Metal Adsorption Controls Stability of Layered Manganese Oxides

May 22, 2019 - Hexagonal birnessite, a typical layered Mn oxide (LMO), can adsorb and oxidize Mn(II) and thereby transform to Mn(III)-rich hexagonal ...
0 downloads 0 Views 1MB Size
Subscriber access provided by BOSTON UNIV

Environmental Processes

Metal Adsorption Controls Stability of Layered Manganese Oxides Peng Yang, Jeffrey E. Post, Qian Wang, Wenqian Xu, Roy H. Geiss, Patrick R McCurdy, and Mengqiang Zhu Environ. Sci. Technol., Just Accepted Manuscript • Publication Date (Web): 22 May 2019 Downloaded from http://pubs.acs.org on May 28, 2019

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 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 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.

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 35

Environmental Science & Technology

1

Metal Adsorption Controls Stability of Layered Manganese Oxides

2 3

Peng Yang,† Jeffrey E. Post,‡ Qian Wang,† Wenqian Xu,§ Roy Geiss,∥ Patrick R. McCurdy,∥ and

4

Mengqiang Zhu†,*

5 6

†Department

7

Wyoming 82071, United States

8

‡Department

9

20013, United States

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

of Mineral Sciences, Smithsonian Institution, Washington, District of Columbia

10

§X-ray

11

Illinois 60439, United States

12

∥Department of Chemistry, Colorado State University, Fort Collins, Colorado 80523, United States

Science Division, Advanced Photon Source, Argonne National Laboratory, Lemont,

13 14

*Corresponding author: E-mail: [email protected]

ACS Paragon Plus Environment

1

Environmental Science & Technology

15

Page 2 of 35

Abstract

16

Hexagonal birnessite, a typical layered Mn oxide (LMO), can adsorb and oxidize Mn(II)

17

and thereby transform to Mn(III)-rich hexagonal birnessite, triclinic birnessite, and/or tunneled Mn

18

oxides (TMOs), remarkably changing the environmental behavior of Mn oxides. We have

19

determined the effects of co-existing cations on the transformation by incubating Mn(II)-bearing

20

δ-MnO2 at pH 8 under anoxic conditions for 25 d (dissolved Mn < 11 μM). In the Li+, Na+ or K+

21

chloride solution, the Mn(II)-bearing δ-MnO2 first transforms to Mn(III)-rich -MnO2 and/or

22

triclinic birnessite (T-bir) due to the Mn(II)-Mn(IV) comproportionation, most of which eventually

23

transform to a 4  4 TMO. In contrast, Mn(III)-rich -MnO2 and T-bir form and persist in the Mg2+

24

or Ca2+ chloride solution. However, in the presence of surface adsorbed Cu(II), Mn(II)-bearing

25

δ-MnO2 turns into Mn(III)-rich -MnO2 without forming T-bir or TMOs. The stabilizing power of

26

the cations on the -MnO2 structure positively correlates with their binding strength to -MnO2

27

(Li+, Na+ or K+ < Mg2+ or Ca2+ < Cu(II)). Since metal adsorption decreases the surface energy of

28

minerals, our finding suggests that the surface energy largely controls the thermodynamic stability

29

of LMOs. Our study indicates that the adsorption of divalent metal cations, particularly transition

30

metals, can be an important cause of the high abundance of LMOs, rather than the more stable

31

TMO phases, in the environment.

ACS Paragon Plus Environment

2

Page 3 of 35

32

Environmental Science & Technology

INTRODUCTION

33

Manganese (Mn) oxides are ubiquitous in soils, sediments and ocean nodules.1 They are

34

metal scavengers and strong oxidants,2-5 thereby affecting the fate of metals, nutrients and organic

35

compounds in the environment.5-13 The most common naturally-occurring Mn oxide is birnessite,1

36

a layered Mn oxide (LMO) consisting of stacked layers constructed by edge-sharing MnO6

37

octahedra. Mn(IV) is dominant in MnO6 layers, but a portion of Mn sites are vacant or substituted

38

by Mn(III).1,14,15 Mn(III) affects the transformation of birnessite to tunneled Mn oxides (TMOs),16-

39

18

40

than birnessite. Both Mn(III) and vacancies can also strongly affect birnessite metal sorption

41

properties, oxidizing activity, and bandgap energies pertinent to photochemical reduction of

42

birnessite.19-24

another family of common naturally-occurring Mn oxides but more thermodynamically stable

43

Mn(II) often encounters birnessite, such as during oxidative precipitation of Mn(II),25,26

44

partial reductive dissolution of birnessite,6,13 and in many other environmental scenarios.27,28 In

45

these situations, inevitable adsorption and oxidation of Mn(II) by birnessite can increase Mn(III)

46

but decrease vacancy concentration of birnessite, or even transform birnessite to other Mn oxide

47

phases.29-33 The oxidation of Mn(II) by Mn(IV) of birnessite is also called comproportionation.

48

The products of the Mn(II)-birnessite reactions depend on the Mn(II)/MnO2 ratios and pH.25,29-33

49

At relatively low Mn(II)/MnO2 ratios (0.05 – 0.24) and pH ≥ 8, the reactions increase Mn(III)

50

concentration but decrease vacancy concentration in the layers of hexagonal birnessite, eventually

51

leading to formation of triclinic birnessite (T-bir) that contains 1/3 Mn as Mn(III) without

52

vacancies.32 The incubation of Mn(II)-bearing δ-MnO2 at pH 6 – 8 and 21 oC leads to surprisingly

53

rapid formation of 4 × 4 TMO with T-bir or Mn(III)-rich δ-MnO2 as intermediate products.16 Upon

54

aging at pH 4 with Mn(II)/MnO2 ratios of 0.04 – 0.54, birnessite rearranges into a superlattice

ACS Paragon Plus Environment

3

Environmental Science & Technology

Page 4 of 35

55

structure by rotated stacking or displacing the layers.13,34 In contrast, at relatively high

56

Mn(II)/MnO2 ratios (e.g., 0.30 – 3.04) and pH 7.0 – 8.5, birnessite transforms to feitknechtite

57

(-MnOOH), manganite (-MnOOH) and hausmannite (Mn3O4).25,29-31

58

The birnessite structure can host large numbers of metal cations due to its strong metal

59

adsorption capability for alkali, alkaline earth and transition metals.3-5,35-37 The strong metal

60

sorption likely affects the rate and products of Mn(II)-induced birnessite transformation. At high

61

Mn(II)/MnO2 ratios (0.43 – 1.04) and in the presence of Ni(II) and Zn(II), birnessite transforms

62

into Ni-substituted feitknechtite and Zn-substituted hausmannite at pH 7.5 – 8, respectively,36,38,39

63

and Ni substitution inhibits further transformation of feitknechtite to manganite.38 With an

64

Mn(II)/MnO2 ratio of 0.04 – 0.54, Ni(II) and Zn(II) mainly compete with Mn(II) for adsorption

65

without changing the mineral phase of birnessite at pH 4 and 7.40 However, the effects of metal

66

cations on the Mn(II)-birnessite reactions at relatively low Mn(II)/MnO2 ratios remain unknown,

67

although the low ratios are common in the natural environment.41,42

68

In the present study, we determined the effects of selected alkali (Li+, Na+, and K+), alkaline

69

earth (Mg2+ and Ca2+) and transition metal (Cu(II)) cations on the rate and products of

70

Mn(II)-induced birnessite transformations at Mn(II)/MnO2 ratios of ≤ 0.14, which are considered

71

as low Mn(II)/MnO2 ratios based on the Mn(II) adsorption loading.16 Mn(II)-bearing δ-MnO2 was

72

prepared first and then incubated in solutions of each alkali and alkaline earth metal cation at pH

73

8 in a N2/H2 atmosphere (95% N2 + 5% H2) for 25 d. In terms of Cu(II) effects, Mn(II)-bearing

74

δ-MnO2 with Cu(II) adsorbed on the surface was incubated in NaCl solution. The three categories

75

of cations were selected because they are common in the natural environment and their adsorption

76

behaviors on birnessite differ markedly. Li+, Na+ and K+ occupy the interlayers of birnessite by

