conditions,

Page 4 of 36

Page 5 of 36

Environmental Science & Technology

90

that the C/Fe coprecipitates could only be produced in an acidic environment when

91

the C/metal molar ratios ranged from 10 to 33.3.35,

92

examined the productive conditions of the metal-organic coprecipitates, the mechanic

93

mechanisms of structural development at the molecular scale and stabilizations of

94

coprecipitates as influenced by pH and C/(C+Fe) ratios during the coprecipitation

95

processes remain unclear.

96

Understanding

the

structures

and

40

Even if these studies had

stabilization/persistence

of

DOM/Fe

97

coprecipitates is an essential component of evaluating their contributions to carbon

98

conservation in terrestrial systems. Structural and surface properties of DOM/Fe

99

coprecipitates have been previously determined using several methods such as X-ray

100

diffraction (XRD),41 Fourier-transform infrared spectroscopy (FTIR),31 and X-ray

101

photoelectron spectroscopy (XPS).33 Nonetheless, the effects of Fe proportion on the

102

molecular-level structural development and changes of DOM/Fe coprecipitates should

103

also be investigated to determine changes in carbon stabilization as the function of

104

compositions for DOM/Fe coprecipitates.

105

Lately, synchrotron-based techniques such as X-ray absorption spectroscopy (XAS),

106

including X-ray absorption near edge structure (XANES) and extended X-ray

107

absorption fine structure (EXAFS) spectroscopy, have received significant scientific

108

attention because of their capability to determine the local coordination environments 5

ACS Paragon Plus Environment

Environmental Science & Technology

109

on the molecular scale. For example, while using synchrotron-based techniques,

110

Karlsson et al.42 indicated that a mononuclear Fe/C bond was formed in the Fe/OM

111

complex with a carboxyl/Fe ratio ≥ 22.5. However, mixtures of mononuclear Fe/C

112

bonding and Fe hydroxides were found with a carboxyl/Fe ratio < 22.5. Previous

113

studies focused mainly on the effects of organic molecules on the crystalline Fe

114

hydroxides and the changes in the lengths and coordination numbers of the Fe-Fe

115

bond with increasing C dose.42 However, the local structural coordination including

116

Fe-C and Fe-Fe bonds in DOM/Fe coprecipitates and whether such structural

117

attributes affect C stabilization as functions of C/(C+Fe) ratios in an acidic solution

118

still remained unexplored. In this study, we aimed to determine the structural

119

stabilization of DOM/Fe coprecipitates in relation to changes in pH and C/(C+Fe)

120

ratios using Fe K-edge XAS, C (1s) near edge X-ray absorption fine structure

121

(C-NEXAFS), and other conventional spectroscopic techniques. The C-stabilization

122

and Fe-C bonds associated with structural changes of Fe-DOM coprecipitates due to

123

Fe intrusion were systematically examined.

124 125

Materials and Methods

126

Extraction of DOM. The DOM was extracted from the Changhua (CHA) peat soil

127

collected from central Taiwan. The vegetation type is weed. Briefly, the CHA soil was 6

ACS Paragon Plus Environment

Page 6 of 36

Page 7 of 36

Environmental Science & Technology

128

mixed with de-ionized (DI) water in a reaction vessel with a solid to liquid ratio of 20

129

g L-1. After shaking for 96 h, the suspension was passed through a 0.45 µm membrane

130

filter,16 and the filtrate that contained the DOM was used for coprecipitation

131

experiments. In total, 200 mg L-1 of NaN3 was added into the DOM solutions to

132

inhibit microbe growth. The C concentrations in the DOM were measured by a total

133

organic carbon (TOC) analyzer (multi N/C 2100, analytikjena s). The DOM was also

134

determined by an elemental analyzer (Heraeus CHN-O-S Rapid Analyzer), NMR

135

(nuclear magnetic resonance, Bruker DSX400WB NMR), and FTIR (Thermo Nicolet

136

Nexus).

137

,

Coprecipitation

and

solubilisation

of

DOM/Fe(III).

The

DOM/Fe(III)

138

coprecipitates (DFC) were synthesized by adding 0.5, 1, 2.5, or 5 mM Fe(III) to a

139

DOM solution containing 100 mg L-1 of DOC (8.33 mM C). The C/(C+Fe) molar

140

ratios of initial addition were 0.62, 0.77, 0.89, and 0.94, which were denoted as

141

DFC62, DFC77, DFC89 and DFC94, respectively. During the coprecipitation, the

142

suspension pH was maintained at 3, 4.5, or 6 individually using 0.1 M KOH/HCl with

143

constant stirring at 150 rpm for 12 h. Hereafter, the coprecipitation pH is denoted in

144

front of the abbreviation of DFC. After 12 h, suspensions were centrifuged at 7155 g

145

for 20 min, and concentrations of C and Fe(III) in the supernatants were measured by

146

the TOC analyzer and inductively coupled plasma - atomic emission spectrometer 7

ACS Paragon Plus Environment

Environmental Science & Technology

147

(ICP-AES, Spectro Genesis), respectively. The coprecipitates were washed 3-5 times

148

using DI water and then freeze-dried for elemental and structural analyses. The

149

solubilisation of C and Fe from DFC samples was examined at a solid concentration

150

of 60 mg C L-1 at the same pH used in the coprecipitation. The concentration of 60 mg

151

C L-1 was chosen as the naturally dissolved OM from the peat soil generally ranges

152

from 20 to 60 mg C L-1.43 After 24 h, the suspensions were centrifuged at 7155 g for

153

20 min and then passed through 0.45 µm filter membranes. The concentrations of C

154

and Fe in the filtrate were determined using TOC and ICP-AES.

155

Characterizations of DOM/Fe coprecipitates. A subsample of each coprecipitates

156

was washed with deionized water to remove excess salts, and freeze-dried or

157

preserved as the moist paste prior to analyses of elemental analyzer (EA) and

158

spectroscopic techniques, including FTIR, nuclear magnetic resonance (NMR), XRD,

159

XPS, Fe-XAS, and C-NEXAFS [see “Chemical Compositions, Mineralogical,

160

Structural, and Surface Analyses of DOM/Fe Coprecipitates” in the Supporting

161

Information (SI) for details]. Briefly, EA was used to determine the contents for C, H,

162

N, and O in coprecipitates. The functional groups on DOM and DFC samples were

163

determined using FTIR, NMR, and C-NEXAFS analyses. Elemental distribution on

164

near surfaces of DFC samples was determined using XPS analysis. The XRD patterns

165

of DFC samples were collected to determine the mineralogical attributes. The Fe 8

ACS Paragon Plus Environment

Page 8 of 36

Page 9 of 36

Environmental Science & Technology

166

K-edge XAS Spectroscopy including XANES and EXAFS was used to determine the

167

Fe species and local structures surrounding central Fe atoms, respectively. The

168

synchrotron based analyses of C-NEXAFS and Fe-XAS were conducted at the

169

National Synchrotron Radiation Research Center (NSRRC), Taiwan.

