Structural Transformation of Birnessite by Fulvic Acid under Anoxic

Jan 22, 2018 - Thus, the system is changing toward favoring FA adsorption with increasing time, which keeps increasing the FA adsorption capacity of t...
1 downloads 11 Views 802KB Size
Subscriber access provided by READING UNIV

Article

Structural Transformation of Birnessite by Fulvic Acid under Anoxic Conditions Qian Wang, Peng Yang, and Mengqiang Zhu Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b04379 • Publication Date (Web): 22 Jan 2018 Downloaded from http://pubs.acs.org on January 27, 2018

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 31

Environmental Science & Technology

1 2

Structural Transformation of Birnessite by Fulvic Acid

3

under Anoxic Conditions

4 5

Qian Wang, Peng Yang, and Mengqiang Zhu*

6

Department of Ecosystem Science and Management, University of Wyoming, Laramie, WY

7

82071, United States

8 9 10

Submitted to Environmental Science & Technology

11

January 10, 2018

12 13 14

*Corresponding author: Mengqiang Zhu

15

E-mail: [email protected]

16

Phone: 307-766-5523

17 18 19 20

1 ACS Paragon Plus Environment

Environmental Science & Technology

21

Page 2 of 31

ABSTRACT

22

The structure and Mn(III) concentration of birnessite dictate its reactivity, and can be

23

changed by birnessite partial reduction, but effects of pH and reductant/birnessite ratios on the

24

changes by reduction remain unclear. We found that the two factors strongly affect structure of

25

birnessite (δ-MnO2) and its Mn(III) content during its reduction by fulvic acid (FA) at pH 4 – 8

26

and FA/solid mass ratios of 0.01 – 10 under anoxic conditions over 600 hours. During the

27

reduction, the structure of δ-MnO2 is increasingly accumulated with both Mn(III) and Mn(II) but

28

much more with Mn(III) at pH 8, whereas the accumulated Mn is mainly Mn(II) with little

29

Mn(III) at pH 4 and 6. Mn(III) accumulation, either in layers or over vacancies, is stronger at

30

higher FA/solid ratios. At FA/solid ratios ≥ 1 and pH 6 and 8, additional hausmannite and

31

MnOOH phases form. The altered birnessite favorably adsorbs FA because of the structural

32

accumulation of Mn(II, III). Like during oxidative precipitation of birnessite, the dynamic

33

changes during its reduction are ascribed to the birnessite-Mn(II) redox reactions. Our work

34

suggests low reactivity of birnessite co-existing with organic matter and severe decline of its

35

reactivity by partial reduction in alkaline environment.

36 37

INTRODUCTION

38

Birnessite is a layered manganese (Mn) oxide and common in sediments and soils.1, 2 It is

39

a metal scavenger and a strong oxidizer, and affects the environmental fate of toxic metals and

40

organic pollutants,3-7 and biogeochemical cycles of carbon, nitrogen, sulfur and iron.5, 8-11 The

41

high reactivity of birnessite stems from its open-framework (layered) structure,12 abundant

42

defects (vacancies), and the high oxidizing potential of tetravalent Mn (Mn(IV)).1, 13-15

43

Partial reduction of birnessite by organic matter can remarkably increase Mn(III) content

2 ACS Paragon Plus Environment

Page 3 of 31

Environmental Science & Technology

44

of birnessite,4, 16-19 which can profoundly change birnessite metal sorption capacity and behavior,

45

oxidizing reactivity and bandgap, and affect its transformation to tunneled structures.10, 12, 20-23

46

Reduction of birnessite by humate or low molecular weight (LMW) organics under oxic

47

conditions results in enrichment of both Mn(II) and Mn(III) in the reacted solid.16-18,

48

Mn(III) proportion can be increased to surprisingly high levels (37.9% – 56%) based on analyses

49

of X-ray photoelectron spectroscopy (XPS) and Mn K-edge X-ray absorption near-edge structure

50

(XANES) spectroscopy.4, 16-19 However, the mechanism for the accumulation of such high levels

51

of Mn(III), and the effects of pH and reductant/birnessite ratios on the Mn(III) accumulation and

52

birnessite structural transformation remain unclear. It is poorly understood if Mn(III) resides on

53

vacancies, in the layers, or both in the altered birnessite structure, which is important to know as

54

the specific location of Mn(III) affects birnessite properties differently.20-22 It is unknown

55

whether the reacted solid remains as birnessite with structural alteration, or converts to other Mn

56

mineral phases of lower Mn oxidation states, such as hausmannite (Mn3O4) and MnOOH phases.

57

The lack of the solid phase information also hinders understanding of the reaction mechanisms

58

behind the changes in Mn oxidation state composition. Moreover, these studies were conducted

59

in air where O2 could participate in the redox reaction, limiting the transferability of the results to

60

suboxic environment.

24

The

61

Birnessite and aqueous divalent Mn (Mn(II)(aq)) coexist during reduction of birnessite and

62

can react with each other. Adsorption and oxidation of Mn(II)(aq) by birnessite can dramatically

63

change birnessite structure and Mn(III) concentration. Under circumneutral or alkaline

64

conditions, Mn(III) is formed with a concomitant decrease in vacancy concentration at a low

65

Mn(II)(aq)/MnO2 ratio due to the conproportionation or redox reaction between adsorbed Mn(II)

66

and Mn(IV) in birnessite;6,

25-28

at a high Mn(II)(aq)/MnO2 ratio, the birnessite-Mn(II)

3 ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 31

67

conproportionation converts birnessite into hausmannite and MnOOH phases.1, 26, 27, 29-31 The

68

above processes are disfavored under acidic conditions.28, 29 Thus, we hypothesize that the above

69

reductive transformation of birnessite by Mn(II) is responsible for Mn(III) accumulation in the

70

solid phase during birnessite reduction, and the mineral phase of the reacted products depends on

71

pH and the ratio of the reductant to birnessite.

72

In the present study, we characterized temporal changes of the structure and Mn

73

oxidation state composition of δ-MnO2 during its reaction (adsorption and oxidation) with fulvic

74

acid (FA) and determined FA removal kinetics at various FA/solid mass ratios (R) and pH values

75

under anoxic conditions. δ-MnO2, a nanocrystalline hexagonal birnessite, is analogous to

76

natually occuring vernadite and biogenic Mn oxides.2 Fulvic acid, an important fraction of

77

natural organic matter (NOM), was used as a model reductant to approximate NOM. The

78

experiments were conducted at pH 4, 6 or 8, representing the common pH conditions of suboxic

79

environment.28, 32, 33 Unlike in the previous studies,1, 4, 29 pH buffers were not used as they may

80

reduce birnessite and complicate the interpretation of the results.27, 34 The mineral phases, local

81

to intermediate range structures, and Mn oxidation state compositions of the reacted soilds

82

collected at different time were thoroughly characterized using synchrotron-based X-ray

83

diffraction (XRD), X-ray atomic pair distribution function (PDF) analysis, and Mn K-edge

84

XANES spectroscopy. The observations are interpreted in light of the most recent progress on

85

birnessite reductive transformaton by reacting with Mn(II)(aq).1, 26, 29 We find that accumulation

86

and the location of Mn(III) in birnessite structure strongly depends on pH and R (FA/solid mass

87

ratios), and the reacted solids include hausmannite and MnOOH phases in addition to Mn(III)-

88

rich birnessite. These results advance our understanding of the dynamic changes in birnessite

89

structure, composition and reactivity during birnessite reduction, an important process during Mn

4 ACS Paragon Plus Environment

Page 5 of 31

Environmental Science & Technology

90

redox cycling, and shed light on the role of Mn oxides in mineralization and stabilization of

91

organic carbon.