77

forming outer-sphere complexes.43,44 Both Mg2+ and Ca2+ adsorb on birnessite probably as a

ACS Paragon Plus Environment

4

Page 5 of 35

Environmental Science & Technology

78

mixture of inner- and outer-sphere complexes.45,46 Cu(II) adsorbs on vacancies or edges as inner-

79

sphere complexes,12,37 enters the vacancies to be part of the layer,10,12 or forms polynuclear clusters

80

on edges of birnessite.37,47 The binding strength of the three categories of cations to birnessite

81

increases in the order of Li+, Na+, or K+ < Mg2+ or Ca2+ < Cu(II).48 The binding strength of the

82

cations within the alkali or alkaline earth group also differs. As metal adsorption affects energetics

83

of the adsorbent, we expect that the three categories of cations have distinct impacts on the

84

transformation of Mn(II)-bearing δ-MnO2. We also examined the effects of Na+ concentration on

85

the transformation of Mn(II)-bearing δ-MnO2. δ-MnO2 was chosen because it is a synthetic

86

analogue to widely spread vernadite and biogenic Mn oxides in the environment.49 The weakly

87

alkaline condition (pH 8) is prevalent in alkaline soils and marine environment. The N2/H2

88

atmosphere was used to mimic suboxic environments with a low O2 content and to avoid

89

interference of atmospheric O2 with Mn(II)-birnessite reactions.16,32 This study provides important

90

insights into how metal adsorption affects the stability of LMOs with respect to phase

91

transformation to more crystalline birnessite and tunneled structures.

92

MATERIALS AND METHODS

93

Materials. All chemicals were of A.C.S. reagent grade and used as received. Solutions

94

used in incubation experiments were prepared in an anaerobic chamber (Coy Vinyl-A, 5% H2 +

95

95% N2) with degassed deionized (DI) water (18.2 MΩ·cm). δ-MnO2 was prepared via KMnO4

96

reduction of by Mn(NO3)2 (Supporting Information, SI-1).

97

Preparation of Mn(II)-Bearing δ-MnO2. Mn(II)-bearing δ-MnO2 was prepared

98

according to our previous study.16 As-prepared -MnO2 (4 g/L) was equilibrated with 10 mM

99

MnCl2 at pH 4 in 100 mM NaCl solution for 2 h in air to reach the highest Mn(II) sorption loading

100

achievable at pH 4 (Mn(II)/MnO2 = 0.14).16 With such low pH and the short reaction time, minimal

ACS Paragon Plus Environment

5

Environmental Science & Technology

Page 6 of 35

101

oxidation occurred to adsorbed Mn(II) by either atmospheric O2 or Mn(IV) in -MnO2. A control

102

system was prepared using the same procedures but without adding Mn(II). The solids were

103

collected using vacuum filtration, rinsed with DI water acidified to pH 4 by HCl, and subsequently

104

used for the incubation experiments.

105

To examine the effects of Cu(II) on the Mn(II)-birnessite reactions, 7 mM CuCl2 and 5 mM

106

MnCl2 were equilibrated with 4 g/L -MnO2 at pH 4 for 2 h, resulting in a Mn(II)/MnO2 ratio of

107

0.08 and a Cu(II)/MnO2 ratio of 0.14. In another experiment, 4 g/L -MnO2 was equilibrated with

108

3 mM MnCl2 at pH 4 for 2 h, resulting in a similar Mn(II)/MnO2 molar ratio (0.08) as that in the

109

Cu system. The preloading of Cu(II) on -MnO2 was to minimize otherwise precipitation of Cu(II)

110

during incubation at pH 8 (see below) and to mimic naturally-occurring birnessite that usually

111

adsorbs transition metals in its structure.50 With the same Mn(II) loading, a comparison of the two

112

systems can reveal the effects of Cu(II) on the transformation of Mn(II)-bearing -MnO2.

113

Incubation of Mn(II)-Bearing δ-MnO2. The experimental conditions are summarized in

114

Table 1. The incubation procedures were the same as those in our previous study.16 All incubation

115

plastic bottles were wrapped with aluminum foil to prevent potential photoreduction of -MnO2.

116

The Mn(II)-bearing δ-MnO2 prepared with 10 mM Mn(II) was incubated in LiCl, NaCl, KCl,

117

MgCl2, or CaCl2 solutions in the anaerobic chamber to investigate the effects of these cations on

118

the Mn(II)-birnessite reactions. The ionic strength was 100 mM, controlled by 100 mM alkali

119

metal (Li+, Na+ or K+) chloride or 33.3 mM alkaline earth metal (Mg2+ or Ca2+) chloride.

120

Cu/Mn(II)-bearing -MnO2 and Mn(II)-bearing -MnO2 prepared with 3 mM Mn(II)

121

(3Mn) were incubated in 100 mM NaCl solution. As a control, -MnO2 equilibrated with 0 mM

122

Mn(II) (0Mn) was also incubated in 100 mM NaCl solution.

123

Mn(II)-bearing -MnO2 prepared with 10 mM Mn(II) was also incubated in 0 or 10 mM

ACS Paragon Plus Environment

6

Page 7 of 35

124

Environmental Science & Technology

NaCl solutions to determine the effects of NaCl concentration on Mn(II)-birnessite reactions.

125

The suspension pH was controlled at pH 8 by an automatic pH titrator (Metrohm 907) for

126

the first 2 h and by manual adjustment thereafter. The incubation lasted for 25 d, during which

127

suspension aliquots were collected at pre-determined time intervals and filtered. The obtained

128

solids were rinsed with 10 mL DI water of pH 8, dried in the anaerobic chamber, and ground for

129

the following characterization. The filtrates were acidified for measuring dissolved Mn

130

concentration using the formaldoxime colorimetric method51 whereas the dissolved Mn and Cu

131

concentrations in the Cu system were measured using inductively coupled plasma-optical emission

132

spectrometry. The concentrations of both metals were very low (Table S1), indicating negligible

133

release of the metals into solution during the incubations.

134

Solid Characterization. The obtained solids from the incubation experiments were subject

135

to X-ray diffraction (XRD), atomic pair distribution function (PDF), transmission electron

136

microscopy (TEM), Raman spectroscopic, and X-ray absorption spectroscopic analyses.

137

TEM images were collected on a FEI Tecnai G2 F20 200 kV (at the Department of Geology

138

and Geophysics, University of Wyoming) or a JEOL 2100F 200 kV (at the Department of

139

Chemistry, Colorado State University) transmission electron microscope. The XRD patterns and

140

the total X-ray scattering data for the PDF analysis were collected, respectively, using X-rays of

141

0.4521 Å at beamline 17-BM-B and 0.2114 Å at beamline 11-ID-B at the Advanced Photon Source

142

(APS), Argonne National Laboratory. Raman spectra were collected using a HORIBA LabRAM

143

HR Evolution Raman microscope in the Department of Mineral Sciences at the Smithsonian

144

Institution.

145

To estimate the fraction of each Mn oxidation state in the incubated solids, selected solids

146

were analyzed by linear combination fitting (LCF) analysis of Mn K-edge X-ray absorption near

ACS Paragon Plus Environment

7

Environmental Science & Technology

Page 8 of 35

147

edge structure (XANES) spectra. The XANES data were collected at beamline 10-BM-B at APS.

148

Cu K-edge extended X-ray absorption fine structure (EXAFS) spectra were collected at beamline

149

4-1 or 7-3 at the Stanford Synchrotron Radiation Lightsource (SSRL) to characterize the local

150

coordination structure of Cu in the Cu system. More details about the solid characterization are

151

provided in the supporting information (SI-2).

152

RESULTS AND DISCUSSION

153

Equilibrating -MnO2 with Mn(II) and/or Cu(II) solutions at pH 4 slightly changes the

154

structure and composition of -MnO2. The obtained Mn(II)-bearing or Mn(II)/Cu(II)-bearing

155

δ-MnO2 has a similar mineral phase and morphology to as-prepared δ-MnO2 according to the XRD

156

(Figures 1 and S1a) and HR-TEM analyses (Figures 2 and S2). Important changes pertinent to

157

Mn(II)/Cu(II) adsorption occur to the peak at 5.3 Å in the PDFs (Figures 3, S3 and S4) and the dip

158

at 2.0 Å in the XRD patterns (Figures 1 and S1a).52 The 5.3 Å peak corresponds to the atomic pairs

159

between the Mn and/or Cu adsorbed on a vacancy and the second nearest layer Mn atoms

160

surrounding the vacancy.6,16 The increased intensity of this PDF peak is consistent with adsorption

161

of Mn(II) and/or Cu(II) on vacancies after the equilibration. The adsorption of the metals on

162

vacancies also makes the dip at 2.0 Å in the XRD pattern more pronounced (Figures 1 and S1a).

