Interactions in Ternary Mixtures of MnO2, Al2O3 ... - ACS Publications

Feb 4, 2016 - Department of Civil and Environmental Engineering, Temple University 1947 .... works,8,23 the oxidative reactivity of MnO2 in all mixtur...
0 downloads 0 Views 380KB Size
Subscriber access provided by ORTA DOGU TEKNIK UNIVERSITESI KUTUPHANESI

Article

Interactions in Ternary Mixtures of MnO2, Al2O3, and Natural Organic Matter (NOM) and the Impact on MnO2 Oxidative Reactivity Saru Taujale, Laura R. Baratta, Jianzhi Huang, and Huichun (Judy) Zhang Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.5b05314 • Publication Date (Web): 04 Feb 2016 Downloaded from http://pubs.acs.org on February 4, 2016

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

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

Page 1 of 21

1 2

Environmental Science & Technology

Interactions in Ternary Mixtures of MnO2, Al2O3, and Natural Organic Matter (NOM) and the Impact on MnO2 Oxidative Reactivity

3

Saru Taujale, Laura R. Baratta, Jianzhi Huang and Huichun Zhang*

4

Department of Civil and Environmental Engineering, Temple University

5

1947 North 12th Street, Philadelphia, PA 19122

6

*Corresponding Author, contact e-mail: [email protected], phone: (215)204-4807, fax: (215)204-4696

7 8

ABSTRACT

9

Our previous work reported that Al2O3 inhibited the oxidative reactivity of MnO2 through

10

heteroaggregation between oxide particles and surface complexation of the dissolved Al ions with MnO2

11

(S. Taujale and H. Zhang, “Impact of interactions between metal oxides to oxidative reactivity of

12

manganese dioxide” Environ. Sci. Technol. 2012, 46, 2764-2771). The aim of the current work was to

13

investigate interactions in ternary mixtures of MnO2, Al2O3, and NOM and how the interactions affect

14

MnO2 oxidative reactivity. For the effect of Al ions, we examined ternary mixtures of MnO2, Al ions, and

15

NOM. Our results indicated that an increase in the amount of humic acids (HAs) increasingly inhibited Al

16

adsorption by forming soluble Al-HA complexes. As a consequence, there was less inhibition on MnO2

17

reactivity than by the sum of two binary mixtures (MnO2+Al ions and MnO2+HA). Alginate or

18

pyromellitic acid (PA) – two model NOM compounds – did not affect Al adsorption, but Al ions

19

increased alginate/PA adsorption by MnO2. The latter effect led to more inhibition on MnO2 reactivity

20

than the sum of the two binary mixtures. In ternary mixtures of MnO2, Al2O3, and NOM, NOM inhibited

21

dissolution of Al2O3. Zeta potential measurements, sedimentation experiments, TEM images, and

22

modified DLVO calculations all indicated that HAs of up to 4 mg-C/L increased heteroaggregation

23

between Al2O3 and MnO2 while higher amounts of HAs completely inhibited heteroaggregation. The

24

effect of alginate is similar to that of HAs although not as significant, while PA had negligible effects on

25

heteroaggregation. Different from the effects of Al ions and NOMs on MnO2 reactivity, the MnO2

26

reactivity in ternary mixtures of Al2O3, MnO2, and NOM was mostly enhanced. This suggests MnO2

1 ACS Paragon Plus Environment

Environmental Science & Technology

27

reactivity was mainly affected through heteroaggregation in the ternary mixtures because of the limited

28

availability of Al ions.

Page 2 of 21

29 A) No Humic Acid, binary oxide Al

B) [Humic Acid] < 4 mg-C/L, ternary Al

C) [Humic Acid] ≥ 4 mg-C/L, ternary Al

Al2O3

30 31

MnO2 Surface

Humic Acid

Al Al ions

TOC Art

32 33 34

INTRODUCTION Manganese oxides have a vital role in the fate and transport of organic contaminants in the

35

environment, mostly via adsorption, hydrolysis, and redox reaction of the contaminants.[1-7] Although

36

numerous work has examined the reactivity of MnO2 as a single oxide, the obtained results cannot be

37

directly extrapolated to natural soil-water systems because they are more complex mixtures containing

38

various metal oxides and natural organic matter (NOM). A previous study of ours revealed that among the

39

metal oxides examined, Al2O3 had the most negative impact on the oxidative reactivity of MnO2, mostly

40

through both heteroaggregation between the oxide particles and complexation of the Al ions released

2 ACS Paragon Plus Environment

Page 3 of 21

Environmental Science & Technology

41

from Al2O3 with MnO2.[8] Both interactions block the reactive sites on MnO2 surface to make it less

42

reactive, with the complexation of Al ions with MnO2 the dominant inhibition mechanism.