92

MATERIALS AND METHODS

93

Materials. δ-MnO2 was synthesized via reduction of KMnO4 by Mn(NO3)2 (Supporting

94

Information, SI-1). Suwannee River FA (1S101F) was purchased from the International Humic

95

Substance Society (SI-1). All other chemicals used were of analytical grade. The O2-free

96

deionized (DI) water was used for preparation of chemical solutions (SI-1).

97

Reaction of FA with δ-MnO2. Batch experiments were conducted in a COY vinyl

98

anaerobic chamber (95% N2/5% H2) over a Pd catalyst at 22 ± 0.1 oC. A total 500 mL suspension

99

in a capped polyethylene bottle containing δ-MnO2 (0.03, 0.3 or 3 g/L, equivalent to 0.343, 3.43

100

or 34.3 mM Mn), FA (0, 0.03 or 0.3 g/L, equivalent to 0, 0.0142 or 0.142 mM FA based on the

101

empirical molecular weight of 2114 Da35) and NaCl (50 mM) was maintained at pH 4, 6 or 8

102

over 600 hours (h) while being vigorously stirred with a magnetic stirring bar. Thus, R was 0,

103

0.01 (0.03: 3), 0.1 (0.3:3), 1 (0.3:0.3) or 10 (0.3: 0.03). Those conducted at R = 0 were the

104

control experiments. The bottles were wrapped with aluminum foil to prevent exposure to light

105

that can promote reductive dissolution of birnessite.8, 24, 36

106

Prior to mixing the reactants, both the FA solution and the δ-MnO2 suspension were

107

maintained at the target pH in the anaerobic chamber for 24 h by manually adding 0.1 M NaOH

108

or HCl. An aliquot of the δ-MnO2 suspension was sampled (0-h sample) right before the mixing.

109

The pH of the mixed suspension was quickly raised to the target value within ~ 0.02 h using an

110

auto pH titrator (Metrohm 907 Titrando) when another aliquot of the suspension was sampled

111

(0.02-h sample). Then the pH of the suspension was maintained for another 5 h using the titrator,

112

during which the pH fluctuated slightly (± 0.01), before switching to manual pH adjustment for 5 ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 31

113

the remaining period of the experiments that lasted for 600 h. At pre-determined time intervals,

114

aliquots (~ 20 mL) were collected and the solid and solution were separated using 0.2 µm

115

membrane filters. The solid particles aggregated extensively and the amount of particles passing

116

through the filter was negligible. The concentration of dissolved Mn ([Mn]dis) in the filtrates was

117

measured by inductively coupled plasma optical emission spectrometry. The obtained solids

118

were air dried in the anaerobic chamber, ground in an agate mortar and stored in a closed glass

119

vial in the chamber prior to the following characterization. The FA remained in the solution was

120

quantified as total dissolved organic carbon (DOC) (SI-2). Both specific UV absorbance at 254

121

nm (SUVA254) and the ratio of the UV absorbance at 250 nm to that at 365 nm (E2/E3 index)

122

based on UV-vis spectra of the filtrates were used to indicate compositional changes of FA (SI-2).

123

A thorough characterization of reacted FA is beyond the scope of this work. LMW organic

124

products in the filtrates were not measured.

125

Characterization of Reacted Solids. Both XRD and total X-ray scattering data for the

126

PDF analysis were collected from the air-dried solids using X-rays of 58.6491 keV (λ= 0.2114 Å)

127

at beamline 11-ID-B at the Advanced Photon Source (APS), with the sample-to-detector

128

distances of ~ 95 cm and ~ 15 cm, respectively (SI-3). The data collection and processing were

129

as described in Wang et al.12

130

The relative molar fractions of Mn(II), Mn(III) and Mn(IV) in the reacted solids were

131

estimated using linear combination fitting (LCF) of Mn K-edge XANES spectra. The XANES

132

spectra were collected from the solids in fluorescence mode using a monochromator equipped

133

with Si (111) double crystal at beamline 10-BM-B at APS (SI-4). The LCF analysis was similar

134

to the Combo method,37 and four Mn reference standards were finally chosen, i.e., MnSO4 solid

135

for Mn(II), feitknechtite (β-MnOOH) for Mn(III), and a combination of potassium birnessite

6 ACS Paragon Plus Environment

Page 7 of 31

Environmental Science & Technology

136

(KBi) and polymeric manganese oxides (PolyMnO2) for Mn(IV). The first three reference

137

spectra were reported in Manceau et al.37 PolyMnO2 contains entirely Mn(IV)38, 39 and has a

138

particle size smaller than that of δ-MnO2.40 As particle sizes can affect Mn K-edge XANES

139

spectra,12 using PolyMnO2 in the fitting was to account for decreased particle size of δ-MnO2

140

due to the reduction. A sensitivity analysis of the fitting strategy to the relative fraction of Mn(III)

141

was conducted following the method in Grangeon et al.,41 showing the fitting results were

142

reliable (SI-5). The obtained Mn oxidation state composition and the measured dissolved Mn(II)

143

concentration were combined to quantify the net electron transfer from FA to δ-MnO2 (SI-6).

144

A portion of the Mn(III) in the layers of birnessite migrates into interlayers upon

145

acidification,6 leading to structural changes of birnessite. The changes can be used as another

146

indicator for the presence of Mn(III) in the layers of birnessite. The end solid products reacted

147

for 600 h at pH 6 and 8 were acidified and the structural changes were characterized (SI-7).

148

RESULTS

149

Dissolved Mn. In the control experiments (R = 0), < 1 µM dissolved Mn, likely Mn(II),

150

is detected at pH 4 but at neither pH 6 nor pH 8 (Figure S1). The presence of dissolved Mn is due

151

to both destabilization of structural Mn(III) in δ-MnO2 to form Mn(II) and disfavored Mn(II)

152

adsorption under highly acidic conditions (pH 4).6

153

At pH 6 and 8, R = 0.01, low [Mn]dis is detected because the low concentration of FA

154

produces little Mn(II), most of which is resorbed onto δ-MnO2 (Figure 1B and 1C). The

155

resorption also results in the initial low [Mn]dis under other conditions (Figure 1). Higher [Mn]dis

156

is detected at R = 0.1 and 1, which increases with time but does not reach plateaus within 600 h

157

(Figure 1). At R = 10 and pH 8, the solid is completely dissolved within 72 h (Figure 1C).

158

Overall, at a given pH, a higher R results in higher [Mn]dis because more FA participates in the

7 ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 31

159

reduction (Figure 1); for a given R, a higher pH leads to much lower [Mn]dis and Mn release rates

160

(Figure 1), ascribed to both disfavored FA adsorption due to increasing negativity of both FA

161

and surface, producing less Mn(II), and enhanced Mn(II) sorption at the higher pH.

162

Adsorption and Oxidation of FA. The DOC and solid C concentrations at different pH

163

and R are given in Figure 2A, 2B, S2 and S3. Both solid C and DOC concentrations display slow

164

removal kinetics of FA from solution and neither reach plateaus within 600 h (Figure 2A, 2B, S2

165

and S3). Both lower pH and lower R lead to faster removal kinetics, a higher C removal from

166

solution, and more C accumulated in the solids (Figure 2A, 2B, S2 and S3).

167

The UV-vis absorbance of the FA in the filtrates decreases at all wavelengths

168

simultaneously with increasing time (Figure S4), suggesting that FA adsorption on the solid

169

plays a more important role for the FA removal from solution, rather than FA oxidative

170

decomposition. Otherwise, the UV-vis absorbance would change only at the wavelengths

171

belonging to the oxidized FA moieties.