163

In addition, the equilibration of -MnO2 with Mn(II) solutions slightly increases Mn(III)

164

percentages by 4 – 6% (Figure 4a) estimated by the XANES-LCF analysis (Table S2) and leads to

165

the formation of minor feitknechtite (Figures 1 and S1a). These changes of -MnO2 upon reaction

166

with Mn(II) and/or Cu(II) are consistent with our previous study.16

167

As a control system, the δ-MnO2 prepared by reacting 0 mM Mn(II) (i.e., 0Mn) at pH 4

168

was incubated at pH 8 for 25 d under anoxic conditions. Both the 2.0 Å XRD dip (Figure S1b) and

169

the 5.3 Å PDF peak (Figures 3b and S3a) become slightly weaker with increasing incubation time,

ACS Paragon Plus Environment

8

Page 9 of 35

Environmental Science & Technology

170

indicating decreasing numbers of Mn(II,III) adsorbed on vacancies,2,6,16,52 which is caused by

171

Mn(III) incorporation into vacancies.6,16 As shown and discussed below, Mn(II)-bearing δ-MnO2,

172

however, undergoes remarkable changes in mineral phases and Mn oxidation state composition

173

during incubation in solutions of various background electrolytes. The transformation kinetics and

174

products depend on the type of the metal cations, which can be ascribed to the different binding

175

strength of the cations to -MnO2.

176

Effects of Li+, Na+, and K+. Upon incubation at pH 8 for 25 d, Mn(II)-bearing δ-MnO2,

177

which was prepared by equilibrating 4 g/L -MnO2 with 10 mM Mn(II) at pH 4 for 2 h, transforms

178

into T-bir within 15 or 25 d and partially transforms into 4 × 4 TMO after 15 d in all Li+, Na+, and

179

K+ systems (Figure 1a); minor pyrochroite (Mn(OH)2) also forms in these systems (Figure 1a).

180

MnOOH and Mn3O4 phases were not observed. The formation of T-bir, indicated by the splitting

181

of the diffraction peaks around 2.4 and 1.4 Å and the arising of new peaks between (Figure 1a), is

182

due to the comproportionation between adsorbed Mn(II) and Mn(IV) of -MnO2 at alkaline pH

183

(i.e., pH 8),16 producing Mn(III) (Figure 4a) and decreasing Mn(II) and Mn(IV) concentrations

184

(Figures 4b and S5). Particle growth via oriented attachment between T-bir particles further

185

increases particle sizes, resulting in well crystallized T-bir.32 These observations are consistent

186

with our previous study conducted in NaCl solution,16 suggesting that the three monovalent cations

187

impose similar impacts on the transformation in spite of their different binding strength. However,

188

the three systems differ in the crystallinity, amount and formation rate of the transformation

189

products.

190

The crystallinity of T-bir decreases in the order of Li+ > Na+ > K+, indicated by the

191

decreasing sharpness of the characteristic XRD peaks of T-bir (Figure 1a). The high crystallinity

192

of Li+-exchanged T-bir was also observed previously.53 However, the amount of T-bir formed in

ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 35

193

the three systems does not follow that order. The Na+ system produces the largest amount of T-bir,

194

followed by the Li+ and then K+ systems according to the XRD patterns (Figure 1a). The amounts

195

of T-bir formed in those systems are positively proportional to the Mn(III) percentages (Figure 4a)

196

estimated by the XANES-LCF analysis (Figure S6 and Table S2) and to the amounts of the

197

decrease in the PDF peak area at 5.3 Å (Figures 3a, S3e,f, and S4a). However, the solids in the Li+

198

system still contain high numbers of Mn(III) (26% at 25 d), which is very close to that in the Na+

199

system (Figure 4a). Providing that the XRD peaks of T-bir are relatively weaker in the Li+ system

200

than in the Na+ system, we may conclude that a small portion of Mn(III) adsorbs on vacancies

201

without being incorporated into the layers, i.e., formation of Mn(III)-rich -MnO2.16

202

The formation of the TMO phase occurs earlier in both Li+ and K+ systems than in the Na+

203

system, which is clearly indicated in the earlier increase of the PDF peak area at 5.3 Å (Figure 3a)

204

and earlier appearance of needle-like crystals (Figure S2). The PDF peak area at 5.3 Å increases

205

because TMO contains abundant Mn-Mn corner-sharing atomic pairs with an interatomic distance

206

of 5.3 Å.16 The faster transformation to TMO in the Li+ and K+ systems than that in the Na+ system

207

can be ascribed to the formation of the Mn(III)-rich -MnO2 with Mn(III) adsorbed on vacancies,

208

which acts as a viable precursor for the transformation.16 The transformation of well-crystallized

209

T-bir to TMO in the Na+ system is slower. The XRD peak intensities of T-bir decrease significantly

210

in both Na+ and K+ systems at 25 d, indicating more extended transformation of T-bir to TMO. In

211

contrast, T-bir persists in the Li+ system, probably due to its very high crystallinity. The greater

212

increase of the PDF peak area at 5.3 Å (Figure 3a) and more needle-like particles at 25 d (Figure

213

2) in the K+ system suggest that more TMO forms in the K+ system than in the Na+ and Li+ systems.

214

However, the strong XRD peak at 12 Å suggests that the TMO formed in the Na+ system has the

215

highest crystallinity (Figure 1a), which could be due to the slower transformation with well-

ACS Paragon Plus Environment

10

Page 11 of 35

216

Environmental Science & Technology

crystallized T-bir precursor, favoring formation of the well-crystallized TMO.

217

Raman spectroscopy was used to further identify the solid phases. In the Li+ system, Raman

218

spectrum of the 10-h solid has two dominant bands at 574 and 651 cm–1 (Figure 5), indicating a

219

layered structure.54 The spectrum of the 25-d solid has an absorption band around 730 cm–1,

220

probably corresponding to tunneled structures.54,55 Similarly, the solids at 75 h or 10 h in both Na+

221

and K+ systems show more spectral features of δ-MnO2 or T-bir, while the 25-d samples have

222

more spectral features of tunneled structures (Figure 5). These results agree with the XRD analyses

223

(Figure 1a).

224

Concentrations of monovalent cations also affect the Mn(II)-induced birnessite

225

transformations. Higher NaCl concentration favors the formation of T-bir. Less T-bir forms in 10

226

mM NaCl than in 100 mM NaCl solution (Figure 1a,b); and well-crystallized T-bir is not detected

227

in 0 mM NaCl solution (Figure 1b). However, poorly crystalline T-bir or Mn(III)-rich -MnO2

228

might be abundant in the solutions of low NaCl concentrations, indicated by the sharp decrease of

229

PDF peak area at 5.3 Å within 5 h (Figure 3b). The favorability of high NaCl concentration for

230

T-bir formation could be caused by a sufficient number of Na+ that is needed to compensate the

231

negative layer charges of T-bir and by a higher ionic strength that favors particle aggregation and

232

thus particle growth via oriented attachment.32,56 4 × 4 TMO forms at all NaCl concentrations,

233

starting earlier at the lower NaCl concentration (Figure S2), while at 25 d, the transformation to

234

TMO is more extended and the TMO is more crystalline at the higher NaCl concentration. The

235

earlier formation of TMO is consistent with the abundance of poorly crystalline T-bir and/or

236

Mn(III)-rich -MnO2 as the precursors.16 Once primary TMO particles form, the growth of TMO

237

crystals via oriented attachment9 can be favored at higher ionic strength,56 leading to more

238

extended transformation to the TMO at 25 d in the 100 mM NaCl solution. The transformations

ACS Paragon Plus Environment

11

Environmental Science & Technology

239

Page 12 of 35

are consistent with the changes of the PDF peak area at 5.3 Å (Figures 3 and S3c,d,f).