170 171

Results

172

Characterizations of DOM. The results of the elemental analysis indicated that

173

the C content in the DOM was approximately 38 % (Table S1 in the supporting

174

information, SI). The C13 NMR result revealed that the DOM contains 29.4 % alkyl,

175

36.7 % O-alkyl, 14.0 % aromatic, 5.0 % phenolic, 12.6 % carboxyl, and 2.2 %

176

carbonyl functional groups (Figure S1 in SI). The FTIR spectrum of DOM showed –

177

CH2 and –CH3 stretching in the aliphatic region at 2918 and 2849 cm–1, symmetric

178

C=O stretch of esters and asymmetric –C=O stretching derived from the carboxyl and

179

carbonyl groups at 1730 and 1610 cm-1, the symmetric COO- band of the carboxyl

180

group at 1400 cm-1, O–H stretching of the phenolic OH and/or OH deformation of

181

COOH at 1270 cm-1, and the C–O stretch of carbohydrate group at 1050 cm-1 (Figure

182

S2 in SI).29, 38, 44-48 Both the FTIR and NMR results indicated that the DOM consists

183

mainly of aliphatic carbon enriched with O-containing groups.

184

Precipitation, solubilisation, crystallization, and distribution of C and Fe in 9

ACS Paragon Plus Environment

Environmental Science & Technology

185

DFC samples. At each precipitation pH, the amounts of precipitated C in the DFC

186

samples tended to increase with increasing C/(C+Fe) ratios of initial addition,

187

whereas the precipitated Fe showed an opposite trend (Table 1 and Figure S3). With

188

the exception of the 3DFC62 and 3DFC77 samples, the final (bulk) C/(C+Fe) molar

189

ratios in the DFC samples were very close to the C/(C+Fe) ratios of their initial

190

counterparts (Table 1). The DFC samples precipitated at pH 3 seemed to have

191

relatively higher C/(C+Fe) ratios, which could be attributed to the flocculation of C

192

domains that facilitated the formation of C/Fe coprecipitates at acidic conditions.49, 50

193

However, such C enrichment at pH 3 was not found when the initial C/(C+Fe) ratios

194

were greater than 0.77.37 On the other hand, the relatively lower C/(C+Fe) ratios were

195

observed for the DFC samples prepared at pH 6 than those prepared at pH 3 and pH

196

4.5 (Table 1). Such results were likely due to the occurrence of Fe hydrolysis, forming

197

Fe hydroxides at a higher pH, that might hinder the association of C and Fe(III).

198

The stabilization of the DFC samples was determined based on the degree of C

199

and Fe solubilisation upon incubation with 0.01 M KCl for 12 h at the same pH in

200

which the DFC was synthesized during the precipitation process. In general, the

201

solubilisation of both C and Fe was more pronounced for the DFC samples prepared

202

at pH 6 (Table 1). For instance, up to 24.9 and 8.5 % of C and Fe, respectively, were

203

dissolved from the 6DFC94 sample and the released C and Fe were 3.5 and 2.2 times 10

ACS Paragon Plus Environment

Page 10 of 36

Page 11 of 36

Environmental Science & Technology

204

greater than that of the 3DFC94 sample (Table 1). In addition, the solubilisation of C

205

and Fe tended to decrease with decreasing C/(C+Fe) molar ratios (Table 1), indicating

206

that the DFC samples with relatively lower C/(C+Fe) ratios were more stable. The

207

significant variations in the amounts of C and Fe solubilisation suggested the presence

208

of structural and compositional differences among these DFC samples.

209

The X-ray diffraction patterns demonstrated a variation in the degree of

210

crystallinity for the DFC samples (Figure 1). For the sample precipitated at pH 3 with

211

a C/(C+Fe) molar ratio of 0.71 (i.e., 3DFC62), two broad peaks centered at 2.6 and

212

1.5 Å indicated the presence of two-line ferrihydrite.51 With increasing precipitation

213

pH, these two peaks could be observed in the samples with a relatively higher

214

C/(C+Fe) ratio. For instance, the ferrihydrite structure could be observed in the

215

sample with a C/(C+Fe) ratio up to 0.89 at pH 6 (6DFC89) (Figure 1). That is, in a

216

system with a higher proportion of impurities, e.g., C, the poorly crystalline Fe

217

hydroxides were readily formed at relatively alkaline conditions.52

218

Given that the surface compositions might affect the physicochemical properties

219

of the DFC, we used the XPS technique to determine distributions of C and Fe on

220

near surfaces of the individual DFC samples. Figure 2 displays the C/(C+Fe) molar

221

ratio on near surfaces as a function of that of the bulk precipitate. Regardless of the

222

bulk C/(C+Fe) molar ratios, C instead of Fe seemed to dominate near surfaces of each 11

ACS Paragon Plus Environment

Environmental Science & Technology

223

synthesized DFC sample (Figure 2). When bulk C/(C+Fe) molar ratios increased from

224

0.62 to 0.95, the ratios on near surfaces were essentially greater than 0.94 for all DFC

225

particles (Figure 2). That is, particularly at relatively lower bulk C/(C+Fe) ratios, C

226

tended to precipitate on particle surfaces. On the contrary, a nearly 1:1 relationship

227

between bulk and surface C/(C+Fe) ratios was found when the bulk C/(C+Fe) ratio

228

was > 0.92. Such results implied a relatively more homogeneous distribution of C and

229

Fe in the DFC samples with a greater C/(C+Fe) molar ratio. The significant

230

discrepancies in the correlation between bulk and surface C/(C+Fe) ratios for

231

individual DFC samples suggested the essential changes in the structural association

232

between C and Fe at different C/(C+Fe) ratios.

233

Speciation and local structures of Fe in DFC samples. The Fe speciation in

234

the DFC samples was determined using linear combination fitting (LCF) of Fe K-edge

235

XANES spectra.53 The LCF results presented in Table 2 revealed that the Fe species

236

in all DFC samples could be fit as a combination of 2-line ferrihydrite and Fe(III)

237

complexed with citrate.54 Figure S4 in the SI contains examples of the LCF fitting for

238

the DFC samples coprecipitated at pH 3. While ferrihydrite dominated the Fe species

239

in the DFC sample with a bulk C/(C+Fe) molar ratio ≤ 0.65, the proportion of Fe(III)

240

complexed with organic groups increased as the C/(C+Fe) molar ratio increased. For

241

instance, the percentages of Fe(III) bound to organic groups in the DFC samples 12

ACS Paragon Plus Environment

Page 12 of 36

Page 13 of 36

Environmental Science & Technology

242

increased from 18.7 to 96.3 when the bulk C/(C+Fe) molar ratios increased from 0.71

243

to 0.95 at pH 3. These results were in line with the previous study31, which reported a

244

mixture of 2-line ferrihydrite and insoluble Fe(III)-organic complexes with a C/(C+Fe)

245

molar ratio ≥ 0.74.