172

The band at 254 nm, corresponding to aromatic functional groups, becomes sharper with

173

time (Figure S4), suggesting that remaining FA differs in composition from the initial FA. The

174

SUVA254 and E2/E3 indices are used to characterize the compositional changes. Generally,

175

SUVA254 decreases and E2/E3 increases significantly with increasing time, indicating that both

176

aromaticity and molecular weight of remaining FA decrease as the reaction proceeds (Figure 2C,

177

2D and S5), i.e., preferentially removing the fraction of FA high in both aromaticity and

178

molecular weight, consistent with previous studies.42,

179

remained in the solution are caused by both adsorption and oxidation of FA, probably more by

180

the former which is well known to cause FA fractionation.42,

181

fractionation are stronger at both lower R and lower pH, at which more FA is also removed

43

The compositional changes of FA

43

The FA oxidation and

8 ACS Paragon Plus Environment

Page 9 of 31

182

Environmental Science & Technology

(Figure 2C, 2D and S5).

183

Mineral Phases. The XRD patterns of the solids in the control experiments in the

184

absence of FA (R = 0) consist of the characteristic peaks of δ-MnO2 (Figure S6A).40 After 600 h,

185

these peaks change slightly, indicating subtle structural alteration of δ-MnO2. The changes of the

186

XRD patterns at R = 0.01 are also subtle because of much lower concentration of FA used than

187

that of δ-MnO2 (Figure S6B and S6C).

188

Figure 3 shows the XRD patterns of the reacted δ-MnO2 at R > 0.01. The (002) peak

189

broadens significantly with increasing reaction time under all conditions, indicating fewer

190

stacked MnO6 layers and/or a decrease of the degree of parallelism between adjacent layers,12

191

which can be caused by Mn(II) and/or FA adsorption on δ-MnO2.29 At R = 0.1, the solid remains

192

as δ-MnO2 at all pH values during the reaction (Figure 3A-3C). A dip at ~ 1.97 Å becomes

193

increasingly pronounced, indicating more Mn (II, III) capping the vacancies as the reaction

194

proceeds.4, 23, 41, 44 At a given reaction time, the dip is more pronounced at lower pH, indicating

195

more Mn (II, III) on the vacancies,6, 41, 44, 45 as higher [Mn]dis is produced.

196

Similar but stronger changes are observed at R = 1 than at R = 0.1 (Figure 3D and 3E). In

197

addition, a small portion of δ-MnO2 is transformed to other mineral phases at R = 1. At pH 6,

198

minor hausmannite and groutite (α-MnOOH) form at 10 h (Figure 3D), but groutite disappears

199

with time. At pH 8, feitknechtite (β-MnOOH) is detected at 24 h and then disappears over time;

200

hausmannite forms at 120 h and becomes more abundant with time (Figure 3E). Hausmannite

201

forms earlier, but its amount is less at pH 6 than at pH 8.

202

At R = 10, both MnOOH phases form in addition to hausmannite. Hausmannite forms

203

more rapidly and extensively at R = 10 than at R = 1, indicated by the earlier appearance of the

204

stronger XRD peaks of hausmannite (Figure 3F). No solids remain after 72 h, indicating that 9 ACS Paragon Plus Environment

Environmental Science & Technology

205

Page 10 of 31

both hausmannite and MnOOH phases are reductively dissolved completely by FA.

206

Local Structural Changes. The PDF analysis is sensitive to structural changes of short

207

and intermediate range, which may not be revealed unambiguously by XRD. The two peaks at ~

208

2.86 Å ((Mn-Mn)L) and ~ 5.27 Å (MnL-MnIL), derived from the edge-sharing correlation in the

209

layer (L) and the corner-sharing correlation between layers and interlayers (IL) (SI-3),

210

respectively, are used to indicate the abundance of layer Mn(III) and interlayer Mn(II, III) of the

211

reacted solids that remain as birnessite. The peak position corresponds to the interatomic distance

212

of the two atoms involved in the correlation. A higher concentration of layer Mn(III) leads to a

213

longer (Mn-Mn)L interatomic distance because Mn(III) has a larger ionic radius than Mn(IV).40

214

Likewise, the MnL-MnIL peak is located at a higher r with a higher proportion of Mn(II) on

215

vacancies because Mn(II) has a larger radius than Mn(III). In addition, when the vacancies are

216

capped with more Mn(II, III), the MnL-MnIL peak becomes stronger. Thus, the PDF analysis

217

allows for detailed characterization of the relative proportions of vacancy-adsorbed Mn(II) and

218

Mn(III), and of the location (layers versus vacancies) of the accumulated Mn(III) with time

219

during δ-MnO2 reduction.

220

No changes in PDFs are observed for the control samples (R = 0) at pH 6 and 8 (Figure

221

S7B). However, at pH 4, the (Mn-Mn)L peak slightly shifts to lower r, and the MnL-MnIL peak

222

becomes stronger after 600 h of reaction. These slight changes suggest migration of layer Mn(III)

223

onto vacancies (Figure S7B). Similarly at R = 0.1, with increasing time, the Mn(III) migration

224

leads to the left shift (towards lower r) of the (Mn-Mn)L peak at pH 4 and 6 with less shift at pH

225

6 (Figure 4A, 4B). In contrast, at R = 0.1 and pH 8, the (Mn-Mn)L peak shifts to the right

226

(towards higher r) with time (Figure 4C), indicating that the layers contain more Mn(III).

227

With increasing time, the MnL-MnIL peak shifts to the right and the peak intensity

10 ACS Paragon Plus Environment

Page 11 of 31

Environmental Science & Technology

228

increases for all pH at R = 0.1 and 0.01 with less changes at R = 0.01 (Figure 4 and S8),

229

indicating more Mn (II, III) capping the vacancies and a higher proportion of Mn(II) as the

230

reaction proceeds. At a given reaction time, with increasing pH from 4 to 8 at R = 0.1, the MnL-

231

MnIL peak area decreases and the peak position shifts to lower r (Figure 4 and 5). These changes

232

indicate that at a higher pH, less Mn is adsorbed on vacancies, and the adsorbed Mn contains a

233

higher proportion of Mn(III).

234

To determine the kinetics of the structural alteration of δ-MnO2, the (Mn-Mn)L peak

235

position and the MnL-MnIL peak area are plotted versus reaction time (Figure 5). Both change

236

rapidly within 72 h and then slow down (Figure 5). At R = 1 and pH 6, the (Mn-Mn)L peak

237

position shifts to the shorter distance up to 5 h, ascribed to the migration of layer Mn(III); the

238

abrupt jump to a longer distance at 10 h is ascribed to the presence of groutite (Figure 3D, 5A

239

and S8D) that contains longer Mn-Mn pairs. At R = 1 and pH 8, the (Mn-Mn)L interatomic

240

distance increases before the appearance of feitknechtite at 5 h, indicating more layer Mn(III) in

241

δ-MnO2 (Figure 5A and S8E). After 5 h, the increase of the distance is due to both Mn(III)

242

enrichment in δ-MnO2 and the presence of feitknechtite and hausmannite.

243

Mn Oxidation States and Net Electron Transfer. The LCF analysis is used to

244

determine the molar fraction of each Mn oxidation state in the reacted solids. A comparison of

245

the fits to the spectra shows good agreement (Figure S10). At R = 0.1, the Mn(II) fraction

246

reaches 9.3% at pH 6 and 10.9% at pH 4 after 600 h; however, the Mn(III) fractions (0 – 3.6%)

247

are very low and change negligibly (Figure 6 and Table S1). At pH 8 and R = 0.1 (Figure 6 and

248

Table S1), the Mn(III) fraction increases from 0.4% to 14.2% with negligible changes in the

249

Mn(II) fraction (0.4 to 4.0%). At pH 8 and R = 1, the fractions of Mn(III) and Mn(II) reach

250

27.9% and 11.2%, respectively, within 600 h (Figure 6 and Table S1). Our fitting may

11 ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 31

251

underestimate the Mn(III) fractions. For example, the LCF analysis shows that triclinic

252

birnessite, which ideally consists of one third of Mn as Mn(III), contains only 26.3% Mn(III);

253

and a chemical titration analysis shows that as-synthesized δ-MnO2 contains 6% Mn(III) (SI-1),

254

but Mn(III) is not detected by the LCF analysis (Table S1).