43

NOM is ubiquitous in the environment and can interact extensively with various metal oxides.

44

For instance, sorption of NOM by metal oxides such as aluminum and iron oxides has been extensively

45

reported to occur through ligand exchange, complexation with the acidic (-COOH) and hydroxyl (-OH)

46

functional groups, hydrogen bonding, electrostatic interaction, and cation bridging among others.[9-13]

47

NOM can also significantly affect the redox reactivity of MnO2 [1, 4] and sorption of metal ions by metal

48

oxides. In ternary mixtures of metal ions, NOM, and metal oxides, multiple interactions exist: i)

49

formation of soluble complexes in solution, ii) formation of ternary surface complexes, iii) competition

50

for surface sites, and iv) changing electrostatic properties of the oxide surface.[14] Metal ions such as Al3+,

51

Fe3+ and Cu2+ are known to form strong soluble complexes with NOM or NOM model compounds,[15-18]

52

leading to decrease in their sorption by oxide surfaces.[19-20] For strongly complexing ions (Me) such as

53

Cu2+, both type A (surface-Me-HA) and type B (surface-HA-Me) complexes can form, but the relative

54

contribution of each complex depends greatly on the solution condition.[18] For weakly complexing ions

55

such as Ca2+, electrostatic interactions dominate the sorption so there is weak sorption of Ca2+ by the

56

positively charged goethite but strong sorption of Ca2+ by the negatively charged fulvic acid (FA).[14, 21]

57

Alginate, a widely used model compound for polysaccharides, was also found to complex with divalent

58

metal ions such as Ca2+, Sr2+, and Ba2+. [22]

59

Despite the above knowledge, we know little about how NOM affects oxide redox reactivity in

60

mixed oxide systems. Our recent work examined how NOM affects oxidative reactivity of MnO2 in the

61

presence of Fe(III) oxides.[23] The primary goal of this study was to examine the reactivity of MnO2 in

62

ternary mixtures with Al2O3 and NOM . To examine how MnO2 reactivity was affected by

63

heteroaggregation with Al2O3 in the presence of NOM, we conducted sedimentation experiments for the

64

extent of aggregation, measured zeta potentials for oxide surface charges, examined transmission electron

65

microscopic (TEM) images of oxide particles, conducted modified DLVO calculations of interaction

66

energy profiles, obtained adsorption isotherms of NOM, and collected ART-FTIR images of soluble vs

67

adsorbed NOM. To examine how MnO2 reactivity was affected by Al complexation in the presence of 3 ACS Paragon Plus Environment

Environmental Science & Technology

68

NOM, we measured dissolution of Al2O3 in binary and ternary mixtures and examined adsorption of

69

NOM and Al ions in ternary systems of MnO2, NOM, and Al ions. Similar to our previous works, [8, 23] the

70

oxidative reactivity of MnO2 in all mixtures was measured based on the oxidation kinetics of triclosan – a

71

widely used antibacterial agent.

72

The NOMs used in this study are Aldrich humic acid (AHA), Leonardite humic acid (LHA),

73

alginate, and pyromellitic acid (PA). Alginate, a natural polymer, is commonly studied as a model NOM

74

and has shown to stabilize fullerene nanoparticles. Its reported molecular weight is 12- 80 kDa and is

75

composed of blocks of 1,4-linked β-D-mannuronic acid and α-L-guluronic acid. Similar to HA, adsorbed

76

alginate changes the surface charges of metal oxides and affects the stability of the oxides.[9-10] PA

77

(C10H6O8) is used as a model NOM, as previous studies have reported its adsorption behavior to be

78

analogous to that of naturally occurring NOM.[24] Compared to the other NOM used in this study, PA is a

79

much smaller molecule (molecular weight: 254.15 g/mol).[24]