246

We subsequently used Fe-EXAFS analysis to investigate the local structures

247

surrounding the central Fe atom in the DFC samples. The Fe K-edge EXAFS spectra

248

in Figure S5a in the SI showed that oscillation amplitudes at k ≈ 5.0 and 7.5 Å-1

249

generally decreased with an increasing proportion of C, corresponding to a diminish

250

in high-shell backscattering signals between 2.3 and 3.4 Å in the Fourier transformed

251

(FT) spectra (Figure S5b in the SI).51 However, oscillation amplitudes at 6.3 and 8.5

252

Å-1 showed the opposite trend, suggesting high C proportions affected the

253

wave-beating patterns of ferrihydrite featured structures and the Fe coordination

254

environment.55 The EXAFS data were fit with an EXAFS model derived from

255

ferrihydrite56 that generated the single-scattering Fe-O and Fe-Fe paths out to 3.45 Å

256

and Fe(III)-NOM complexes57 that generated the single-scattering Fe-C paths in the

257

range 2.82-2.98 Å (Table 3). Coordination numbers of the first-shell Fe-O paths that

258

were fit at 1.96-1.99 Å across all DFC samples were fixed at six (Table 3) because

259

each Fe atom was expected to be surrounded by six oxygen atoms under the moist

260

paste conditions of the prepared EXAFS samples, and the surface defects of Fe 13

ACS Paragon Plus Environment

Environmental Science & Technology

261

hydroxides were presumed to be covered by water molecules.58 The σ2 values of the

262

Fe-C path as well as the edge- and corner-sharing Fe-Fe paths for DFC samples were

263

floated according to the precipitated pH due to the structural ordering difference.51, 59

264

With the exception of the DFC62 samples precipitated at pH 4.5 and 6, whose

265

C/(C+Fe) ratios were ≤ 0.65, the second-shell coordination surrounding the central

266

Fe atom for the other DFC samples were designated as the Fe-C bonds with an

267

average distance of 2.97 Å, and there were no significant differences in the

268

coordination number (CN) for these Fe-C bonds (Table 3). For the Fe-Fe paths, we

269

found two average distances at 3.04 (3.01~3.07) and 3.42 (3.40~3.45) Å,

270

corresponding to edge- and corner-shared FeO6 octahedral linkages, respectively.60, 61

271

While the corner-shared (CS) FeO6 octahedra were found across all DFC samples, the

272

edge-shared (ES) FeO6 octahedra might only occur in samples with a C/(C+Fe) molar

273

ratio ≤ 0.89. The CN for ES and CS Fe-Fe paths varied in the range of 0.82-1.77 and

274

0.39-1.87, respectively, and no regular changes were found along with either

275

coprecipitation pH or C/(C+Fe) molar ratios (Table 3). The CN for both ES and CS

276

FeO6 octahedra in our DFC samples was essentially less than that of two-line

277

ferrihydrite.41 Based on a structural model proposed elsewhere,61, 62 the decrease in Fe

278

octahedral linkages could be translated as a decrease in the domain size of Fe

279

polyhedra. That is, the Fe domain in our DFC samples is possibly smaller than ∼2 nm, 14

ACS Paragon Plus Environment

Page 14 of 36

Page 15 of 36

Environmental Science & Technology

280

which is the general domain size of two-line ferrihydrite.56 According to the EXAFS

281

results, the local coordination of Fe in the DFC samples could be plausibly grouped

282

into the following three types: (1) the ferrihydrite-like Fe domain with C/(C+Fe) ≤

283

0.65; (2) the mixture of ferrihydrite-like Fe domains or ES/CS FeO6 octahedra with

284

Fe-C bonds when C/(C+Fe) ranged between 0.71 and 0.89; and (3) the CS FeO6

285

octahedra associated with Fe-C bonds when C/(C+Fe) ≥ 0.92. The vicissitude of the

286

ferrihydrite-like structure as a function of C/(C+Fe) molar ratios in DFC samples

287

obtained from EXAFS analyses was consistent with LCF and XRD results.

288

Changes of organic functional groups in DFC samples. The changes in

289

organic functional groups in DFC samples were first examined using FTIR analysis

290

(Figure 3). Generally, each major absorption peak of the DOM became broad or even

291

disappeared with decreasing C/(C+Fe) molar ratios. Disappearance of the symmetric

292

–C=O stretch at 1730 and 1400 cm-1 for DFC62 and DFC77 samples at all pH values

293

indicated that the carboxyl groups of the DOM might strongly associate with Fe

294

during coprecipitation processes. However, the association seemed incomplete for the

295

DFC89 and DFC94 samples at all pH values, probably due to the presence of a higher

296

number of carboxyl groups in these samples. Upon Fe complexation, the

297

characteristic absorption peak of the carboxyl group (–C=O band) was shifted to a

298

lower frequency, broadening the asymmetric –C–O stretch of the carbonyl group in 15

ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 36

299

the region from 1610 to 1645 cm-1.16 In addition, the O–H stretching vibration of the

300

phenolic OH at 1270 cm-1 disappeared across all DFC samples with the addition of Fe,

301

suggesting the phenolic groups might also participate in the coprecipitation reactions.

302

The variation in the C–O stretching of carbohydrates at 1050 cm-1 was relatively

303

insignificant with increasing C/(C+Fe) ratios, demonstrating low affinity between this

304

functional group and Fe(III).31, 39

305

The C NEXAFS analysis was used to further examine the changes in C

306

functional groups of the DFC samples (Figure 4). Three major peaks of carbon

307

materials in the DFC samples indicated the bands of H and C-substituted aromatic

308

carbon (π*C=C, 285.1 ± 0.2 eV), O-substituted aromatic carbon (phenolic carbon, π

309

*C=C-O, 286.5 ± 0.2 eV), and carboxyl carbon (π*C=O, 288.5 ± 0.3 eV).31,

310

Compared with the original DOM, the carboxyl C peak and aromatic C peak for DFC

311

samples became broader and less intense, suggesting the interaction between Fe(III)

312

and these two groups was significant in all DFC samples.38, 39 The phenolic carbon

313

disappeared for all of the 6DFC samples due to their strong associations with Fe(III)

314

at pH 6.65 A slight decrease in the intensity of the phenolic peak was observed for the

315

3DFC and 4.5DFC samples, which may be the collective result of low affinity or

316

lesser amounts of bonding between Fe(III) and the phenolic groups and flocculation

317

of some unreacted phenolic groups under acidic conditions, i.e., pH 3 and 4.5. 16

ACS Paragon Plus Environment

63, 64

Page 17 of 36

Environmental Science & Technology

318

However, subtle changes in the phenolic groups upon the coprecipitation of DOM and

319

Fe at various pH values were not observed in the FTIR spectra with a lower energy.

320

Accordingly, the current results were consistent with the previous studies in that the

321

carboxyl groups and aromatic moieties of DOM exhibited various degrees of high

322

affinity with Fe(III).31, 32

323 324

Discussion

325

Interactions with Fe(III) control the stabilization and biogeochemical cycling of

326

C. A significant decrease in Fe(III) solubilisation was observed with increasing Fe

327

proportion in the DFC samples among all tested pH values (Table 1). We presumed

328

that the Fe domains became unstable due to the loss of edge-sharing Fe octahedra

329

with an increase in the C proportion, leading to an elevation of Fe solubilisation

330

(Table 1). According to Fe K-edge XAS results contained in the supplementary

331

information provided by other spectroscopic analyses, structural developments of

332

DOM/Fe(III) coprecipitates could be generally categorized into three types as a

333

function of the C/(C+Fe) ratio.