255

Figure 7 shows the estimated net electron transfer from FA to δ-MnO2 during the reaction

256

based on dissolved Mn(II) and the LCF-determined Mn oxidation state composition (Table S1).

257

The net electron transfer versus reaction time better represents the progress of the redox reaction

258

than based on only dissolved Mn(II) because the solids contain high reduced Mn (i.e., Mn(II) and

259

Mn(III)). At R = 0.1, the net electron transfer at pH 4 is much higher than at pH 6 and pH 8; and

260

interestingly, no significant difference occurs between pH 6 and pH 8 at R = 0.1 (Figure 7).

261

DISSCUSSION

262

Adsorption and Oxidation of FA on δ-MnO2. Adsorption of NOM on mineral surfaces

263

can involve electrostatic interactions, ligand exchange, cation bridging, and hydrophobic

264

interactions.42, 46 The electrostatic interactions make little contribution to adsorption of FA on

265

δ-MnO2 because both FA and δ-MnO2 are negatively charged at pH 4 - 8. The carboxylic and/or

266

phenolic groups of FA are responsible for ligand exchange, likely with >Mn-OH at the edge sites

267

of the δ-MnO2 layers because the hydroxyls associated with vacancies are inert for ligand

268

exchange as each is coordinated to two Mn atoms. However, adsorbed Mn(II, III) on vacancies

269

can enhance FA adsorption via cation bridging.47 Hydrophobic interactions could also contribute

270

to the adsorption of FA on δ-MnO2.42

271

Both the surface properties of the solid and the FA composition (because of FA oxidation)

272

are changing during reaction of FA with δ-MnO2, affecting FA removal from solution by

273

adsorption. The enrichment of Mn(II, III) increases PZC of δ-MnO2, and both hausmannite and 12 ACS Paragon Plus Environment

Page 13 of 31

Environmental Science & Technology

274

the MnOOH phases have higher PZC values than birnessite.48 These lead to less negatively

275

charged solids, favoring FA adsorption. In addition, complexation of FA with dissolved Mn in

276

solution decreases the negative charge of FA and favors FA adsorption as well. Thus, the system

277

is changing towards favoring FA adsorption with increasing time, which keeps increasing the FA

278

adsorption capacity of the solid. Consequently, it is hard for FA adsorption to reach plateaus with

279

time, i.e., displaying the slow FA removal kinetics (Figure 2).

280

In addition, the reduction rate of δ-MnO2 by FA decreases over time, which is also

281

observed during birnessite reduction by other reductants.18, 23, 36, 49 The decreased rate can be

282

caused by the depletion of oxidizable functional groups of FA, and the reduced oxidizing

283

potential of MnO2/Mn(II)(aq).49 Surface passivation of δ-MnO2 by adsorbed Mn(II, III) and

284

adsorption of oxidized FA could also contribute.18, 23 In fact, the accumulation of Mn(III) in δ-

285

MnO2 with time (e.g., R = 0.1 and pH 8 in Figure 5) indicates lower reactivity of Mn(III) than

286

Mn(IV) in oxidizing FA. The accumulation of MnOOH and hausmannite at R = 0.1 (Figure 3)

287

also indicates the lower oxidizing reactivity of these phases than birnessite, and thus their

288

formation slows down the overall reduction.

289

Low pH favors both adsorption and oxidation of FA by δ-MnO2. Both FA and δ-MnO2

290

surfaces become more protonated at lower pH, decreasing their negative charge and favoring FA

291

adsorption and further oxidation (Figure 1A and 2A), consistent with Allard et al.42 Lower pH

292

also favors FA oxidation from a thermodynamic perspective because the redox reaction

293

consumes H+.50 However, similar amounts of electron transfer at pH 6 and 8 are observed

294

(Figure 7), probably due to uncertainties in the calculation of electron transfer that relies on Mn

295

oxidation state composition. In terms of dissolved Mn, low pH supresses Mn(II) back adsorption

296

to δ-MnO2,6 increasing dissolved Mn concentration and thus favorable to FA adsorption.

13 ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 31

297

Moreover, structural modification and phase transformation of δ-MnO2 during oxidizing FA

298

highly depend on pH, which further affects FA adsorption and oxidation as the reaction proceeds.

299

Mn(III) in the Structure of δ-MnO2. Trivalent Mn remarkably affects birnessite

300

properties.10, 12, 20-23 Our XRD and PDF analyses indicate that at R = 0.1, δ-MnO2 is enriched in

301

Mn(III) at pH 8 but not at pH 4 and 6 under anoxic conditions (Figure 3, 4 and 5). The

302

accumulation of Mn(III) in the layers at pH 8 is further supported by the acidification experiment

303

showing that a considerable amount of the layer Mn(III) in the solids at pH 8 migrates out of

304

layers upon acidification (SI-7).

305

Two mechanisms can lead to the Mn(III) accumulation in δ-MnO2 at pH 8 during

306

δ-MnO2 reduction by FA. In the first mechanism, Mn(III) is formed as an intermediate product

307

during Mn(IV) reduction to Mn(II) by FA. A portion of the Mn(III) is not reactive enough to

308

oxidize FA while being stabilized on the surface by FA complexation.16, 17 Since FA most likely

309

attacks the edge sites, the intermediate Mn(III) initially formed must also be located at the edge

310

sites. Our PDF data indicate that a significant portion of the accumulated Mn(III) resides in the

311

layers of δ-MnO2. Therefore, if this mechanism applies, the Mn(III) formed at the edge sites

312

must be redistributed in the layers, likely through intra-layer electron transfer between Mn(IV)

313

and Mn(III). However, such mechanism cannot explain the presence of abundant Mn(III) on

314

vacancies. In the second mechanism, Mn(III) forms as a result of the conproportionation between

315

Mn(IV) and Mn(II).23 In this mechanism, Mn(II) produced by Mn(IV) reduction adsorbs on a

316

vacancy and then reacts with a Mn(IV) cation surrounding the vacancy to form two Mn(III)

317

cations, with one in the layer and the other adsorbed on the vacancy or incorporated into the

318

layer.6 This conproportionation mechanism does not require FA complexation to stabilize the

319

Mn(III). Numerous recent studies support the conproportionation mechanism.1,

26, 27

This 14

ACS Paragon Plus Environment

Page 15 of 31

Environmental Science & Technology

320

mechanism can also better explain the presence of Mn(III) on vacancies than the intermediate

321

mechanism.

322

Interestingly, at pH 4 and 6, the content of layer Mn(III) decreases but the content of the

323

all structural Mn(III) (both layered and interlayered) remains almost constant or increases

324

slightly at low levels with increasing time. This suggests that layer Mn(III) migrates into the

325

interlayer without further disproportionation. Grangeon et al. shows that adsorption of Zn2+ and

326

Ni2+ on vacancies induces the migration of layer Mn(III) at pH < 7.51, 52 Mn(II) adsorbs on

327

vacancies at low pH is not supposed to reduce layer Mn(IV) and it probably behaves similarly as

328

Zn2+ and Ni2+ at pH 4 and 6, leading to decreased layer Mn(III) content. Alternatively, the

329

leftwards peak shifts at ~ 2.87 Å in the PDFs are not derived from the decrease of the layer

330

Mn(III) content. The shift could be caused by decreasing particle sizes that lead to structural

331

contraction.12 Regardless of the mechanism leading to the PDF peak shift, we do not observe

332

significant accumulation of Mn(III) in the structure of δ-MnO2 at either pH 4 or 6 during FA

333

interaction with δ-MnO2.