80 81 82

MATERIALS AND METHODS Details on the chemicals, oxide preparation, experimental setup and analysis for triclosan

83

oxidation kinetics, TEM images, sedimentation, zeta potentials, adsorption of NOM, and ATR-FTIR

84

images are shown in Text S1 in the supporting information (SI). Briefly, for triclosan oxidation kinetics,

85

the reactors contained 5 mg/L of MnO2, 10 µM of triclosan, 0.01 M NaCl, and 25 mM acetic acid to

86

maintain a pH of 5.0. Aliquots of reaction suspensions were added into centrifuge tubes containing

87

enough 1 M NaOH to raise the suspension pH to greater than 10. This was followed by centrifugation at

88

12,100 ×g for 20 minutes. Previous studies have shown that triclosan oxidation is very slow at alkaline

89

pH thus quenching the reaction, and also pH>10 desorbs >98% triclosan from the metal oxides.[1] This

90

method allowed us to measure loss in triclosan due to oxidation only. After centrifugation, the

91

supernatants were transferred to separate vials for triclosan analysis by an Agilent 1200 HPLC.

92

TEM images were collected on a JEOL (JEM 1400) transmission electron microscope. Details on

93

sample preparation are in Text S1. Sedimentation experiments of Al2O3 and Al2O3+MnO2 in the presence

94

of various NOMs were conducted at pH 5.0 by using a UV-vis spectrophotometer. The optical absorbance 4 ACS Paragon Plus Environment

Page 4 of 21

Page 5 of 21

Environmental Science & Technology

95

by Al2O3 or Al2O3+MnO2 were measured at 508 nm as a function of time.[25] Zeta potentials and pHzpc

96

were measured using a Zetasizer Nano ZS (Malvern Instruments). The observed effects of NOM on the

97

surface charge of Al2O3 are in Text S3. ATR-FTIR spectra of the soluble and adsorbed PA were collected

98

with a Perkin-Elmer Spectrum 100 FTIR spectrometer equipped with a deuterated triglycine sulfate

99

(DTGS) detector. Details on the instrument conditions, sample preparations along with the results are

100

described in Texts S1 and S4.

101

Adsorption of NOMs by either Al2O3 or Al2O3+MnO2 was carried out under experimental

102

conditions identical to kinetic experiments while varying the initial concentrations of NOM. The amounts

103

of NOM adsorbed by Al2O3 or Al2O3+MnO2 were calculated by subtracting the amounts of NOM

104

measured in the filtered samples from the initial NOM concentrations (Ci). Note that our control

105

experiments have showed negligible adsorption of the NOMs by the filters. AHA, LHA, and PA

106

concentrations were measured by UV-vis while alginate concentrations were measured by a TOC

107

analyzer.

108

Adsorption experiments of Al ions by MnO2 were carried out as a function of NOM

109

concentration. To prepare the reactors, 5 mg/L of MnO2 was added to 50 mL DI water with 0.01 M NaCl

110

and 25 mM of acetate buffer. Al ion was added to suspensions to maintain an initial concentration of

111

0.009 mM or 0.03 mM. After equilibrating the mixture for an hour by stirring on a magnetic stir plate,

112

NOM were added to the reactors. After allowing the mixtures to stir overnight, they were filtered using

113

0.22 µm filters. The filtrates were acidified using 1 M HCl and the Al concentrations in the filtrates were

114

measured using an ICP-MS.

115

Dissolution experiments of Al2O3 in single oxide systems, i.e., only Al2O3, and in binary oxide

116

mixtures of Al2O3+MnO2 were carried out as a function of NOM concentration. 0.1 g/L of Al2O3 and 5

117

mg/L of MnO2 were added to reactors containing 50 mL DI water, 0.01 M NaCl, and 25 mM acetic acid

118

buffer. The reactors were allowed to equilibrate for an hour before adding the NOM. The mixtures were

119

then equilibrated overnight on a magnetic stir plate and centrifuged followed by filtration through 0.22

120

µm filters. The filtrates were acidified using 1 M HCl and Al concentrations were measured using an ICP-

121

MS. 5 ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 21

122 123

RESULTS AND DISCUSSION

124

NOM adsorption by Al2O3 and Al2O3+MnO2

125

NOM adsorption by Al2O3 and Al2O3+MnO2 was determined while varying the concentration of