334

In the system with C/(C+Fe) molar ratios ≤ 0.65, XRD and Fe-XAS analyses

335

(Figure 1, Table 2 and 3) implied that such DFC samples precipitated as a core-shell

336

structure. The ferrihydrite-like Fe domains precipitated as the cores, and these 17

ACS Paragon Plus Environment

Environmental Science & Technology

337

Fe-cores were covered with DOM molecules as indicated by XPS analysis, which

338

showed an approximate 1.5-fold enrichment of near-surface C proportion relative to

339

bulk C proportion (Figure 2, Table S2). The association between the Fe cores and the

340

C shells was evidenced by the significantly decreased peak intensity of the carboxyl

341

functional groups as shown in the FTIR and C-NEXAFS results (Figure 3 and 4).

342

These samples also had the least amounts of C and Fe solubilisation (Table 1),

343

suggesting a relatively stable structure in the Fe-core/C-shell type.

344

When the C/(C+Fe) molar ratio ranges between 0.71 and 0.89, the formation of

345

Fe-C bonds suggested a more substantial association between Fe domains including

346

ES and CS FeO6 octahedra and DOM (Table 3). In accordance with the variation in

347

the local structure of the Fe domains, the LCF results of Fe-XANES spectra showed

348

the proportion of ferrihydrite in the Fe inventory decreased concomitantly with the

349

increasing proportion of Fe(III) complexed with organic functional groups (Table 2).

350

Moreover, the less discernible ferrihydrite patterns in the XRD analyses for samples

351

with C/(C+Fe) molar ratios from 0.71 to 0.89 also attested to the decreasing

352

proportion of ferrihydrite in the total Fe inventory and/or the more disrupted structural

353

ordering of ferrihydrite (Figure 1).31, 41

354

In DFC samples with C/(C+Fe) molar ratios ≥ 0.92, only CS FeO6 octahedra

355

along with Fe-C bonding was found in coprecipitates (Table 3). Such non-crystalline 18

ACS Paragon Plus Environment

Page 18 of 36

Page 19 of 36

Environmental Science & Technology

356

Fe hydroxides were also evidenced by the loss of diagnostic ferrihydrite peaks

357

observed in XRD analysis (Table 3). Due to the inherent speciation uncertainty caused

358

by the fitting models that can only approximate reality and/or the limits in the

359

separation of spectral features in the LCF of XANES spectra,66 our LCF results still

360

indicate ferrihydrite in the DFC samples with C/(C+Fe) molar ratios ≥ 0.92 although

361

the proportions were essentially less than those in other DFC samples. Moreover, the

362

loss of ES Fe octahedral linkages in these DFC samples compared with other DFC

363

samples implied a decrease in the Fe domain size when the C/(C+Fe) molar ratio was

364

≥ 0.92. In contrast to DFC samples with limited C [C/(C+Fe) molar ratios ≤ 0.89],

365

wherein Fe hydroxides served as a heterogeneous nucleation template for surface

366

outgrowth of DOM molecules, Fe and C in samples with C/(C+Fe) molar ratios ≥

367

0.92 tended to be distributed homogeneously on the external portions of the DFC

368

structures, resulting in the nearly 1:1 relationship between bulk and surface C/(C+Fe)

369

ratios (Figure 2). Due to the associations between Fe and C, interferences of the

370

growths for C and Fe domains caused unstable structures for Fe hydroxides and

371

subsequently enhanced the Fe solubilisation (Table 1). Such Fe solubilisation might

372

result in the concurrent release of C from those Fe sites with bonded C.

373 374

Implications 19

ACS Paragon Plus Environment

Environmental Science & Technology

375

Coprecipitation of NOM and Fe is an important process that stabilizes C and Fe

376

in the soil system. Many studies also indicated that the variation in the coprecipitate

377

structures is effected by the C/(C+Fe) ratios. Our collective results demonstrated the

378

C/(C+Fe) ratio-driven transformation of DFC coprecipitated structures. While the

379

outgrowth of the DOM on Fe hydroxides at C/(C+Fe) ratios between 0.71 and 0.89

380

stabilized the Fe domains, Fe hydroxides were subject to solubilisation at C/(C+Fe)

381

ratios ≥ 0.92 as a consequence of the homogeneously distributed unstable structures

382

between Fe and C domains. Once C/Fe coprecipitates formed in soil systems, such

383

association of organic matter to soil inorganic minerals increases the resistance of

384

organic matter to microbial decomposition In addition, C/Fe coprecipitates also serve

385

as carriers controlling the mobility and bioavailability of environmental inorganic

386

pollutants such as chromium and arsenic. Recognition of the structural composition

387

and stability of C/Fe coprecipitates could lead to a better quantification of the

388

dynamics and mass balances of Fe and C in soils, resulting in an improved prediction

389

of the geochemistry of nutrients such as phosphorus and contaminants such as arsenic

390

and heavy metals.

391 392 393

Acknowledgements The authors are grateful to Dr. Jyh-Fu Li for the assistance at beamline 17C and 20

ACS Paragon Plus Environment

Page 20 of 36

Page 21 of 36

Environmental Science & Technology

394

Dr. Jin-Ming Chen for the assistance at beamline 20A, NSRRC. This work is

395

financially supported by the Ministry of Science and Technology, ROC under the

396

project

397

103-2112-M-007-022-MY3, and 104-2311-B-005-016-MY3.

number

of

101-2313-B-005-047-MY3,

102-2313-B-029-005-MY2,

398 399

Supporting information

400

Additional information includes elemental composition, NMR, and FTIR results

401

of original CHA DOM, chemical compositions, mineralogical, structural, and surface

402

analyses of DOM/Fe Coprecipitates, as well as XAS analysis including XANES-LCF

403

results and EXAFS fitting results. This information is available free of charge via the

404

Internet at http://pubs.acs.org.

405 406 407 408 409 410 411 412 21

ACS Paragon Plus Environment

Environmental Science & Technology

413

References

414

1.

415 416 417

security. Science 2004, 304, (5677), 1623-1627. 2. Lal, R., Carbon sequestration. Philosophical Transactions of the Royal Society B-Biological Sciences 2008, 363, (1492), 815-830.

418 419 420

3. Hayes, M.; Swift, R., The chemistry of soil organic colloids. The chemistry of soil constituents 1978, 179-320. 4. Aiken, G. R.; McKnight, D. M.; Wershaw, R. L.; MacCarthy, P., Humic substances

421 422 423

in soil, sediment, and water: geochemistry, isolation and characterization. John Wiley & Sons: 1985. 5. Kaiser, K.; Guggenberger, G., The role of DOM sorption to mineral surfaces in the

424 425 426 427 428

preservation of organic matter in soils. Organic Geochemistry 2000, 31, (7-8), 711-725. 6. Edwards, A. P.; Bremner, J., MICROAGGREGATES IN SOILS1. Journal of Soil Science 1967, 18, (1), 64-73. 7. Elliott, E., Aggregate structure and carbon, nitrogen, and phosphorus in native

429 430 431

and cultivated soils. Soil Sci. Soc. Am. J. 1986, 50, (3), 627-633. 8. Jastrow, J. D., Soil aggregate formation and the accrual of particulate and mineral-associated organic matter. Soil Biology & Biochemistry 1996, 28, (4-5),

432 433 434

665-676. 9. Tisdall, J.; Oades, J. M., Organic matter and water-stable aggregates in soils. J. Soil Sci. 1982, 33, (2), 141-163.