334

The low Mn(III) content in δ-MnO2 structure at pH 4 and 6 contrast to the results of the

335

previous studies under oxic conditions based on XPS analyses, showing 34% - 56% Mn(III) after

336

δ-MnO2 reacted with humate and oxalate at pH ~ 5.2.16,

337

exceptionally strong Mn(III)-humate or Mn(III)-oxalate complexes are proposed to form on the

338

solid surface, inhibiting further reduction of Mn(III) to Mn(II).16, 17 However, both humate and

339

oxalate are weak ligands, and their complexes with Mn(III) are not strong enough to be stable at

340

acidic pH.53 Indeed, even the highly stable Mn(III)-Desferrioxamine B complexes exist only at

341

alkaline pH.54, 55

342

17

In those previous studies,

The previously observed high Mn(III) fractions16-18 are more likely due to the presence of

15 ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 31

343

O2, formation of low valence of Mn phases, uncertainties in Mn(III) quantification, and/or

344

properties of starting materials. Oxidation of adsorbed Mn(II) on birnessite surfaces by

345

atmospheric O2, and formation of MnOOH and Mn3O4 can both increase Mn(III) concentration.

346

The XPS analysis based on Mn 2p3/2 spectra and its probing chemical composition of surface

347

structures, not the bulk structure, may lead to overestimation of Mn(III) fractions.56, 57 In contrast,

348

a more recent work on the oxidation of bisphenol A by Mn(III)-rich δ-MnO2 at pH 7 under oxic

349

conditions shows essentially no accumulation of either bulk or surface Mn(III) using XANES

350

spectroscopy and more accurate XPS analyses based on Mn 3p spectra.4 No Mn(III)

351

accumulation in the previous study4 could be caused by the already high Mn(III) content (31%)

352

in the starting δ-MnO2 material that cannot accommodate additional Mn(III), and/or by pH 7 that

353

is not high enough to cause significant Mn(III) accumulation.

354

Formation of MnOOH and Hausmannite. The transformation product of δ-MnO2

355

depends on the molar ratio of produced Mn(II) to δ-MnO2. High Mn(II)/MnO2 ratios result in

356

transformation of birnessite to MnOOH and hausmannite (Mn3O4),1, 29, 58 consistent with the

357

observations in the present work (Figure 3). The formation of MnOOH needs a lower

358

Mn(II)/MnO2 ratio than the formation of Mn3O4,29 accounting for the observed earlier formation

359

of MnOOH. However, groutite and feitknechtite are metastable and can convert into

360

hausmannite in the presence of Mn(II)(aq),29, 59 or they are dissolved by FA reduction (e.g., at R =

361

10), both leading to their disappearance with time.

362

Groutite and hausmannite also form at R = 1 and pH 6 (Figure 3) where the molar ratio of

363

[Mn]dis to the remaining Mn(IV) is 0.03 and 0.05 – 0.5, respectively. However, Lefkowitz et al.29

364

find that hausmannite forms at pH 8 but not at pH < 7 during the reaction of [Mn]dis (0.2 – 2.2

365

mM) with acid birnessite (0.05 g/L) ([Mn(II)]/[Mn(IV)] = ~ 0.35 – 3.8) under anoxic conditions.

16 ACS Paragon Plus Environment

Page 17 of 31

Environmental Science & Technology

366

The difference could be caused by the different Mn(II)/Mn(IV) ratios in the two studies or the

367

lower reactivity of acid birnessite than δ-MnO2.26 The adsorbed FA may also affect the

368

transformation products.

369

Environmental Implications. Both birnessite and NOM are common compounds in

370

suboxic sediments.1,

2, 60

371

concentration, and accordingly decrease birnessite metal sorption and oxidizing reactivity and

372

change metal sorption behavior.20, 21, 23 Our work shows that pH and FA/birnessite mass ratios

373

strongly affect both the content and location of the accumulated Mn(III) in birnessite structure,

374

which was not revealed previously.4, 16-18 Results indicate that Mn(III) accumulates much more

375

both in layers and on vacancies with less vacancy-adsorbed Mn(II) under alkaline than acidic

376

conditions. Such detailed information is important because Mn(III) adsorbs more strongly than

377

Mn(II) and the presence of abundant Mn(III) on vacancies can strongly decrease birnessite

378

sorption capacities for metals and force the metals to adsorb on birnessite edge sites.20, 21 Thus,

379

the finding suggests that the decline of birnessite reactivity by partial reduction is more severe in

380

alkaline environment (e.g. ocean sediments) than in acidic environment (e.g., acid mine

381

drainage). On the other hand, more Mn(III) accumulated in birnessite in alkaline environment is

382

more favorable to birnessite transformation to todorokite as Mn(III) is required for the

383

transformation.22 Our work for the first time shows that high NOM/birnessite ratios lead to not

384

only Mn(III)-rich birnessite but also Mn3O4 and MnOOH phases around circumneutral pH.

385

These low-valence Mn oxides can contribute to the decline of birnessite oxidizing reactivity

386

during reduction, and the reduction-induced formation from birnessite is a new formation

387

pathway of these mineral phases in natural environment. Our study implies that Mn oxides co-

388

existing with NOM in suboxic environment, either as Mn(III,II)-rich birnessite or the low-

Partial reduction of birnessite by NOM can increase Mn(III)

17 ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 31

389

valence Mn oxides, have low metal adsorption and oxidizing reactivity. This work also shows

390

that the altered birnessite due to partial reduction has a high NOM adsorption capacity because

391

adsorbed Mn(II, III) makes the birnessite surface less negative and favors NOM adsorption via

392

cation bridging. Thus, birnessite not only can change quality of organic carbon but also is a

393

potent adsorbent for organic carbon.

394

Dynamic mineralogical and Mn oxidation state changes of Mn oxides during

395

microbially-mediated oxidative precipitation have been ascribed to the Mn(II)-birnessite redox

396

reactions.15, 61, 62 Birnessite partial reduction by FA leads to similar dynamic changes, which can

397

also be well explained by the redox reactions. Such reaction mechanisms depend less on the type

398

of the reductant, thus the findings in this study are likely applicable to reduction of birnessite by

399

other reducing substances, including both organics and inorganics, although the extent and rate

400

of the changes during reduction may depend on the type of reductants. For instance, NOM

401

occurring in sediments are more reducing than surface-environment extracted FA used in this

402

work,60, 63 and thus faster reduction and phase transformation of birnessite in sediments may

403

occur.

404

ASSOCIATED CONTENT

405

Supporting Information.

406

Synthesis of δ-MnO2, acidification experiments, Mn XANES spectra and LCF fits,

407

organic matter characterization, and some of the XRD and PDF data are available free of charge

408

at http://pubs.acs.org.

409

AUTHOR INFORMATION

410

Corresponding Author

411

*M. Zhu. E-mail: [email protected]; Phone: 307-766-5523

18 ACS Paragon Plus Environment

Page 19 of 31

412 413 414

Environmental Science & Technology

Notes The authors declare no competing financial interest. ACKNOWLEDGMENTS

415

This work was supported by the U.S. National Science Foundation under Grant

416

EAR-1529937. We acknowledge beamline scientists Olaf J. Borkiewicz, Kevin A. Beyer and

417

Karena W. Chapman at beamline 11-ID-B and John Katsoudas and Carlo Segre at beamline 10-

418

BM-B at APS for their technical assistance in data collection. This research used resources of the

419

Advanced Photon Source, a U.S. Department of Energy (DOE) Office of Science User Facility

420

operated for the DOE Office of Science by Argonne National Laboratory under Contract No.

421

DE-AC02-06CH11357. We thank four anonymous reviewers and Associate Editor T. David

422

Waite for constructive comments and suggestions.