240

The observed differences among Li+, Na+ and K+ systems may be caused by their different

241

magnitudes of electrostatic attraction with MnO6 layers. The ionic potential, φH = Z/rH, where rH

242

is the radius of hydrated cations,57 measures the charge density of a cation,58 and the electrostatic

243

forces between these hydrated cations and birnessite layers increase in the order of Li+ < Na+ < K+

244

(Table S3). However, except for the crystallinity of T-bir, other aspects of the transformation, such

245

as the amounts of T-bir and TMO and their formation rates, do not correlate with the order of the

246

ionic potentials, suggesting existence of other controlling factors, which could be hydration energy

247

and ionic radius of cations, in addition to electrostatic interactions.

248

Effects of Mg2+ and Ca2+. Different from alkali metal cations, both Mg2+ and Ca2+ lead to

249

formation of T-bir but not TMO. T-bir forms earlier, faster and more extensively and is much more

250

crystalline in the Mg2+ system than in the Ca2+ system (Figure 1c). Consistently, the peak area at

251

5.3 Å in the PDFs decreases less in the Ca2+ system (Figure 3a). The continuous decrease of PDF

252

peak area at 5.3 Å in both systems (Figures 3a and S4b,c) is consistent with the XRD and TEM

253

results showing that T-bir forms and increases in quantity and no tunneled phase forms. The Raman

254

spectra of the solids in the Mg2+ system show a gradual transformation from δ-MnO2 to T-bir with

255

incubation time (Figure 5). The 25-d solid in the Ca2+ system has weak Raman spectral features of

256

T-bir. None of the solids in Mg2+ nor Ca2+ systems possess spectral features of tunneled structures

257

(Figure 5).

258

The disfavored formation of T-bir in the Ca2+ system might result from the exchange of

259

adsorbed Mn(II) by Ca2+ so that less Mn(III) formed via the Mn(II)-Mn(IV) comproportionation

260

reaction.59 The formation of minor hausmannite (4.9 Å XRD peak in Figure 1c) suggests that the

261

released Mn(II) may react with the edge sites of -MnO2 because the edge sites are negatively

ACS Paragon Plus Environment

12

Page 13 of 35

Environmental Science & Technology

262

charged at pH 8 (pKa of Mn-OH groups on the edges is lower than 6).35,60 Due to weaker

263

adsorption, Mg2+ may not exchange adsorbed Mn(II) as strongly as Ca2+ does; otherwise, the

264

Mn(III) percentage in the Mg2+ system would not reach 34% (Figure 4a), the percentage value of

265

an ideal T-bir.61 In addition, Ca2+ can induce a drastic Mn(III) rearrangement in MnO6 layers and

266

interstratification of MnO6 layers with different Mn(III) distribution in T-bir,62 which may further

267

decrease the crystallinity of T-bir.

268

T-bir formed in the Mg2+ system has a high order of layer stacking, indicated by strong

269

(00l) XRD peaks (Figure 1c), which can be caused by the strong hydrogen bonds between MnO6

270

layers and hydrated Mg2+.46 The high hydration energy of Mg2+ prevents dehydration of Mg2+ upon

271

air drying, leading to the (003) diffraction peak at 3.2 Å.63,64 Compared to the high stacking order,

272

the degree of the intra-layer ordering of the a-b plane of T-bir in the Mg2+ system is surprisingly

273

low, as indicated by the broad XRD peaks between 2.4 – 1.4 Å (Figure 1c). The strong interaction

274

between Mg2+ and the layers may cause the poor Mn(III) arrangement in the layers.46

275

Effects of Cu(II). Mn(II)/Cu(II)-bearing δ-MnO2 does not undergo much phase

276

transformation. The XRD pattern of the 25-d solid is very similar to that of -MnO2 (Figure 1d),

277

and the Raman spectrum of the solid is dissimilar to both T-bir and TMOs (Figure 5), indicating

278

the absence of both T-bir and TMO phases. The 3Mn system uses a similar Mn(II)/MnO2 ratio as

279

the Cu system, but a small amount of poorly crystalline T-bir forms and decent layer stacking is

280

developed, as indicated by the (001) and (002) XRD peaks (Figure 1d) and the strong decrease of

281

the PDF peak area at 5.3 Å (Figures 3b and S3b). In spite of no mineral phase changes, the

282

weakened dip at 2.0 Å in the XRD patterns (Figure 1d) and the reduced PDF peak area at 5.3 Å

283

within 10 h (Figures 3b and S4d) suggest that a portion of Mn(II) is oxidized and incorporated into

284

vacancies.2,52,59 Thus, the Cu(II) system leads to formation of Mn(III)-rich -MnO2 which,

ACS Paragon Plus Environment

13

Environmental Science & Technology

Page 14 of 35

285

however, does not transform to T-bir and 4 × 4 TMO, suggesting an inhibition of Cu(II) on the

286

transformation to T-bir and 4 × 4 TMO.

287

The speciation of Cu(II) in the solids provides insights into the inhibition of Cu(II) on T-

288

bir formation. Both Cu K-edge XANES and EXAFS spectra show significant differences between

289

the incubated samples and spertiniite (Cu(OH)2) and tenorite (CuO) references (Figure S7a,b),

290

indicating that Cu(II) does not precipitate although the pH is high (pH 8). The EXAFS spectra of

291

the incubated solids differ only slightly from the initial sample (Figure S7b), suggesting that the

292

Cu coordination environment remains largely unchanged during the incubation. EXAFS spectral

293

fits show that each Cu atom is surrounded by about 4 O atoms (CN = 3.7) at 1.95 – 1.96 Å (Figure

294

S7b,c and Table S4), corresponding to four equatorial Cu-O bonds of the Jahn-Teller distorted

295

CuO6 octahedron.65,66 The two axial O atoms were not included in the fitting because of the strong

296

thermal motion of the long axial bonds at room temperature.67 The second shell consists of 1.5 Cu

297

at 2.90 – 2.93 Å, likely due to the formation of polynuclear clusters;

298

exclude the possibility that a small number of Cu incorporates into layer vacancies. The small CNs

299

of the second edge-sharing Cu-Cu/Mn shell at 2.90 – 2.93 Å and the third corner-sharing Cu-

300

Cu/Mn shell at 3.38 – 3.40 Å suggest that Cu mainly adsorbs at edge sites as double edge-sharing

301

complexes and single/double corner-sharing complexes.37 The absence of T-bir can be ascribed to

302

the adsorption of Cu(II) clusters on the edge sites, leading to Mn(III) disordered arrangement in

303

the layers. The Cu(II) in layers can impair intra-layer electron transfer and thus Mn(III)

304

rearrangement as well.

37,47

however, we cannot

305

Role of Metal Adsorption on Stability of Layered Mn Oxides. The degree of the

306

transformation of Mn(II)-bearing -MnO2 decreases in the order of Li+, Na+ or K+ > Mg2+ or Ca2+

307

> Cu(II). The cation impacts on the transformation can be understood from thermodynamic

ACS Paragon Plus Environment

14

Page 15 of 35

Environmental Science & Technology

308

perspectives by considering the contribution of surface energy to the stability of LMOs and how

309

much surface energy metal adsorption decreases.

310

LMOs have large total surface areas mainly from basal planes for well-crystalized LMOs

311

while from both edges and basal planes for small LMO particles. With decreasing basal plane size,

312

edge sites contribute increasingly more additional surface areas that can be substantial, such as for

313

-MnO2. Because of the high surface area to volume ratios of LMOs, surface energy contributes a

314

large portion of the total free energy (including both bulk and surface), elevating the total free

315

energy and probably driving their transformation to TMOs.68,69 Due to the large contribution from

316

edge sites, small LMO particles have high total free energy and readily transform to more

317

crystalline birnessite phase or to TMOs. For example, -MnO2 grows bigger in particle sizes via

318

oriented attachment70 or transforms to cryptomelane (2  2) at room temperature,71,72 while acid

319

birnessite with a larger particle size does not. However, ion adsorption on LMO surfaces (including

320

both edges and basal planes) can decrease the surface energy and thus the total free energy,

321

increasing their thermodynamic stability. The extent to which the surface energy can be decreased

322

depends on metal binding strength.