435 436 437 438 439

10. Six, J.; Elliott, E. T.; Paustian, K., Soil macroaggregate turnover and microaggregate formation: a mechanism for C sequestration under no-tillage agriculture. Soil Biology & Biochemistry 2000, 32, (14), 2099-2103. 11. Six, J.; Paustian, K.; Elliott, E. T.; Combrink, C., Soil structure and organic matter: I. Distribution of aggregate-size classes and aggregate-associated carbon. Soil Science

440 441 442

Society of America Journal 2000, 64, (2), 681-689. 12. Swift, M. J.; Heal, O. W.; Anderson, J. M., Decomposition in terrestrial ecosystems. Univ of California Press: 1979; Vol. 5.

443 444 445 446 447 448 449 450

13. Zech, W.; Senesi, N.; Guggenberger, G.; Kaiser, K.; Lehmann, J.; Miano, T. M.; Miltner, A.; Schroth, G., Factors controlling humification and mineralization of soil organic matter in the tropics. Geoderma 1997, 79, (1-4), 117-161. 14. Hammes, K.; Smernik, R. J.; Skjemstad, J. O.; Schmidt, M. W. I., Characterisation and evaluation of reference materials for black carbon analysis using elemental composition, colour, BET surface area and C-13 NMR spectroscopy. Applied Geochemistry 2008, 23, (8), 2113-2122. 15. Six, J.; Conant, R. T.; Paul, E. A.; Paustian, K., Stabilization mechanisms of soil

Lal, R., Soil carbon sequestration impacts on global climate change and food

22

ACS Paragon Plus Environment

Page 22 of 36

Page 23 of 36

Environmental Science & Technology

451

organic matter: Implications for C-saturation of soils. Plant and Soil 2002, 241, (2),

452

155-176.

453 454 455

16. Chen, K. Y.; Liu, J. C.; Chiang, P. N.; Wang, S. L.; Kuan, W. H.; Tzou, Y. M.; Deng, Y.; Tseng, K. J.; Chen, C. C.; Wang, M. K., Chromate removal as influenced by the structural changes of soil components upon carbonization at different temperatures.

456 457 458

Environmental Pollution 2012, 162, 151-158. 17. Feller, C.; Beare, M. H., Physical control of soil organic matter dynamics in the tropics. Geoderma 1997, 79, (1-4), 69-116.

459 460 461

18. Ladd, J.; Foster, R.; Skjemstad, J., Soil structure: carbon and nitrogen metabolism. Geoderma 1993, 56, (1), 401-434. 19. Sorensen, L. H., Stabilization of newly formed amino acid metabolites in soil by

462 463 464 465 466

clay minerals. Soil Sci. 1972, 114, 5-11. 20. Torn, M. S.; Trumbore, S. E.; Chadwick, O. A.; Vitousek, P. M.; Hendricks, D. M., Mineral control of soil organic carbon storage and turnover. Nature 1997, 389, (6647), 170-173. 21. Bloomfield, C., A study of podzolization. Journal of Soil Science 1953, 4, (1), 5-16.

467 468 469

22. Bloomfield, C., A study of podzolization. Journal of soil Science 1954, 5, (1), 46-56. 23. Browne, B. A.; McFee, W.; Kelly, J. In Toward a new theory of podzolization,

470 471 472 473

Carbon forms and functions in forest soils., 1995; Soil Science Society of America Inc.: 1995; pp 253-273. 24. Dahlgren, R.; Marrett, D., Organic carbon sorption in arctic and subalpine Spodosol B horizons. Soil Science Society of America Journal 1991, 55, (5), 1382-1390.

474 475 476 477

25. Lundstrom, U. S.; van Breemen, N.; Bain, D., The podzolization process. A review. Geoderma 2000, 94, (2-4), 91-107. 26. Petersen, L., Podzols and podzolization. DSR forlag Copenhagen: 1976. 27. Pohlman, A. A.; McColl, J. G., Soluble organics from forest litter and their role in

478 479 480

metal dissolution. Soil Sci. Soc. Am. J. 1988, 52, (1), 265-271. 28. De Coninck, F., Major mechanisms in formation of spodic horizons. Geoderma 1980, 24, (2), 101-128.

481 482 483 484 485 486 487 488

29. Stevenson, F. J., Humus chemistry: genesis, composition, reactions. John Wiley & Sons: 1994. 30. Fuller, C. C.; Davis, J. A.; Waychunas, G. A., Surface chemistry of ferrihydrite: Part 2. Kinetics of arsenate adsorption and coprecipitation. Geochimica et Cosmochimica Acta 1993, 57, (10), 2271-2282. 31. Chen, C. M.; Dynes, J. J.; Wang, J.; Sparks, D. L., Properties of Fe-Organic Matter Associations via Coprecipitation versus Adsorption. Environmental Science & Technology 2014, 48, (23), 13751-13759. 23

ACS Paragon Plus Environment

Environmental Science & Technology

489

32. Eusterhues, K.; Rennert, T.; Knicker, H.; Kogel-Knabner, I.; Totsche, K. U.;

490

Schwertmann, U., Fractionation of Organic Matter Due to Reaction with Ferrihydrite:

491 492 493

Coprecipitation versus Adsorption. Environmental Science & Technology 2011, 45, (2), 527-533. 33. Mikutta, R.; Lorenz, D.; Guggenberger, G.; Haumaier, L.; Freund, A., Properties

494 495 496

and reactivity of Fe-organic matter associations formed by coprecipitation versus adsorption: Clues from arsenate batch adsorption. Geochimica Et Cosmochimica Acta 2014, 144, 258-276.

497 498 499

34. Lalonde, K.; Mucci, A.; Ouellet, A.; Gelinas, Y., Preservation of organic matter in sediments promoted by iron. Nature 2012, 483, (7388), 198-200. 35. Jansen, B.; Nierop, K. G. J.; Verstraten, J. M., Mobility of Fe(II), Fe(III) and Al in

500 501 502 503 504

acidic forest soils mediated by dissolved organic matter: influence of solution pH and metal/organic carbon ratios. Geoderma 2003, 113, (3-4), 323-340. 36. Mokma, D.; Buurman, P., Podzols and podzolization in temperate regions. 1982. 37. Nierop, K. G. J.; Jansen, B.; Verstraten, J. A., Dissolved organic matter, aluminium and iron interactions: precipitation induced by metal/carbon ratio, pH and

505 506 507

competition. Science of the Total Environment 2002, 300, (1-3), 201-211. 38. Gu, B.; Schmitt, J.; Chen, Z.; Liang, L.; McCarthy, J. F., Adsorption and desorption of natural organic matter on iron oxide: mechanisms and models. Environmental

508 509 510 511

Science & Technology 1994, 28, (1), 38-46. 39. Kaiser, K.; Guggenberger, G., Sorptive stabilization of organic matter by microporous goethite: sorption into small pores vs. surface complexation. European Journal of Soil Science 2007, 58, (1), 45-59.