19 ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 31

423

REFERENCES

424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467

(1) Elzinga, E. J. Reductive transformation of birnessite by aqueous Mn(II). Environ. Sci. Technol. 2011, 45 (15), 6366-6372. (2) Post, J. E. Manganese oxide minerals: Crystal structures and economic and environmental significance. Proc. Nat. Acad. Sci. U. S. A. 1999, 96, 3447-3454. (3) Ginder-Vogel, M.; Landrot, G.; Fischel, J. S.; Sparks, D. L. Quantification of rapid environmental redox processes with quick-scanning x-ray absorption spectroscopy (Q-XAS). Proc. Nat. Acad. Sci. U. S. A. 2009, 106 (38), 16124-16128. (4) Balgooyen, S.; Alaimo, P. J.; Remucal, C. K.; Ginder-Vogel, M. Structural transformation of MnO2 during the oxidation of bisphenol A. Environ. Sci. Technol. 2017, 51 (11), 6053-6062. (5) Ehlert, K.; Mikutta, C.; Kretzschmar, R. Impact of birnessite on arsenic and iron speciation during mcrobial reduction of arsenic-bearing ferrihydrite. Environ. Sci. Technol. 2014, 48 (19), 11320-11329. (6) Zhu, M.; Ginder-Vogel, M.; Parikh, S. J.; Feng, X.-H.; Sparks, D. L. Cation effects on the layer structure of biogenic Mn-oxides. Environ. Sci. Technol. 2010, 44, 4465-4471. (7) Pan, C.; Liu, H.; Catalano, J. G.; Qian, A.; Wang, Z.; Giammar, D. E. Rates of Cr(VI) generation from CrxFe1–x(OH)3 solids upon reaction with manganese oxide. Environ. Sci. Technol. 2017, 51 (21), 12416-12423. (8) Sunda, W. G.; Huntsman, S. A.; Harvey, G. R. Photoreduction of manganese oxides in seawater and its geochemical and biological implications. Nature 1983, 301 (5897), 234-236. (9) Sunda, W. G.; Kieber, D. J. Oxidation of humic substances by manganese oxides yields low-molecular-weight organic substrates. Nature 1994, 367 (6458), 62-64. (10) Lin, H.; Taillefert, M. Key geochemical factors regulating Mn(IV)-catalyzed anaerobic nitrification in coastal marine sediments. Geochim. Cosmochim. Acta 2014, 133, 17-33. (11) Tebo, B. M.; Johnson, H. A.; McCarthy, J. K.; Templeton, A. S. Geomicrobiology of manganese(II) oxidation. Trends Microbiol. 2005, 13 (9), 421-428. (12) Wang, Q.; Liao, X.; Xu, W.; Ren, Y.; Livi, K. J.; Zhu, M. Synthesis of birnessite in the presence of phosphate, silicate, or sulfate. Inorg. Chem. 2016, 55 (20), 10248-10258. (13) Grangeon, S.; Fernandez-Martinez, A.; Warmont, F.; Gloter, A.; Marty, N.; Poulain, A.; Lanson, B. Cryptomelane formation from nanocrystalline vernadite precursor: a high energy Xray scattering and transmission electron microscopy perspective on reaction mechanisms. Geochem. Trans. 2015, 16 (1), 12. (14) Webb, S. M.; Tebo, B. M.; Bargar, J. R. Structural characterization of biogenic Mn oxides produced in seawater by the marine bacillus sp. strain SG-1. Am. Mineral. 2005, 90 (8-9), 1342-1357. (15) 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. (16) Banerjee, D.; Nesbitt, H. W. XPS study of dissolution of birnessite by humate with constraints on reaction mechanism. Geochim. Cosmochim. Acta 2001, 65 (11), 1703-1714. (17) Banerjee, D.; Nesbitt, H. W. XPS study of reductive dissolution of birnessite by oxalate: rates and mechanistic aspects of dissolution and redox processes. Geochim. Cosmochim. Acta 1999, 63 (19–20), 3025-3038. (18) Shaikh, N.; Taujale, S.; Zhang, H.; Artyushkova, K.; Ali, A.-M. S.; Cerrato, J. M. 20 ACS Paragon Plus Environment

Page 21 of 31

468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513

Environmental Science & Technology

Spectroscopic investigation of interfacial interaction of manganese oxide with triclosan, aniline, and phenol. Environ. Sci. Technol. 2016, 50 (20), 10978-10987. (19) Jokic, A.; Frenkel, A. I.; Vairavamurthy, M. A.; Huang, P. M. Birnessite catalysis of the Maillard reaction: Its significance in natural humification. Geophys. Res. Lett. 2001, 28 (20), 3899-3902. (20) Simanova, A. A.; Kwon, K. D.; Bone, S. E.; Bargar, J. R.; Refson, K.; Sposito, G.; Peña, J. Probing the sorption reactivity of the edge surfaces in birnessite nanoparticles using nickel(II). Geochim. Cosmochim. Acta 2015, 164, 191-204. (21) Hinkle, M. A. G.; Dye, K. G.; Catalano, J. G. Impact of Mn(II)-manganese oxide reactions on Ni and Zn speciation. Environ. Sci. Technol. 2017, 51 (6), 3187-3196. (22) Feng, X. H.; Zhu, M.; Ginder-Vogel, M.; Ni, C.; Parikh, S. J.; Sparks, D. L. Formation of nano-crystalline todorokite from biogenic Mn oxides. Geochim. Cosmochim. Acta 2010, 74 (11), 3232-3245. (23) Lafferty, B. J.; Ginder-Vogel, M.; Zhu, M.; Livi, K. J. T.; Sparks, D. L. Arsenite oxidation by a poorly crystalline manganese-oxide. 2. Results from X-ray absorption spectroscopy and X-ray diffraction. Environ. Sci. Technol. 2010, 44, 8467-8472. (24) Waite, T. D.; Wrigley, I. C.; Szymczak, R. Photoassisted dissolution of a colloidal manganese oxide in the presence of fulvic acid. Environ. Sci. Technol. 1988, 22 (7), 778-785. (25) Hinkle, M. A. G.; Flynn, E. D.; Catalano, J. G. Structural response of phyllomanganates to wet aging and aqueous Mn(II). Geochim. Cosmochim. Acta 2016, 192, 220-234. (26) Zhao, H.; Zhu, M.; Li, W.; Elzinga, E. J.; Villalobos, M.; Liu, F.; Zhang, J.; Feng, X.; Sparks, D. L. Redox reactions between Mn(II) and hexagonal birnessite change its layer symmetry. Environ. Sci. Technol. 2016, 50, 1750-1758. (27) Elzinga, E. J.; Kustka, A. B. A Mn-54 radiotracer study of Mn isotope solid–liquid exchange during reductive transformation of Vernadite (δ-MnO2) by aqueous Mn(II). Environ. Sci. Technol. 2015, 49 (7), 4310-4316. (28) Elzinga, E. J. 54Mn radiotracers demonstrate continuous dissolution and reprecipitation of vernadite (δ-MnO2) during interaction with aqueous Mn (II). Environ. Sci. Technol. 2016, 50, 8670-8677. (29) Lefkowitz, J. P.; Rouff, A. A.; Elzinga, E. J. Influence of pH on the reductive transformation of birnessite by aqueous Mn(II). Environ. Sci. Technol. 2013, 47 (18), 1036410371. (30) Mandernack, K. W.; Post, J.; Tebo, B. M. Manganese mineral formation by bacterial spores of the marine Bacillus, strain SG-1: Evidence for the direct oxidation of Mn(II) to Mn(IV). Geochim. Cosmochim. Acta 1995, 59 (21), 4393-4408. (31) Tu, S.; Racz, G. J.; Goh, T. B. Transformations of synthetic birnessite as affected by pH and manganese concentration. Clays Clay Miner. 1994, 42 (3), 321-330. (32) Nordstrom, D. K.; Alpers, C. N.; Ptacek, C. J.; Blowes, D. W. Negative pH and extremely acidic mine waters from Iron Mountain, California. Environ. Sci. Technol. 2000, 34 (2), 254-258. (33) Sánchez-Andrea, I.; Knittel, K.; Amann, R.; Amils, R.; Sanz, J. L. Quantification of Tinto River sediment microbial communities: importance of sulfate-reducing bacteria and their role in attenuating acid mine drainage. Appl. Environ. Microbiol. 2012, 78 (13), 4638-4645. (34) Simanova, A. A.; Peña, J. Time-resolved investigation of cobalt oxidation by Mn(III)rich δ-MnO2 using quick X-ray absorption spectroscopy. Environ. Sci. Technol. 2015, 49 (18), 10867-10876. 21 ACS Paragon Plus Environment