323

In the present work, Cu(II) adsorbs very strongly on LMO surfaces and the surface and

324

total free energy of LMOs can be decreased substantially, thereby stabilizing Mn(III)-rich -MnO2

325

without forming T-bir and 4 × 4 TMO (Figure 6). Similarly, the adsorption of transition metals

326

retards the transformation of LMOs to todorokite, a 3×3 TMO, under refluxing conditions.9,73-75

327

Mn(III)O6 is Jahn-Teller distorted, and its presence and ordered distribution in layers, such as in

328

T-bir, can increase the surface and total free energy of LMOs. Cu(II) adsorption decreases the

329

Gibbs free energy of Mn(III)-rich -MnO2,76 and the energy of Mn(III)-rich -MnO2 with Cu(II)

330

adsorbed is likely lower than that of T-bir, disfavoring the transformation of Mn(III)-rich -MnO2

ACS Paragon Plus Environment

15

Environmental Science & Technology

Page 16 of 35

331

to T-bir. Thus, Mn(III)-rich -MnO2 still maintains its hexagonal layer symmetry (Figure 6). The

332

stabilization of hexagonal birnessite by transition metals is also evidenced in the previous studies

333

although unlike Cu(II), vacancies are the main adsorption sites of those metals.5,77 For example,

334

during the biogenic formation of nanoparticulate birnessite, Na+ and Ca2+ favor the formation of

335

T-bir while Ni(II) favors the formation of hexagonal birnessite.33 Zn(II) inhibits Mn(III)

336

accumulation in the structure of -MnO2 during its formation78 or during its partial reduction by

337

dissolved organic matter,39 and Zn(II) adsorption on -MnO2 causes migration of Mn(III) out of

338

layers.79 Note that metal adsorption may cause kinetic hindrance of the transformation of

339

hexagonal birnessite, which could also contribute to the increased stability of the LMOs.

340

As to Mg2+ and Ca2+, their binding to LMOs is much weaker than those of transition metal

341

cations. Thus, Mn(III)-bearing -MnO2 proceeds to form T-bir. The high crystallinity and

342

abundance of T-bir in the Mg2+ system are consistent with its weaker binding strength. However,

343

the adsorption of Ca2+ and Mg2+ in the interlayers stabilizes T-bir, retarding or preventing its

344

transformation to TMOs at room temperature (Figure 6). High temperature would be required for

345

transformation of Mg2+-exchanged T-bir to TMOs.80 Monovalent cations Li+, Na+ and K+ weakly

346

bind to LMOs and their adsorption cannot much decrease the surface and total free energy of

347

Mn(III)-bearing -MnO2. Thus, in their solutions, Mn(III)-rich -MnO2 and T-bir readily form but

348

are unstable and eventually transform to the TMO (Figure 6). However, other environmental

349

factors may also contribute to the high abundance of LMOs in the natural environment, such as

350

the adsorption of oxyanions and the rapid formation of biogenic birnessite.

351

Environmental Implications. Layered Mn oxide minerals are common in the natural

352

environment and play important role in elemental cycles and pollutant dynamics. They have large

353

surface areas on basal planes and edges and are not thermodynamically stable and can transform

ACS Paragon Plus Environment

16

Page 17 of 35

Environmental Science & Technology

354

to TMOs. Our findings suggest that the unusually long lifetimes of LMOs in the natural

355

environment may be caused by the adsorption of divalent cations, probably also trivalent (e.g.,

356

Fe(III)), particularly transition metals, that increases their thermodynamic stability by decreasing

357

their surface energies. Thus, the stabilization of LMOs by metal adsorption may be an important

358

cause for the overall higher abundance of LMOs than TMOs in the natural environments. The

359

increased stabilization by metal adsorption may also decrease their redox potentials because

360

surface energy shifts redox equilibria.81 Our results have important implications for understanding

361

environmental and geochemical impacts of Mn oxides because LMOs are more reactive than

362

TMOs in terms of oxidation and metal adsorption reactivity.82 Effects of other environmental

363

factors, such as adsorption of oxyanions, on the stability of LMOs need to be addressed in future

364

studies.

365

ASSOCIATED CONTENT

366

Supporting Information.

367

The Supporting Information are available free of charge at http://pubs.acs.org, including

368

the details of preparation of δ-MnO2, TEM characterization, and XRD and Raman spectroscopic

369

analyses; XRD patterns, TEM images, PDF data, XANES and EXAFS spectra and corresponding

370

fitting results; dissolved Mn and Cu concentrations; and physicochemical atomic parameters of

371

alkali and alkaline earth metal cations.

372

Corresponding Author

373 374 375

*M. Zhu. E-mail: [email protected]; Phone: 307-766-5523 Notes The authors declare no competing financial interest.

ACS Paragon Plus Environment

17

Environmental Science & Technology

376

Page 18 of 35

ACKNOWLEDGEMENTS

377

This study was supported by the U.S. Department of Energy Experimental Program to

378

Stimulate Competitive Research Office for financial support (DOE-EPSCoR DE-SC0016272). We

379

thank Dr. Xiaoming Wang in the College of Resources and Environment, Huazhong Agricultural

380

University, Wuhan, China for providing Cu K-edge X-ray absorption spectra of spertiniite

381

(Cu(OH)2) and tenorite (CuO). This work utilized resources of APS, a U.S. DOE Office of Science

382

User Facility, operated for the DOE Office of Science by Argonne National Laboratory under

383

Contract No. DE-AC02-06CH11357. Use of SSRL, SLAC National Accelerator Laboratory, was

384

supported by the U.S. DOE, Office of Science, Office of Basic Energy Sciences, under Contract

385

No. DE-AC02-76SF00515.

386

REFERENCES

387 388

(1) Post, J. E. Manganese oxide minerals: Crystal structures and economic and environmental significance. Proc. Natl. Acad. Sci. U. S. A. 1999, 96 (7), 3447-3454.

389

(2) Lafferty, B. J.; Ginder-Vogel, M.; Zhu, M.; Livi, K. J. T.; Sparks, D. L. Arsenite oxidation by

390

a poorly crystalline manganese-oxide. 2. Results from X-ray absorption spectroscopy and X-ray

391

diffraction. Environ. Sci. Technol. 2010, 44 (22), 8467-8472.

392 393 394 395 396 397

(3) Peña, J.; Kwon, K. D.; Refson, K.; Bargar, J. R.; Sposito, G. Mechanisms of nickel sorption by a bacteriogenic birnessite. Geochim. Cosmochim. Acta 2010, 74 (11), 3076-3089. (4) Villalobos, M.; Bargar, J.; Sposito, G. Mechanisms of Pb(II) sorption on a biogenic manganese oxide. Environ. Sci. Technol. 2005, 39 (2), 569-576. (5) Zhu, M.; Ginder-Vogel, M.; Sparks, D. L. Ni(II) sorption on biogenic Mn-oxides with varying Mn octahedral layer structure. Environ. Sci. Technol. 2010, 44 (12), 4472-4478.

ACS Paragon Plus Environment

18

Page 19 of 35

398 399 400 401 402 403

Environmental Science & Technology

(6) Wang, Q.; Yang, P.; Zhu, M. Structural transformation of birnessite by fulvic acid under anoxic conditions. Environ. Sci. Technol. 2018, 52 (4), 1844-1853. (7) Lin, H.; Taillefert, M. Key geochemical factors regulating Mn(IV)-catalyzed anaerobic nitrification in coastal marine sediments. Geochim. Cosmochim. Acta 2014, 133, 17-33. (8) Yao, W.; Millero, F. J. Adsorption of phosphate on manganese dioxide in seawater. Environ. Sci. Technol. 1996, 30 (2), 536-541.

404

(9) Atkins, A. L.; Shaw, S.; Peacock, C. L. Nucleation and growth of todorokite from birnessite:

405

Implications for trace-metal cycling in marine sediments. Geochim. Cosmochim. Acta 2014, 144,

406

109-125.