126

AHA or LHA from 0 to 52.4 or to 82.9 mg-C/L, respectively. The concentrations of both alginate and PA

127

were varied from 0 to 200 mg/L. For both HAs and the model NOMs, we did not see a significant

128

difference in the amount of NOM adsorbed by Al2O3 and by Al2O3+MnO2 (Figure 1). This was mostly

129

because of the negligible adsorption of the NOMs by MnO2.[23] Therefore, any adsorption of NOM by

130

Al2O3+MnO2 is mostly due to Al2O3 only. A difference observed among the adsorption trends of the

131

NOMs was that the HA systems had reached an equilibrium between 20 to 25 mg-C/L while the

132

adsorption trends for both alginate and PA were still linear even at the highest initial NOM concentration

133

of 200 mg/L. A comparison between alginate and PA adsorption reveals a higher alginate adsorption by

134

both systems. A lower adsorption of HAs compared to alginate/PA at high concentrations is most likely

135

associated with the significantly larger size of the HA molecules that prevents further adsorption of HAs

136

on oxide surfaces due to steric effects.

a)

25

60

15

b)

40

10

Al2O3 (AHA) Al2O3+MnO2 (AHA) Al2O3 (LHA) Al2O3+MnO2 (LHA)

5 0 0

137

Qe (mg/0.1g)

Qe(mg-C/0.1g)

20

Al2O3 (PA) Al2O3+MnO2(PA) Al2O3 (Alginate) Al2O3+MnO2(Alginate)

80

20

40 60 Ce(mg-C/L)

80

100

20 0 0

20 40 60 80 100 120 140 Ce (mg/L)

138 139 140 141

Figure 1: Adsorption isotherms of a) AHA and LHA and b) PA and alginate by Al2O3 and Al2O3+ MnO2. Qe is the amount adsorbed and Ce is the equilibrium solution concentration. Conditions: 0.1 g/L Al2O3, 5 mg/L MnO2, 0.01 M NaCl, 25 mM acetate buffer, pH 5.0.

142

Effects of NOM on the homo- and heteroaggregation of metal oxides

6 ACS Paragon Plus Environment

Page 7 of 21

143

Environmental Science & Technology

TEM images were used in analyzing the extent of homo and hetero-aggregation between the

144

oxide particles at pH 5.0 and 0.01 M NaCl. Images of single oxides and oxide mixtures were obtained in

145

the presence of different concentrations of NOM. Selected area diffraction (image not shown) indicated

146

that the MnO2 (Figure S3a) used in our study is amorphous, Figure S3b shows Al2O3 as distinct

147

nanoparticles, and Figure S3c suggests intensive heteroaggregation between MnO2 and Al2O3 under the

148

examined conditions. TEM analysis was also conducted for MnO2 and MnO2+Al2O3 in the presence of

149

0.4 mg-C/L of AHA. Addition of AHA seemed to have little effect on either the homoaggregation

150

between the MnO2 particles (Figure S3d) or the heteroaggregation between Al2O3 and MnO2 (Figure

151

S3e). It must be noted that the air drying process of the sample grids may affect the aggregation between

152

the oxide particles, so the TEM images obtained may not represent the oxide interactions in the aqueous

153

suspension. Therefore, we conducted sedimentation experiments as an alternative method to study the

154

effect of NOMs on the extent of both homo and heteroaggregation. It is also worthwhile to mention that

155

light scattering techniques were not used to study the aggregation due to the wide range of particle size

156

distribution in our systems.

157

The effect of NOM on the homo and hetero-aggregation of the oxide particles was studied by

158

monitoring the sedimentation rates of Al2O3 and Al2O3+MnO2 at varying NOM concentrations. To

159

measure the extent of homo or heteroaggregation between the oxide particles, rather than the rate of

160

aggregation, the systems were allowed to equilibrate for overnight before UV-vis analysis. Therefore, the

161

systems had reached a pseudo steady-state of aggregation after the pre-equilibrium. The rate of

162

sedimentation is correlated to the size of the aggregates. Faster sedimentation corresponds to a higher

163

extent of aggregation and slower sedimentation corresponds to a lesser extent of aggregation. As shown in

164

Figures 2a and 2c, AHA and LHA affected the sedimentation rates of Al2O3 to a certain extent.