Environmental Science & Technology

514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559

Page 22 of 31

(35) Her, N.; Amy, G.; Foss, D.; Cho, J. Variations of molecular weight estimation by HPSize exclusion chromatography with UVA versus online DOC detection. Environ. Sci. Technol. 2002, 36 (15), 3393-3399. (36) Stone, A. T.; Morgan, J. J. Reduction and dissolution of manganese(III) and manganese(IV) oxides by organics: 2. Survey of the reactivity of organics. Environ. Sci. Technol. 1984, 18 (8), 617-624. (37) Manceau, A.; Marcus, M. A.; Grangeon, S. Determination of Mn valence states in mixedvalent manganates by XANES spectroscopy. Am. Mineral. 2012, 97 (5-6), 816-827. (38) Perez-Benito, J. F.; Brillas, E.; Pouplana, R. Identification of a soluble form of colloidal manganese(IV). Inorg. Chem. 1989, 28 (3), 390-392. (39) Luther, G., III; Popp, J. Kinetics of the abiotic reduction of polymeric manganese dioxide by nitrite: an anaerobic nitrification reaction. Aquat. Geochem. 2002, 8 (1), 15-36. (40) Zhu, M.; Farrow, C. L.; Post, J. E.; Livi, K. J. T.; Billinge, S. J. L.; Ginder-Vogel, M.; Sparks, D. L. Structural study of biotic and abiotic poorly-crystalline manganese oxides using atomic pair distribution function analysis. Geochim. Cosmochim. Acta 2012, 81, 39-55. (41) Grangeon, S.; Lanson, B.; Miyata, N.; Tani, Y.; Manceau, A. Structure of nanocrystalline phyllomanganates produced by freshwater fungi. Am. Mineral. 2010, 95 (11-12), 1608-1616. (42) Allard, S.; Gutierrez, L.; Fontaine, C.; Croué, J.-P.; Gallard, H. Organic matter interactions with natural manganese oxide and synthetic birnessite. Sci. Total Environ. 2017, 583, 487-495. (43) Chorover, J.; Amistadi, M. K. Reaction of forest floor organic matter at goethite, birnessite and smectite surfaces. Geochim. Cosmochim. Acta 2001, 65 (1), 95-109. (44) Grangeon, S.; Lanson, B.; Lanson, M.; Manceau, A. Crystal structure of Ni-sorbed synthetic vernadite: a powder X-ray diffraction study. Mineral. Mag. 2008, 72 (6), 1279-1291. (45) Drits, V.; Lanson, B.; Gaillot, A.-C. Birnessite polytype systematics and identification by powder X-ray diffraction. Am. Mineral. 2007, 92 (5-6), 771 - 788. (46) Gu, B.; Schmitt, J.; Chen, Z.; Liang, L.; McCarthy, J. F. Adsorption and desorption of natural organic matter on iron oxide: mechanisms and models. Environ. Sci. Technol. 1994, 28 (1), 38-46. (47) Taujale, S.; Baratta, L. R.; Huang, J.; Zhang, H. Interactions in ternary mixtures of MnO2, Al2O3, and natural organic matter (NOM) and the impact on MnO2 oxidative reactivity. Environ. Sci. Technol. 2016, 50 (5), 2345-2353. (48) Feng, X. H.; Zhai, L. M.; Tan, W. F.; Liu, F.; He, J. Z. Adsorption and redox reactions of heavy metals on synthesized Mn oxide minerals. Environ. Pollut. 2007, 147 (2), 366-373. (49) Lin, K.; Liu, W.; Gan, J. Oxidative removal of Bisphenol A by manganese dioxide: Efficacy, products, and pathways. Environ. Sci. Technol. 2009, 43 (10), 3860-3864. (50) Nico, P. S.; Anastasio, C.; Zasoski, R. J. Rapid photo-oxidation of Mn(II) mediated by humic substances. Geochim. Cosmochim. Acta 2002, 66 (23), 4047-4056. (51) Grangeon, S.; Manceau, A.; Guilhermet, J.; Gaillot, A.-C.; Lanson, M.; Lanson, B. Zn sorption modifies dynamically the layer and interlayer structure of vernadite. Geochim. Cosmochim. Acta 2012, 85, 302-313. (52) Grangeon, S.; Fernandez-Martinez, A.; Claret, F.; Marty, N.; Tournassat, C.; Warmont, F.; Gloter, A. In-situ determination of the kinetics and mechanisms of nickel adsorption by nanocrystalline vernadite. Chem. Geol. 2017, 459, 24-31. (53) Chen, W.-R.; Liu, C.; Boyd, S. A.; Teppen, B. J.; Li, H. Reduction of carbadox mediated by reaction of Mn(III) with oxalic acid. Environ. Sci. Technol. 2013, 47 (3), 1357-1364. 22 ACS Paragon Plus Environment

Page 23 of 31

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

Environmental Science & Technology

(54) Klewicki, J. K.; Morgan, J. J. Kinetic behavior of Mn(III) complexes of pyrophosphate, EDTA, and citrate. Environ. Sci. Technol. 1998, 32 (19), 2916-2922. (55) Duckworth, O. W.; Sposito, G. Siderophore-manganese(III) interactions. I. Air-oxidation of manganese(ll) promoted by desferrioxamine B. Environ. Sci. Technol. 2005, 39 (16), 6037-44. (56) Ilton, E. S.; Post, J. E.; Heaney, P. J.; Ling, F. T.; Kerisit, S. N. XPS determination of Mn oxidation states in Mn (hydr)oxides. Appl. Surf. Sci. 2016, 366, 475-485. (57) Yin, H.; Tan, W.; Zheng, L.; Cui, H.; Qiu, G.; Liu, F.; Feng, X. Characterization of Nirich hexagonal birnessite and its geochemical effects on aqueous Pb2+/Zn2+ and As(III). Geochim. Cosmochim. Acta 2012, 93 (0), 47-62. (58) Lefkowitz, J. P.; Elzinga, E. J. Impacts of aqueous Mn(II) on the sorption of Zn(II) by hexagonal birnessite. Environ. Sci. Technol. 2015, 49 (8), 4886-4893. (59) Colón, W. C. E. In situ pE-pH analysis on the oxidation kinetics of manganese bearing solution. M. S. Thesis, The Pennsylvania State University, University Park, PA, 2014. (60) Canfield, D. E.; Thamdrup, B.; Hansen, J. W. The anaerobic degradation of organic matter in Danish coastal sediments: Iron reduction, manganese reduction, and sulfate reduction. Geochim. Cosmochim. Acta 1993, 57 (16), 3867-3883. (61) Webb, S. M.; Dick, G. J.; Bargar, J. R.; Tebo, B. M. Evidence for the Presence of Mn(III) Intermediates in the Bacterial Oxidation of Mn(II). Proc. Nat. Acad. Sci. U. S. A. 2005, 102, 5558-5563. (62) 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. Am. Mineral. 2005, 90 (1), 143-154. (63) Segarra, K. E. A.; Comerford, C.; Slaughter, J.; Joye, S. B. Impact of electron acceptor availability on the anaerobic oxidation of methane in coastal freshwater and brackish wetland sediments. Geochim. Cosmochim. Acta 2013, 115, 15-30.