512 513 514 515

40. Riedel, T.; Zak, D.; Biester, H.; Dittmar, T., Iron traps terrestrially derived dissolved organic matter at redox interfaces. Proceedings of the National Academy of Sciences of the United States of America 2013, 110, (25), 10101-10105. 41. Eusterhues, K.; Wagner, F. E.; Hausler, W.; Hanzlik, M.; Knicker, H.; Totsche, K. U.;

516 517 518

Kogel-Knabner, I.; Schwertmann, U., Characterization of Ferrihydrite-Soil Organic Matter Coprecipitates by X-ray Diffraction and Mossbauer Spectroscopy. Environmental Science & Technology 2008, 42, (21), 7891-7897.

519 520 521 522 523 524 525 526

42. Karlsson, T.; Persson, P., Complexes with aquatic organic matter suppress hydrolysis and precipitation of Fe(III). Chemical Geology 2012, 322, 19-27. 43. Blodau, C., Carbon cycling in peatlands A review of processes and controls. Environmental Reviews 2002, 10, (2), 111-134. 44. Orlov, D. S., Humus acids of soils. AA Balkema: 1985. 45. Piccolo, A., The supramolecular structure of humic substances: A novel understanding of humus chemistry and implications in soil science. In Advances in Agronomy, Vol 75, Sparks, D. L., Ed. Elsevier Academic Press Inc: San Diego, 2002; Vol. 24

ACS Paragon Plus Environment

Page 24 of 36

Page 25 of 36

Environmental Science & Technology

527

75, pp 57-134.

528

46. Stevenson, F., Humus Chemistry, J. Wiley&Sons, NY 1982.

529 530 531

47. Swift, R. S.; Sparks, D.; Page, A.; Helmke, P.; Loeppert, R.; Soltanpour, P.; Tabatabai, M.; Johnston, C.; Sumner, M., Organic matter characterization. Methods of soil analysis. Part 3-chemical methods. 1996, 1011-1069.

532 533 534

48. Niemeyer, J.; Chen, Y.; Bollag, J.-M., Characterization of humic acids, composts, and peat by diffuse reflectance Fourier-transform infrared spectroscopy. Soil Science Society of America Journal 1992, 56, (1), 135-140.

535 536 537

49. Alvarez-Puebla, R. A.; Garrido, J. J., Effect of pH on the aggregation of a gray humic acid in colloidal and solid states. Chemosphere 2005, 59, (5), 659-667. 50. Sieliechi, J. M.; Lartiges, B. S.; Kayem, G. J.; Hupont, S.; Frochot, C.; Thieme, J.;

538 539 540 541 542

Ghanbaja, J.; de la Caillerie, J. B. D.; Barres, O.; Kamga, R.; Levitz, P.; Michot, L. J., Changes in humic acid conformation during coagulation with ferric chloride: Implications for drinking water treatment. Water Research 2008, 42, (8-9), 2111-2123. 51. Liu, Y. T.; Hesterberg, D., Phosphate Bonding on Noncrystalline Al/Fe-Hydroxide

543 544 545

Coprecipitates. Environmental Science & Technology 2011, 45, (15), 6283-6289. 52. Cornell, R. M.; Schwertmann, U., The iron oxides: structure, properties, reactions, occurrences and uses. John Wiley & Sons: 2003.

546 547 548 549

53. Kelly, S.; Hesterberg, D.; Ravel, B., Analysis of soils and minerals using X-ray absorption spectroscopy. Methods of soil analysis. Part5 2008, 5, 387-463. 54. Mikutta, C.; Frommer, J.; Voegelin, A.; Kaegi, R.; Kretzschmar, R., Effect of citrate on the local Fe coordination in ferrihydrite, arsenate binding, and ternary arsenate

550 551 552 553

complex formation. Geochim. Cosmochim. Acta 2010, 74, (19), 5574-5592. 55. Mikutta, C., X-ray absorption spectroscopy study on the effect of hydroxybenzoic acids on the formation and structure of ferrihydrite. Geochim. Cosmochim. Acta 2011, 75, (18), 5122-5139.

554 555 556

56. Michel, F. M.; Ehm, L.; Antao, S. M.; Lee, P. L.; Chupas, P. J.; Liu, G.; Strongin, D. R.; Schoonen, M. A. A.; Phillips, B. L.; Parise, J. B., The structure of ferrihydrite, a nanocrystalline material. Science 2007, 316, (5832), 1726-1729.

557 558 559 560 561 562 563 564

57. van Schaik, J. W.; Persson, I.; Kleja, D. B.; Gustafsson, J. P., EXAFS study on the reactions between iron and fulvic acid in acid aqueous solutions. Environ. Sci. Technol. 2008, 42, (7), 2367-2373. 58. Manceau, A.; Gates, W., Surface structural model for ferrihydrite. Clays Clay Miner. 1997, 45, (3), 448-460. 59. Karlsson, T.; Persson, P., Coordination chemistry and hydrolysis of Fe(III) in a peat humic acid studied by X-ray absorption spectroscopy. Geochimica Et Cosmochimica Acta 2010, 74, (1), 30-40. 25

ACS Paragon Plus Environment

Environmental Science & Technology

565

60. Manceau, A.; Combes, J., Structure of Mn and Fe oxides and oxyhydroxides: a

566

topological approach by EXAFS. Phys. Chem. Miner. 1988, 15, (3), 283-295.

567 568 569

61. Rose, J.; Cortalezzi-Fidalgo, M. M.; Moustier, S.; Magnetto, C.; Jones, C. D.; Barron, A. R.; Wiesner, M. R.; Bottero, J.-Y., Synthesis and characterization of carboxylate-FeOOH nanoparticles (ferroxanes) and ferroxane-derived ceramics. Chem.

570 571 572

Mater. 2002, 14, (2), 621-628. 62. Thoral, S.; Rose, J.; Garnier, J.; Van Geen, A.; Refait, P.; Traverse, A.; Fonda, E.; Nahon, D.; Bottero, J., XAS study of iron and arsenic speciation during Fe (II) oxidation

573 574 575

in the presence of As (III). Environ. Sci. Technol. 2005, 39, (24), 9478-9485. 63. Schumacher, M.; Christl, I.; Scheinost, A. C.; Jacobsen, C.; Kretzschmar, R., Chemical heterogeneity of organic soil colloids investigated by scanning transmission

576 577 578 579 580

X-ray microscopy and C-1s NEXAFS microspectroscopy. Environ. Sci. Technol. 2005, 39, (23), 9094-9100. 64. Liang, B.; Lehmann, J.; Solomon, D.; Sohi, S.; Thies, J. E.; Skjemstad, J. O.; Luizao, F. J.; Engelhard, M. H.; Neves, E. G.; Wirick, S., Stability of biomass-derived black carbon in soils. Geochim. Cosmochim. Acta 2008, 72, (24), 6069-6078.

581 582 583

65. Sparks, D. L., Kinetics and mechanisms of chemical reactions at the soil mineral/water interface. Soil physical chemistry 1999, 2, 135-191. 66. Hsu, L.-C.; Liu, Y.-T.; Tzou, Y.-M., Comparison of the spectroscopic speciation and

584 585

chemical fractionation of chromium in contaminated paddy soils. J. Hazard. Mater. 2015, 296, 230-238.