165

Sedimentation of Al2O3 increased as the concentrations of AHA were increased. In the case of LHA, the

166

sedimentation rate was higher when the concentration of LHA was less than 4.6 mg-C/L but much lower

167

when LHA concentration reached 12.5 mg-C/L. The sedimentation of the Al2O3 particles occurs as a

168

result of homoaggregation. The higher sedimentation rate when the HAs were added is likely due to the

169

neutralization of surface charges of Al2O3 by the adsorbed HA molecules, leading to more aggregation. 7 ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 21

170

Upon further increase in the concentration of the HAs, there was enough HA to reverse the surface charge

171

(Figure S1b) such that the colloids were electrostatically stabilized. 1.2

Al2O3

a)

Al2O3+MnO2

b)

1.0

A/A0

0.8 0.6

AHA

AHA

10 4 0 0.2 0.4 2

0.2 0 0.4 2 4 10

0.4 0.2 0.0 1.2

Al2O3

1.0

Al2O3+MnO2

c)

d)

0.8

A/A0

0.6

LHA

0.4 0.2 0.0

172

LHA

12.5 0 0.5 2 4.6

0

20

12.5 4.6 0 0.5 2.0

40 60 80 Time (minutes)

100 120 0

20

40 60 80 Time (minutes)

100 120

173 174 175 176

Figure 2: Sedimentation of Al2O3 and Al2O3+MnO2 in the presence of varying concentrations of AHA (a and b) and LHA (c and d). Conditions: 0.1 g/L MnO2, 0.1 g/L Al2O3, 0-10 mg-C/L as AHA, 0-12.5 mgC/L as LHA, 0.01 M NaCl, and pH 5.0.

177

Al2O3+MnO2 had higher sedimentation rates at 0 to 2 mg-C/L AHA as compared to the

178

sedimentation rates at 4-10 mg-C/L AHA (Figure 2b). Similar results were seen in sedimentation of the

179

binary oxide at varying concentrations of LHA (Figure 2d). Given the opposite surface charges of Al2O3

180

and MnO2 at pH 5 (Figure S1a), in the absence of HA or at low HA concentrations, a higher extent of

181

heteroaggregation is expected to occur between the oxides. At HA concentration ≥ 10 mg-C/L, the binary

182

oxide suspensions did not sediment over time indicating there was negligible homo or heteroaggregation

183

between the particles. Although there was sedimentation in the Al2O3 system at ≥ 10 mg-C/L HA, the

184

Al2O3+MnO2 system with ≥ 10 mg-C/L of HAs was more stable. The difference in the stabilities of the

185

two systems is most likely due to the difference in Al2O3 concentration (0.2 vs 0.1 g/L in single Al2O3 vs.

186

binary oxide systems). Since there is insignificant AHA adsorption by MnO2, there is a higher amount of 8 ACS Paragon Plus Environment

Page 9 of 21

Environmental Science & Technology

187

AHA adsorbed by Al2O3 in the binary oxide system which imparts more negative charge on the oxide

188

surface and hence decreases the homo-aggregation within the Al2O3. As shown in Figure 1a, the majority

189

of AHA at concentrations less than 10 mg-C/L was adsorbed by 0.1 g/L of Al2O3. With 0.2 g/L of Al2O3,

190

majority of AHA should be adsorbed as well. Therefore, the adsorption density of AHA is less for 0.2 g/L

191

of Al2O3, which would lead to less impact of AHA on the surface charge of Al2O3 at 0.2 g/L than at 0.1

192

g/L.

193

Sedimentation results with alginate are similar to those with the HAs (Figures S4e-f). In both

194

single and binary oxide systems, sedimentation rates initially increased with increasing alginate

195

concentration but decreased again when alginate concentration was above 10 mg/L. These results show

196

that there was a decrease in the extent of heteroaggregation between MnO2 and Al2O3 when a larger

197

amount of alginate (≥ 10 mg/L) was added to the system. Figures S4g-h show that, although PA

198

adsorption by the oxides is only slightly lower than alginate adsorption, varying concentrations of PA did

199

not affect the sedimentation rates of Al2O3 and Al2O3+MnO2. One of the possible reasons for the

200

difference in sedimentation patterns is the steric hindrance effect in the case of alginate. This effect could

201

be lower for PA due to its much smaller molecular size. However, further studies are needed to confirm

202

this possibility.