585

23 ACS Paragon Plus Environment

Environmental Science & Technology

Page 24 of 31

586 6 (A)

pH 4

4

R = 0.1

2 0

(mM)

dis [Mn]

R=1

0.5

R = 0.1 R = 0.01

0.0 (C)

0.1 Log scale

pH 6

(B)

1.0

R = 10

pH 8 R=1

0.01 1E-3

R = 0.1 R = 0.01

1E-4 1E-5 0

587 588 589 590 591 592 593 594 595 596

100

200

300

600

Time (h)

Figure 1. Dissolved Mn concentration ([Mn]dis) during reduction of δ-MnO2 by fulvic acid (FA) at pH 4, 6 and 8 and FA/solid mass ratios (R) of 0.01 - 10. (A) pH 4, (B) pH 6, and (C) pH 8. Note that the logarithmic scale is used for [Mn]dis in panel C. The error bars result from repeated concentration measurements using ICP-OES.

24 ACS Paragon Plus Environment

Page 25 of 31

Environmental Science & Technology

150

(A)

(C) R=1

140

R=1

3.6

pH 8

pH 8

3.5 pH 6

3.4 pH 6

SUVA254

DOC (mg/L)

130

120 (B)

150

R = 0.1

3.3 4.0

(D) R = 0.1

pH 8

pH 8

3.2 pH 6

100 pH 6

2.4

pH 4

pH 4

50

1.6 0

597 598 599 600 601 602

100

200

300

600

0

100

200

300

600

Time (h)

Time (h)

Figure 2. Dissolved organic carbon (DOC) (A, B) and SUVA254 (C, D) of fulvic acid (FA) remained in the solution during reduction of δ-MnO2 by FA at pH 4, 6 and 8 and FA/solid mass ratios (R) of 0.1 and 1. The error bars for DOC and SUVA254 result from repeated concentration measurements using a TOC analyzer and UV-vis spectroscopy.

25 ACS Paragon Plus Environment

Environmental Science & Technology

(A)

(B)

pH 4, R = 0.1

Page 26 of 31

(C)

pH 6, R = 0.1

pH 6, R = 1

(11, 20) (31, 20)

H

(002) (22, 40) 600 h

600 h

600 h

120 h

240 h

72 h 24 h

24 h

1h

5h

120 h G

24 h 10 h

Intensity (a. u.)

1h 0.02 h 0h

(D)

(E)

pH 8, R = 0.1

1h

0.02 h

0.02 h

0h

0h

pH 8, R = 1

(F) H

600 h 600 h

5h

5h

0.02 h

0h

605 606 607 608 609 610 611

0.02 h

1h

0.02 h

d spacing (Å)

24 h

24 h

24 h

603 604

H G

120 h

120 h

2

FG

240 h F

360 h

86 4

pH 8, R = 10

H

H

0h

0h

86 4

2

d spacing (Å)

86 4

2

d spacing (Å)

Figure 3. Synchrotron-based X-ray diffraction (XRD) patterns of δ-MnO2 reacted with fulvic acid (FA) at pH 4, 6 and 8 and FA/solid mass ratios (R) of 0.1, 1 and 10. For each condition, XRD patterns are normalized by adjusting (11, 20) peak height and superimposed with the patterns of the 0-h sample. The arrow indicates the position of the dip at ~ 1.97 Å. The labels of H, G and F represent hausmannite, groutite and feitknechtite, respectively.

26 ACS Paragon Plus Environment

Page 27 of 31

Environmental Science & Technology

(A)

(Mn-O)

2.857 Å

(Mn-Mn)L

pH 4, R = 0.1 (MnL-MnIL)

5.318 Å

600 h 24 h 1h 0.02 h 0h

(B)

(Mn-O)

2.858 Å

(Mn-Mn)L

pH 6, R = 0.1

G(r) (Å ) (a.u.)

(MnL-MnIL)

5.308 Å

600 h

-2

240 h 72 h 24 h 5h 0.5 h 0.02 h 0h

(C)

(Mn-O)

2.876 Å

(Mn-Mn)L

pH 8, R = 0.1 (MnL-MnIL) 5.272 Å 600 h 240 h 72 h 24 h 5h 1h 0.02 h 0h

1.6

612 613 614 615 616 617 618

2.4

3.2

4.0

4.8

5.6

r (Å)

Figure 4. X-ray atomic pair distribution functions within 6 Å for δ-MnO2 reacted with fulvic acid (FA) at FA/solid mass ratio (R) of 0.1 and pH 4 (A), 6 (B) or 8 (C). The PDF data are superimposed with the 0-h sample at each pH (black). The highlighted peaks show clear shifts with increasing reaction time under each condition, and the number is the peak position of the 600-h sample. The data for other conditions are provided in Figure S8.

27 ACS Paragon Plus Environment

Environmental Science & Technology

2.885

0.30

(A)

(B) pH 8, R = 1

pH 8, R = 1

2.880

0.25 pH 6, R = 1

pH 8, R = 0.1

2.875

Area

Distance (Å)

Page 28 of 31

2.870

0.20 pH 4, R = 0.1

0.15

pH 6, R = 0.1

0.10

pH 8, R = 0.1

pH 6, R = 1

2.865 pH 6, R = 0.1

2.860

pH 4, R = 0.1

0

619 620 621 622 623 624 625

50

0.05 100

150

200

250 600

0

50

Time (h)

100

150

200

250 600

Time (h)

Figure 5. The nearest (Mn-Mn)L interatomic distances (A) and the area of the MnL-MnIL peak located at ~ 5.27 Å (B) derived from the atomic pair distribution functions for δ-MnO2 reacted with fulvic acid (FA) at 4, 6 or 8 and FA/solid mass ratios (R) of 0.1 or 1.

28 ACS Paragon Plus Environment

Page 29 of 31

Environmental Science & Technology

(A)

80

Mn(II) Mn(III) Mn(IV)

60

Red: pH 4 Black: pH 6

40

80 60

Mn(II) Red: R = 0.1 Mn(III) Black: R = 1 Mn(IV)

40

0

0 0

627 628 629

pH 8, R = 0.1 and 1

20

20

626

(B)

100

pH 4 and 6, R = 0.1

Mn Fractions (%)

Mn Fractions (%)

100

100

200

300

400 600

0

100

200

300

400 600

Time (h)

Time (h)

Figure 6. Molar fractions of Mn(II), Mn(III) and Mn(IV) during δ-MnO2 reduction by fulvic acid at pH 4, 6 or 8 and FA/solid mass ratios (R) of 1 and 0.1, determined from Mn K-edge XANES linear combination fitting analysis. (A) R = 0.1, pH 4 and 6. (B) pH 8, R = 1 and 0.1.

29 ACS Paragon Plus Environment

Environmental Science & Technology

Page 30 of 31

Electron transfer (mM)

18 16

pH 4, R = 0.1

14 12 10

pH 6, R = 0.1

8 6

pH 8, R = 0.1

4 pH 8, R = 1

2 0 0

630 631 632 633

100

200

300

400 600

Time (h)

Figure 7. The amount of net electron transfer from 0.3 g/L fulvic acid (FA) to 3 g/L (R = 0.1) or 0.3 g/L (R = 1) δ-MnO2 during reduction of δ-MnO2 by FA at pH 4, 6 or 8.

30 ACS Paragon Plus Environment

Page 31 of 31

634

Environmental Science & Technology

TOC Art

635

636

31 ACS Paragon Plus Environment