407 408 409 410 411 412

(10) Manceau, A.; Lanson, M.; Takahashi, Y. Mineralogy and crystal chemistry of Mn, Fe, Co, Ni, and Cu in a deep-sea Pacific polymetallic nodule. Am. Mineral. 2014, 99 (10), 2068-2083. (11) Marcus, M. A.; Manceau, A.; Kersten, M. Mn, Fe, Zn and As speciation in a fast-growing ferromanganese marine nodule. Geochim. Cosmochim. Acta 2004, 68 (14), 3125-3136. (12) Sherman, D. M.; Peacock, C. L. Surface complexation of Cu on birnessite (δ-MnO2): Controls on Cu in the deep ocean. Geochim. Cosmochim. Acta 2010, 74 (23), 6721-6730.

413

(13) Flynn, E. D.; Catalano, J. G. Reductive transformations of layered manganese oxides by

414

small organic acids and the fate of trace metals. Geochim. Cosmochim. Acta 2019, 250, 149-172.

415

(14) Drits, V. A.; Silvester, E.; Gorshkov, A. I.; Manceau, A. Structure of synthetic monoclinic

416

Na-rich birnessite and hexagonal birnessite: I. Results from X-ray diffraction and selected-area

417

electron diffraction. Am. Mineral. 1997, 82 (9-10), 946-961.

418

(15) Silvester, E.; Manceau, A.; Drits, V. A. Structure of synthetic monoclinic Na-rich birnessite

419

and hexagonal birnessite: II. Results from chemical studies and EXAFS spectroscopy. Am. Mineral.

420

1997, 82 (9-10), 962-978.

ACS Paragon Plus Environment

19

Environmental Science & Technology

Page 20 of 35

421

(16) Yang, P.; Lee, S.; Post, J. E.; Xu, H.; Wang, Q.; Xu, W.; Zhu, M. Trivalent manganese on

422

vacancies triggers rapid transformation of layered to tunneled manganese oxides (TMOs):

423

Implications for occurrence of TMOs in low-temperature environment. Geochim. Cosmochim.

424

Acta 2018, 240, 173-190.

425

(17) Cui, H.; Liu, X.; Tan, W.; Feng, X.; Liu, F.; Ruan, H. D. Influence of Mn(III) availability

426

on the phase transformation from layered buserite to tunnel-structured todorokite. Clays Clay

427

Miner. 2008, 56 (4), 397-403.

428

(18) Cui, H.; Qiu, G.; Feng, X.; Tan, W.; Liu, F. Birnessites with different average manganese

429

oxidation states synthesized, characterized, and transformed to todorokite at atmospheric pressure.

430

Clays Clay Miner. 2009, 57 (6), 715-724.

431 432 433 434

(19) 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. (20) Nico, P. S.; Zasoski, R. J. Importance of Mn(III) availability on the rate of Cr(III) oxidation on δ-MnO2. Environ. Sci. Technol. 2000, 34 (16), 3363-3367.

435

(21) Landrot, G.; Ginder-Vogel, M.; Livi, K.; Fitts, J. P.; Sparks, D. L. Chromium(III) oxidation

436

by three poorly-crystalline manganese(IV) oxides. 1. Chromium(III)-oxidizing capacity. Environ.

437

Sci. Technol. 2012, 46 (21), 11594-11600.

438 439 440 441 442 443

(22) Kwon, K. D.; Refson, K.; Sposito, G. On the role of Mn(IV) vacancies in the photoreductive dissolution of hexagonal birnessite. Geochim. Cosmochim. Acta 2009, 73 (14), 4142-4150. (23) Zhao, W.; Cui, H.; Liu, F.; Tan, W.; Feng, X. Relationship between Pb2+ adsorption and average Mn oxidation state in synthetic birnessites. Clays Clay Miner. 2009, 57 (5), 513-520. (24) Zhao, W.; Wang, Q. Q.; Liu, F.; Qiu, G. H.; Tan, W. F.; Feng, X. H. Pb2+ adsorption on birnessite affected by Zn2+ and Mn2+ pretreatments. J. Soils Sediments 2010, 10 (5), 870-878.

ACS Paragon Plus Environment

20

Page 21 of 35

Environmental Science & Technology

444

(25) Zhang, T.; Liu, L.; Tan, W.; Suib, S. L.; Qiu, G.; Liu, F. Photochemical formation and

445

transformation of birnessite: Effects of cations on micromorphology and crystal structure. Environ.

446

Sci. Technol. 2018, 52 (12), 6864-6871.

447 448 449 450

(26) Tebo, B. M.; Johnson, H. A.; McCarthy, J. K.; Templeton, A. S. Geomicrobiology of manganese(II) oxidation. Trends Microbiol. 2005, 13 (9), 421-428. (27) Graybeal, A. L.; Heath, G. R. Remobilization of transition metals in surficial pelagic sediments from the eastern Pacific. Geochim. Cosmochim. Acta 1984, 48 (5), 965-975.

451

(28) Och, L. M.; Müller, B.; Voegelin, A.; Ulrich, A.; Göttlicher, J.; Steiniger, R.; Mangold, S.;

452

Vologina, E. G.; Sturm, M. New insights into the formation and burial of Fe/Mn accumulations in

453

Lake Baikal sediments. Chem. Geol. 2012, 330, 244-259.

454 455

(29) Elzinga, E. J. Reductive transformation of birnessite by aqueous Mn(II). Environ. Sci. Technol. 2011, 45 (15), 6366-6372.

456

(30) Elzinga, E. J.; Kustka, A. B. A Mn-54 radiotracer study of Mn isotope solid–liquid exchange

457

during reductive transformation of vernadite (δ-MnO2) by aqueous Mn(II). Environ. Sci. Technol.

458

2015, 49 (7), 4310-4316.

459 460

(31) 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), 10364-10371.

461

(32) Zhao, H.; Zhu, M.; Li, W.; Elzinga, E. J.; Villalobos, M.; Liu, F.; Zhang, J.; Feng, X.; Sparks,

462

D. L. Redox reactions between Mn(II) and hexagonal birnessite change its layer symmetry.

463

Environ. Sci. Technol. 2016, 50 (4), 1750-1758.

464 465

(33) 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 (12), 4465-4471.

ACS Paragon Plus Environment

21

Environmental Science & Technology

466 467 468 469 470 471 472 473 474 475

Page 22 of 35

(34) 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. (35) Peacock, C. L.; Sherman, D. M. Sorption of Ni by birnessite: Equilibrium controls on Ni in seawater. Chem. Geol. 2007, 238 (1), 94-106. (36) 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. (37) Peña, J.; Bargar, J. R.; Sposito, G. Copper sorption by the edge surfaces of synthetic birnessite nanoparticles. Chem. Geol. 2015, 396, 196-207. (38) Lefkowitz, J. P.; Elzinga, E. J. Structural alteration of hexagonal birnessite by aqueous Mn(II): Impacts on Ni(II) sorption. Chem. Geol. 2017, 466, 524-532.

476

(39) Wang, Q.; Yang, P.; Zhu, M. Effects of metal cations on coupled birnessite structural

477

transformation and natural organic matter adsorption and oxidation. Geochim. Cosmochim. Acta

478

2019, 250, 292-310.

479 480 481 482

(40) 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. (41) Landing, W. M.; Bruland, K. W. Manganese in the North Pacific. Earth. Planet. Sci. Lett. 1980, 49 (1), 45-56.

483

(42) Thomas, N. E.; Kan, K. T.; Bray, D. I.; MacQuarrie, K. T. B. Temporal changes in

484

manganese concentrations in water from the Fredericton aquifer, New Brunswick. Groundwater

485

1994, 32 (4), 650-656.

486 487

(43) Post, J. E.; Heaney, P. J.; Hanson, J. Rietveld refinement of a triclinic structure for synthetic Na-birnessite using synchrotron powder diffraction data. Powder Diffr. 2002, 17 (3), 218-221.

ACS Paragon Plus Environment

22

Page 23 of 35

Environmental Science & Technology

488

(44) Lopano, C. L.; Heaney, P. J.; Post, J. E.; Hanson, J.; Komarneni, S. Time-resolved structural

489

analysis of K- and Ba-exchange reactions with synthetic Na-birnessite using synchrotron X-ray

490

diffraction. Am. Mineral. 2007, 92 (2-3), 380-387.