586 587 588 589 590 591 592 593 594 595 596 597 26

ACS Paragon Plus Environment

Page 26 of 36

Page 27 of 36

Environmental Science & Technology

598

Table captions Table 1. The initial added, precipitated, and dissolved C and Fe for DOM/Fe(III) coprecipitates containing various C/(C+Fe) molar ratios at pH 3, 4.5, and 6. Table 2. Results of Fe-XANES LCF analysis for DFC samples precipitated at pH 3, 4.5, and 6. The initial additions of C/(C+Fe) molar ratios are 0.62, 0.77, 0.89 and 0.94 (DFC62, DFC77, DFC89, and DFC94). Table 3. Fits to Fourier transformed EXAFS data over the k-range 2.5-11.5 Å-1 for 2-line ferrihydrite and DFC samples precipitated at pH 3, 4.5, and 6. For each precipitation condition, the initial additions of C/(C+Fe) molar ratios were 0.62, 0.77, 0.89 and 0.94 (DFC62, DFC77, DFC89, and DFC94). Figure captions Figure 1. X-ray diffraction patterns for DFC samples precipitated at (a) pH 3 (b) pH 4.5 and (c) pH 6. For each precipitation condition, the initial additions of C/(C+Fe) molar ratios were 0.62, 0.77, 0.89 and 0.94 (DFC62, DFC77, DFC89, and DFC94). Data were collected using the incident X-ray wavelength of 0.77 Å. Figure 2. The C/(C+Fe) molar ratio on near surfaces as the function of that on bulk precipitates. The diamond, triangle, and circle symbols represent the DFC samples precipitates at pH 3, 4.5, and 6, respectively. The dashed line

27

ACS Paragon Plus Environment

Environmental Science & Technology

shows the 1:1 relationship of C/(C+Fe) molar ratios on near surfaces vs. bulk particles of DFC samples. Figure 3. Fourier-transform infrared spectra for DFC samples precipitated at (a) pH 3 (b) pH 4.5 and (c) pH 6. For each precipitation condition, the initial additions of C/(C+Fe) molar ratios were 0.62, 0.77, 0.89 and 0.94 (DFC62, DFC77, DFC89, and DFC94). Figure 4. C 1s NEXAFS spectra for DFC samples precipitated at (a) pH 3 (b) pH 4.5 and (c) pH 6. For each precipitation condition, the initial additions of C/(C+Fe) molar ratios were 0.62, 0.77, 0.89 and 0.94 (DFC62, DFC77, DFC89, and DFC94).

28

ACS Paragon Plus Environment

Page 28 of 36

Page 29 of 36

Environmental Science & Technology

Table 1. The initial added, precipitated, and solubilisation C and Fe for DOM/Fe(III) coprecipitates containing various C/(C+Fe) molar ratios at pH 3, 4.5, and 6.a Copreci pitation pH

Sample

Initial added C/(C+Fe)

Precipitated C (mmole g-1)

Precipitated Fe (mmole g-1)

Bulk (precipitated) C/(C+Fe)

Bulk (precipitated) C/Fe

C solubilisation (%)

Fe solubilisation (%)

pH 3

DFC62

0.62

14.25 (0.08)

5.87 (0.02)

0.71 (0.002)

2.43 (0.023)

3.21 (1.83)

0.60 (0.10)

DFC77

0.77

23.88 (0.07)

3.49 (0.06)

0.87 (0.002)

6.84 (0.100)

2.38 (0.29)

1.95 (0.05)

DFC89

0.89

27.51 (0.01)

2.49 (0.03)

0.92 (0.001)

11.07 (0.115)

4.96 (1.16)

6.35 (0.05)

DFC94

0.94

30.08 (0.09)

1.54 (0.11)

0.95 (0.003)

19.59 (1.332)

7.00 (1.56)

3.90 (0.10)

DFC62

0.62

11.29 (0.01)

6.17 (0.01)

0.65 (0.001)

1.83 (0.005)

1.58 (0.01)

NDb

DFC77

0.77

16.75 (0.04)

5.33 (0.04)

0.76 (0.001)

3.14 (0.013)

2.69 (0.17)

0.05 (0.05)

DFC89

0.89

25.85 (0.05)

2.31 (0.02)

0.92 (0.001)

11.21 (0.055)

7.80 (2.22)

0.25 (0.25)

DFC94

0.94

28.60 (0.06)

1.47 (0.06)

0.95 (0.002)

19.43 (0.787)

9.91 (3.38)

0.60 (0.10)

DFC62

0.62

10.30 (0.03)

6.45 (0.02)

0.62 (0.001)

1.60 (0.001)

3.96 (0.44)

NDb

DFC77

0.77

16.04 (0.16)

5.75 (0.05)

0.74 (0.001)

2.79 (0.007)

12.96 (0.17)

2.85 (0.65)

DFC89

0.89

22.93 (0.05)

2.85 (0.03)

0.89 (0.001)

8.05 (0.051)

25.05 (4.23)

7.55 (1.85)

DFC94

0.94

27.90 (0.02)

2.34 (0.05)

0.92 (0.002)

11.93 (0.262)

24.90 (1.74)

8.45 (2.35)

pH 4.5

pH 6

a b

Numbers in parentheses represent standard deviations derived from two replicates. Not detected.

ACS Paragon Plus Environment

Environmental Science & Technology

Page 30 of 36

Table 2. Results of Fe-XANES LCF analysis for DFC samples precipitated at pH 3, 4.5, and 6. The initial additions of C/(C+Fe) molar ratios are 0.62, 0.77, 0.89 and 0.94 (DFC62, DFC77, DFC89, and DFC94). Bulk Ferrihydrite Fe-citrate R factor Sample (%)a C/(C+Fe) (%)a (×103)b pH 3 DFC62 0.71 81.3 (2.0) 18.7 (2.0) 0.40

pH 4.5

pH 6

DFC77

0.87

18.5 (3.6)

81.5 (3.6)

1.31

DFC89

0.92

3.7 (3.6)

96.3 (3.6)

1.28

DFC94

0.95

7.5 (2.5)

92.5 (2.5)

0.59

DFC62

0.65

100 (0.0)

0 (0.0)

4.58

DFC77

0.76

43.8 (2.0)

56.2 (2.0)

0.41

DFC89

0.92

16.5 (2.3)

83.5 (2.3)

0.56

DFC94

0.95

8.4 (1.5)

91.6 (1.5)

0.23

DFC62

0.62

100 (0.0)

0 (0.0)

2.17

DFC77

0.74

30.0 (4.9)

70.0 (4.9)

2.38

DFC89

0.89

26.5 (5.0)

73.5 (5.0)

2.34

DFC94

0.92

24.4 (1.8)

75.6 (1.8)

0.32

a

Numbers in parentheses represent standard deviation. The weighting factors on each fit were summed to 100 ± 1% and were normalized to 100%. Although five Fe species were used as end-members in the LCF analyses, the best fits for all samples were obtained using a combination of ferrihydrite and Fe(III) complexed with citrate. b Normalized sum of the squared residuals of the fit (R-factor = ∑(data-fit)2/∑data2).

ACS Paragon Plus Environment

Page 31 of 36

Environmental Science & Technology

Table 3. Fits to Fourier transformed EXAFS data over the k-range 2.5-11.5 Å-1 for 2-line ferrihydrite and DFC samples precipitated at pH 3, 4.5, and 6. For each precipitation condition, the initial additions of C/(C+Fe) molar ratios were 0.62, 0.77, 0.89 and 0.94 (DFC62, DFC77, DFC89, and DFC94).a Fe-C (SS)

pH 3

pH 4.5

pH 6

Fe-Fe1 (SS)

Sample

CN

R (Å)

σ2 (Å2)b

CN

Ferrihydrite

-

-

-

2.84 (0.48)

DFC62

1.77 (0.14)

2.96 (0.011) 0.004 (0.001)

DFC77

1.65 (0.71)

DFC89

Fe-Fe2 (SS) CN

R (Å)

σ2 (Å2)b

3.03 (0.011) 0.014 (0.002)

0.79 (0.22)

3.45 (0.014)

0.005 (0.002)

1.77 (0.14)

3.07 (0.005) 0.008 (0.001)

1.35 (0.53)

3.43 (0.037)

0.018 (0.002)

2.96 (0.011) 0.004 (0.001)

0.82 (0.31)

3.07 (0.005) 0.008 (0.001)

1.77 (1.06)

3.43 (0.037)

0.018 (0.002)

1.64 (1.10)

2.98 (0.019) 0.004 (0.001)

-

-

-

1.87 (1.74)

3.40 (0.051)

0.018 (0.002)

DFC94

1.58 (0.69)

2.98 (0.019) 0.004 (0.001)

-

-

-

1.66 (0.96)

3.41 (0.041)

0.018 (0.002)

DFC62

-

DFC77

1.13 (0.42)

DFC89

-

σ2 (Å2)b

0.87 (0.21)

3.02 (0.025) 0.009 (0.001)

0.64 (0.32)

3.42 (0.024)

0.008 (0.002)

2.96 (0.033) 0.004 (0.002)

0.88 (0.21)

3.05 (0.008) 0.009 (0.001)

0.87 (0.21)

3.43 (0.008)

0.008 (0.002)

1.88 (1.62)

2.99 (0.026) 0.004 (0.002)

-

-

-

0.88 (0.87)

3.43 (0.033)

0.008 (0.002)

DFC94

1.33 (0.72)

2.99 (0.026) 0.004 (0.002)

-

-

-

0.47 (0.36)

3.43 (0.033)

0.008 (0.002)

DFC62

-

DFC77

1.32 (1.11)

DFC89 DFC94

-

-

R (Å)

-

1.02 (0.20)

3.01 (0.015) 0.010 (0.001)

0.55 (0.15)

3.44 (0.008)

0.005 (0.002)

2.95 (0.023) 0.006 (0.004)

1.66 (0.50)

3.06 (0.008) 0.010 (0.001)

0.47 (0.35)

3.44 (0.008)

0.005 (0.002)

1.30 (0.93)

2.95 (0.023) 0.006 (0.004)

1.16 (0.42)

3.06 (0.008) 0.010 (0.001)

0.39 (0.27)

3.44 (0.008)

0.005 (0.002)

1.55 (1.26)

2.95 (0.023) 0.006 (0.004)

-

0.39 (0.38)

3.43 (0.059)

0.005 (0.002)

-

-

Paths used in EXAFS fitting are all single scattering (SS) paths. S02 (fixed amplitude reduction factor based on first-shell fitting of hematite) = 0.83. ΔE (fitted energy shift parameter) = -2.44 (0.43), -3.74 (0.26), and -2.85 (0.31) eV for DFC samples precipitated at pH 3, 4.5, and 6. For the Fe-O paths among all samples, the interatomic distances were fitted around 1.96-1.99 Å and the fitted σ2 (mean-square displacements of interatomic distances) values ranged from 0.006-0.019 Å2. The CN (coordination number) was fixed at 6.0. Details for fitting parameters of Fe-O paths were tabulated in Table S3 in SI. b 2 σ values were grouped according to precipitated pH and fitted. a

ACS Paragon Plus Environment

Environmental Science & Technology

    (a) pH 3

2.6 Å

1.5 Å DFC62

DFC77

DFC89 DFC94

(b) pH 4.5 2.6 Å 1.5 Å

intensity (cps)

DFC62 DFC77 DFC89 DFC94

(c) pH 6

15000

2.6 Å 1.5 Å DFC62

10000 DFC77 DFC89

5000

DFC94

0 0

10

20

30

40

50

2theta

Figure 1. X-ray diffraction patterns for DFC samples precipitated at (a) pH 3 (b) pH 4.5 and (c) pH 6. For each precipitation condition, the initial additions of C/(C+Fe) molar ratios were 0.62, 0.77, 0.89 and 0.94 (DFC62, DFC77, DFC89, and DFC94). Data were collected using the incident X-ray wavelength of 0.77 Å.

ACS Paragon Plus Environment

Page 32 of 36

Page 33 of 36

Environmental Science & Technology

      1.0

Surface C/(C+Fe) ratio

0.9

0.8 pH 3 pH 4.5

0.7

pH 6 0.6 0.6

0.7

0.8

0.9

1.0

Bulk C/(C+Fe) ratio Figure 2. The C/(C+Fe) molar ratio on near surfaces as the function of that on bulk precipitates. The diamond, triangle, and circle symbols represent the DFC samples precipitates at pH 3, 4.5, and 6, respectively. The dashed line shows the 1:1 relationship of C/(C+Fe) molar ratios on near surfaces vs. bulk particles of DFC samples. 

ACS Paragon Plus Environment

(a) pH 3

1050

1270

1400

1610

1730

Environmental Science & Technology

DFC62 DFC77 DFC89 DFC94 DOM

(b) pH 4.5

DFC62

Absorbance

DFC77 DFC89 DFC94

DOM

3

(c) pH 6  

DFC62 DFC77

2 DFC89 DFC94

1 DOM

0 2000

1800

1600

1400

1200

1000

800

Wavenumber (cm-1)

Figure 3. Fourier-transform infrared spectra for DFC samples precipitated at (a) pH 3 (b) pH 4.5 and (c) pH 6. For each precipitation condition, the initial additions of C/(C+Fe) molar ratios were 0.62, 0.77, 0.89 and 0.94 (DFC62, DFC77, DFC89, and DFC94).  

ACS Paragon Plus Environment

Page 34 of 36

Page 35 of 36

Environmental Science & Technology

phenolic C

(a) pH3 aromatic C carboxyl C DFC62 DFC77 DFC89 DFC94 DOM

(b) pH4.5 Normalized intensity (a.u.)

DFC62 DFC77 DFC89 DFC94 DOM

(c) pH6 1.0

DFC62 DFC77 DFC89

0.5

DFC94 DOM

0.0 280

285

290

295

Energy (eV)

Figure 4. C 1s NEXAFS spectra for DFC samples precipitated at (a) pH 3 (b) pH 4.5 and (c) pH 6. For each precipitation condition, the initial additions of C/(C+Fe) molar ratios were 0.62, 0.77, 0.89 and 0.94 (DFC62, DFC77, DFC89, and DFC94).

ACS Paragon Plus Environment

Environmental Science & Technology

TOC/A Abstract A rt (4.5 cm m*5.88 cm m)

 

ACS Paragon Plus Environment

Page 36 of 36