491 492

(45) Posselt, H. S.; Anderson, F. J.; Weber, W. J. Cation sorption on colloidal hydrous manganese dioxide. Environ. Sci. Technol. 1968, 2 (12), 1087-1093.

493

(46) Post, J. E.; Veblen, D. R. Crystal structure determinations of synthetic sodium, magnesium,

494

and potassium birnessite using TEM and the Rietveld method. Am. Mineral. 1990, 75 (5-6), 477-

495

489.

496

(47) Qin, Z.; Xiang, Q.; Liu, F.; Xiong, J.; Koopal, L. K.; Zheng, L.; Ginder-Vogel, M.; Wang,

497

M.; Feng, X.; Tan, W.; Yin, H. Local structure of Cu2+ in Cu-doped hexagonal turbostratic

498

birnessite and Cu2+ stability under acid treatment. Chem. Geol. 2017, 466, 512-523.

499 500 501 502

(48) Sparks, D. L. 5 - Sorption Phenomena on Soils. In Environmental Soil Chemistry (Second Edition), Sparks, D. L., Ed. Academic Press: Burlington, 2003; Vol. pp 133-186. (49) Manceau, A.; Lanson, B.; Lanson, M.; Grangeon, S. Crystal structure of Ni-sorbed synthetic vernadite: a powder X-ray diffraction study. Mineral. Mag. 2008, 72 (6), 1279-1291.

503

(50) Burns, R. G.; Burns, V. M. The mineralogy and crystal chemistry of deep-sea manganese

504

nodules, a polymetallic resource of the twenty-first century. Phil. Trans. R. Soc. A 1977, 286

505

(1336), 283-301.

506 507 508 509

(51) Chiswell, B.; O'Halloran, K. R. Comparison of three colorimetric methods for the determination of manganese in freshwaters. Talanta 1991, 38 (6), 641-647. (52) Drits, V. A.; Lanson, B.; Gaillot, A.-C. Birnessite polytype systematics and identification by powder X-ray diffraction. Am. Mineral. 2007, 92 (5-6), 771-788.

ACS Paragon Plus Environment

23

Environmental Science & Technology

Page 24 of 35

510

(53) Kuma, K.; Usui, A.; Paplawsky, W.; Gedulin, B.; Arrhenius, G. Crystal structures of

511

synthetic 7 Å and 10 Å manganates substituted by mono- and divalent cations. Mineral. Mag. 1994,

512

58 (3), 425-447.

513 514 515 516 517 518 519 520 521 522 523 524

(54) Julien, C. M.; Massot, M.; Poinsignon, C. Lattice vibrations of manganese oxides. Part I. Periodic structures. Spectrochim. Acta A 2004, 60 (3), 689-700. (55) Julien, C.; Massot, M.; Rangan, S.; Lemal, M.; Guyomard, D. Study of structural defects in γ-MnO2 by Raman spectroscopy. J. Raman Spectrosc. 2002, 33 (4), 223-228. (56) Burrows, N. D.; Hale, C. R. H.; Penn, R. L. Effect of ionic strength on the kinetics of crystal growth by oriented aggregation. Cryst. Growth Des. 2012, 12 (10), 4787-4797. (57) Nightingale, E. R. Phenomenological theory of ion solvation. Effective radii of hydrated ions. J. Phys. Chem. 1959, 63 (9), 1381-1387. (58) Cartledge, G. H. Studies on the periodic system. I. The ionic potential as a periodic function. J. Am. Chem. Soc. 1928, 50 (11), 2855-2863. (59) 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.

525

(60) Tonkin, J. W.; Balistrieri, L. S.; Murray, J. W. Modeling sorption of divalent metal cations

526

on hydrous manganese oxide using the diffuse double layer model. Appl. Geochem. 2004, 19 (1),

527

29-53.

528 529 530 531

(61) Lanson, B.; Drits, V. A.; Feng, Q.; Manceau, A. Structure of synthetic Na-birnessite: Evidence for a triclinic one-layer unit cell. Am. Mineral. 2002, 87 (11-12), 1662-1671. (62) Drits, V. A.; Lanson, B.; Gorshkov, A. I.; Manceau, A. Substructure and superstructure of four-layer Ca-exchanged birnessite. Am. Mineral. 1998, 83 (1-2), 97-118.

ACS Paragon Plus Environment

24

Page 25 of 35

Environmental Science & Technology

532

(63) Bodeï, S.; Manceau, A.; Geoffroy, N.; Baronnet, A.; Buatier, M. Formation of todorokite

533

from vernadite in Ni-rich hemipelagic sediments. Geochim. Cosmochim. Acta 2007, 71 (23), 5698-

534

5716.

535

(64) Atkins, A. L.; Shaw, S.; Peacock, C. L. Release of Ni from birnessite during transformation

536

of birnessite to todorokite: Implications for Ni cycling in marine sediments. Geochim. Cosmochim.

537

Acta 2016, 189, 158-183.

538 539

(65) Oswald, H. R.; Reller, A.; Schmalle, H. W.; Dubler, E. Structure of copper(II) hydroxide, Cu(OH)2. Acta Cryst. C 1990, 46 (12), 2279-2284.

540

(66) Cheah, S.-F.; Brown, G. E.; Parks, G. A. XAFS study of Cu model compounds and Cu2+

541

sorption products on amorphous SiO2, γ-Al2O3, and anatase. Am. Mineral. 2000, 85 (1), 118-132.

542

(67) Manceau, A.; Nagy, K. L. Comment on “Molecular controls on Cu and Zn isotopic

543

fractionation in Fe–Mn crusts” by Little, S. H.; Sherman, D. M.; Vance, D.; Hein, J. R. Earth.

544

Planet. Sci. Lett. 2015, 411, 310-312.

545 546 547 548

(68) Navrotsky, A.; Mazeina, L.; Majzlan, J. Size-driven structural and thermodynamic complexity in iron oxides. Science 2008, 319 (5870), 1635-1638. (69) Luo, W.; Hu, W.; Xiao, S. Size effect on the thermodynamic properties of silver nanoparticles. J. Phys. Chem. C 2008, 112 (7), 2359-2369.

549

(70) Marafatto, F. F.; Lanson, B.; Peña, J. Crystal growth and aggregation in suspensions of δ-

550

MnO2 nanoparticles: implications for surface reactivity. Environ. Sci.: Nano 2018, 5 (2), 497-508.

551

(71) Eitel, E. M.; Zhao, S.; Tang, Y.; Taillefert, M. Effect of manganese oxide aging and structure

552

transformation on the kinetics of thiol oxidation. Environ. Sci. Technol. 2018, 52 (22), 13202-

553

13211.

ACS Paragon Plus Environment

25

Environmental Science & Technology

Page 26 of 35

554

(72) Grangeon, S.; Fernandez-Martinez, A.; Warmont, F.; Gloter, A.; Marty, N.; Poulain, A.;

555

Lanson, B. Cryptomelane formation from nanocrystalline vernadite precursor: a high energy X-

556

ray scattering and transmission electron microscopy perspective on reaction mechanisms.

557

Geochem. Trans. 2015, 16 (1), 12.

558

(73) Cui, H.; Feng, X.; Tan, W.; He, J.; Hu, R.; Liu, F. Synthesis of todorokite-type manganese

559

oxide from Cu-buserite by controlling the pH at atmospheric pressure. Micropor. Mesopor. Mater.

560

2009, 117 (1), 41-47.

561

(74) Cui, H.-J.; Liu, F.; Tan, W.-F.; Feng, X.-H. Effect of cobalt-doped framework on formation

562

of todorokite from layered manganese oxides with Mg2+/Co2+ ions as template. Pedosphere 2011,

563

21 (6), 730-737.

564

(75) Wu, Z.; Peacock, C. L.; Lanson, B.; Yin, H.; Zheng, L.; Chen, Z.; Tan, W.; Qiu, G.; Liu, F.;

565

Feng, X. Transformation of Co-containing birnessite to todorokite: Effect of Co on the

566

transformation and implications for Co mobility. Geochim. Cosmochim. Acta 2019, 246, 21-40.

567 568

(76) Sato, N. Electric double layer at electrode interfaces. In Electrochemistry at metal and semiconductor electrodes, Sato, N., Ed. Elsevier Science: Amsterdam, 1998; Vol. pp 119-199.

569

(77) Manceau, A.; Lanson, B.; Drits, V. A. Structure of heavy metal sorbed birnessite. Part III:

570

Results from powder and polarized extended X-ray absorption fine structure spectroscopy.

571

Geochim. Cosmochim. Acta 2002, 66 (15), 2639-2663.

572

(78) Zhao, S.; Wang, Q.; Sun, J.; Borkiewicz, O. J.; Huang, R.; Saad, E. M.; Fields, B.; Chen, S.;

573

Zhu, M.; Tang, Y. Effect of Zn coprecipitation on the structure of layered Mn oxides. Chem. Geol.

574

2018, 493, 234-245.

ACS Paragon Plus Environment

26

Page 27 of 35

Environmental Science & Technology

575

(79) Grangeon, S.; Manceau, A.; Guilhermet, J.; Gaillot, A.-C.; Lanson, M.; Lanson, B. Zn

576

sorption modifies dynamically the layer and interlayer structure of vernadite. Geochim.

577

Cosmochim. Acta 2012, 85, 302-313.

578 579

(80) Feng, X. H.; Tan, W. F.; Liu, F.; Wang, J. B.; Ruan, H. D. Synthesis of todorokite at atmospheric pressure. Chem. Mater. 2004, 16 (22), 4330-4336.

580

(81) Navrotsky, A.; Ma, C.; Lilova, K.; Birkner, N. Nanophase transition metal oxides show large

581

thermodynamically driven shifts in oxidation-reduction equilibria. Science 2010, 330 (6001), 199-

582

201.

583 584

(82) 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.

585

ACS Paragon Plus Environment

27

Environmental Science & Technology

586

Page 28 of 35

Figures (001)

2.4 1.4

(001)

(a)

2.4

(b) 1.4

4.7 (002)

4.7

10Na 25d

Li 25d

10Na 15d

Intensity

Li 15d Li 5d Li 50h Na 25d Na 15d

(002)

10Na 5d 10Na 50h 10Na 10h 0Na 25d

Na 5d Na 50h

0Na 15d 0Na 5d

K 25d K 15d K 5d 10Mn Init As-prep

2.0 4.6

12

0Na 50h

(001) (002)

6 (001)

(c)

2.4

As-prep

4.6

12

6

1.4

2.4

(d) 1.4

(002)

(003)

Mg 25d

3Mn 25d

Mg 15d

3Mn 15d

Intensity

Mg 5d 2.0

Mg 50h Ca 25d

4.6

Cu 25d

Ca 15d Ca 5d

2.0

10Mn Init

12

587 588 589 590 591 592 593 594 595 596 597

12

6

Cu 15d Cu Init As-prep

As-prep

4.6

3Mn 5d 3Mn Init As-prep

4.9

2.0

0Na 10h 10Mn Init

2.0

6

d-spacing (Å)

d-spacing (Å)

Figure 1. XRD patterns of as-prepared δ-MnO2 (As-prep) and incubated samples collected from the Li+, Na+, K+ (a), 0Na, 10Na (b), Mg2+, Ca2+ (c), Cu(II), and 3Mn (d) systems. 3Mn and 10Mn indicate the Mn(II) concentration of 3 and 10 mM during equilibration with δ-MnO2 at pH 4, respectively. Init stands for δ-MnO2 equilibrated with Mn(II) and Cu(II) at pH 4 without incubation at pH 8. 0Na and 10Na denote the systems with 0 and 10 mM NaCl as background electrolytes, respectively. In c) and d), the vertical solid lines indicate the appearance of T-bir. The vertical dash lines at 4.6, 4.7, and 4.9 Å indicate the appearance of feitknechtite (β-MnOOH), pyrochroite (Mn(OH)2) and hausmannite (Mn3O4), respectively. The XRD pattern of the As-prep sample is overlapped with initial samples in panels a and b and with incubated samples in the Ca2+, 3Mn, and Cu(II) systems in panels c and d to illustrate the changes of XRD patterns.

ACS Paragon Plus Environment

28

Page 29 of 35

Environmental Science & Technology

598 599

Figure 2. The TEM images of starting materials and 25-d samples from different reaction systems.

600

As-prep and 10Mn Init stand for as-prepared δ-MnO2 and that equilibrated with 10 mM Mn(II) at

601

pH 4 but without being incubated at pH 8, respectively.

ACS Paragon Plus Environment

29

Environmental Science & Technology

0.05

0.04

As-prep Li Na K Mg Ca

0.02

0.01

0.0

0.1

0.2

0.3

Page 30 of 35

(a) 0.05

0.04

As-prep 0Mn 3Mn 0Na 10Na Cu

0.02

0.01

0.4

0.0

0.03

0.02

0.02

0.01

0.01

0.2

0.3

0.4

Area

0.03

0.1

(b)

0.00

0.00 0

602

5

10

15

25

0

Time (d)

5

10

15

25

Time (d)

603

Figure 3. Changes of the PDF peak area at 5.3 Å with time in different reaction systems. The

604

insets show the data points within 10 h. As-prep stands for as-prepared δ-MnO2.

ACS Paragon Plus Environment

30

Page 31 of 35

Environmental Science & Technology

(a)

8

Mn(II) percentage (%)

30

Mn(III) percentage (%)

(b)

10

35

25

20

As-prep 0Mn 3Mn

15

Li Na Mg

6 4 2 0

10 0

605

1

2

3

25

0

Time (d)

3

6

9

12

24

Time (d)

606

Figure 4. XANES-LCF derived Mn(III) (a) and Mn(II) (b) molar percentages for selected reaction

607

systems. As-prep stands for as-prepared δ-MnO2.

ACS Paragon Plus Environment

31

Environmental Science & Technology

Page 32 of 35

730

Woodruffite Todorokite T-bir Cu 25d

Intensity (counts)

Ca 25d Mg 25d Mg 15d Mg 5d Mg 50h Mg 10h K 25d K 10h Na 25d Na 75h Li 25d Li 10h 574 651

10Mn Init As-prep

100 200 300 400 500 600 700 800 900 1000 –1

608

Wavenumber (cm )

609

Figure 5. Raman spectra of T-bir, woodruffite (3 × 4), todorokite (3 × 3), and selected samples

610

from the Li+, Na+, K+, Mg2+, Ca2+, and Cu(II) systems.

ACS Paragon Plus Environment

32

Page 33 of 35

Environmental Science & Technology

Triclinic birnessite

4 × 4 TMO

δ-MnO2 Mn(III)-rich δ-MnO2 +

Mn(II)

+,K

a

+,N

Li

Mn(II)-bearing δ-MnO2

Mg2+, Ca2+

Cu

Triclinic birnessite

Mn(II) Mn(III)

(II )

Mn(III)-rich δ-MnO2

Mn(IV) Cu(II)

611 612

Figure 6. Transformation pathways of Mn(II)-bearing δ-MnO2 to other phases in different

613

systems.

ACS Paragon Plus Environment

33

Environmental Science & Technology

Page 34 of 35

614

Table(s)

615

Table 1. A summary of the experimental conditions. All experiments were carried out using

616

δ-MnO2 (4 g/L) as starting material at pH 8 in a N2/H2 atmosphere.

Condition

Initial Mn(II) concentration (mM)

Mn(II)/MnO2 molar ratio in Mn(II)-bearing -MnO2

Background electrolyte

Ionic strength (mM)

0Mn

0

0

100 mM NaCl

100

3Mn

3

0.08

100 mM NaCl

100

0Na

10

0.14

0 mM NaCl

0

10Na

10

0.14

10 mM NaCl

10

Li+

10

0.14

100 mM LiCl

100

Na+

10

0.14

100 mM NaCl

100

K+

10

0.14

100 mM KCl

100

Mg2+

10

0.14

33.3 mM MgCl2

100

Ca2+

10

0.14

33.3 mM CaCl2

100

Cu(II)

5 mM Mn(II) +

0.08

100 mM NaCl

100

7 mM Cu(II)

617

ACS Paragon Plus Environment

34

Page 35 of 35

618

Environmental Science & Technology

TOC

619

ACS Paragon Plus Environment

35