203 204 205

Extended DLVO calculations The DLVO theory has been widely used to examine the interaction mechanisms between particles

206

in aqueous systems. The overall interaction energies between the particles are determined based on the

207

sum of two forces viz. electrostatic repulsive forces and van der Waals attractive forces. However, in the

208

presence of NOM, steric repulsion due to the large NOM molecules attached on the oxide surfaces also

209

contributes to the overall interaction between the particles. Therefore, in our systems the net energy

210

between the particles was determined based on the sum of the three forces. As shown in our previous

211

work, [23] the modified DLVO theory can better explain the aggregation pattern between metal oxides in

212

the presence of NOM.

9 ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 21

213

The interaction energy profiles shown in Figure S5a are based on the classical DLVO theory,

214

which does not consider the steric repulsive energy. They indicate net negative interaction energies of

215

zero at all concentrations of AHA. However, based on the information gathered from the sedimentation

216

experiments, there was a negligible amount of aggregation in the mixed oxide system at higher HA

217

concentrations. This shows the inadequacy of the classical DLVO theory in explaining the stable behavior

218

of the oxides at higher concentrations of AHA. For the calculation purpose, the reported estimated values

219

of 7 nm[26] or 5.7 nm[27] for the thickness of HA adsorbed on the surface of hematite and TiO2,

220

respectively, were used to represent the thickness of the adsorbed AHA layer (Text S5). The energy

221

profiles for the interaction between MnO2 and Al2O3 based on these values (Figures S5b-c) show the

222

overall interaction energies to be positive at 4 and 10 mg-C/L AHA. These results are in agreement with

223

the results seen in our sedimentation experiments, supporting the mixed oxide system to be stable at

224

higher concentrations of HA.

225 226

Dissolution of Al2O3 in the presence of NOM

227

To understand the importance of Al ions on MnO2 reactivity in ternary MnO2+Al2O3+NOM

228

mixtures, we measured the amount of Al dissolved from Al2O3 in the mixtures. The detected Al ion

229

concentrations in the single and binary oxide systems as a function of NOM are shown in Figure 3. Figure

230

3a shows that the amount of Al dissolved in the single Al2O3 suspension mostly decreased as the

231

concentration of NOM was increased. In the absence of NOM, a much lower concentration of Al ions was

232

measured in the binary oxide system (Figure 3b) than in the single oxide system. This is mainly a result of

233

intense hetero-aggregation between Al2O3 and MnO2 as well as the adsorption of Al ions by MnO2.[8]

234

After the addition of NOM to the binary oxide mixture, there was a slight decrease in Al ion

235

concentration with increase in most of the NOMs (Figure 3b). This is again due to the passivation of

236

Al2O3 surfaces upon NOM adsorption to inhibit its dissolution. Indeed, outer-sphere ligands such as PA

237

and HA are known to inhibit dissolution of Al oxides.[28-30] Addition of alginate of > 10 mg/L led to a

238

significantly increased dissolution of Al2O3 in both single and binary oxide systems, further research is

239

however needed to elucidate the Al ion release mechanism. 10 ACS Paragon Plus Environment

Page 11 of 21

Environmental Science & Technology

0.015

0

10 a)

Alginate/PA (mg/L) 20 30 40 Al2O3

60

AHA LHA Alginate PA

0.012

[Al3+] mM

50

0.008

0

10 b)

Alginate/PA (mg/L) 20 30 40 Al2O3+MnO2

0.006

50

60

AHA LHA Alginate PA

0.009 0.004 0.006 0.002

0.003 0.000

0

2

4 6 8 10 AHA/LHA (mg-C/L)

12

14

0.000

0

2

4 6 8 10 AHA/LHA (mg-C/L)

12

14

240 241 242 243 244

Figure 3: Soluble Al ions measured in a) Al2O3 and b) Al2O3+MnO2 in the presence of NOMs. Experimental conditions: 5.0 mg/L MnO2, 0.1 g/L Al2O3, 0-12.5 mg-C/L AHA and LHA, 0-50 mg/L alginate and PA, 0.01 M NaCl, pH 5.0.

245

Adsorption of Al ions by MnO2 with or without NOM

246

Because of the dissolution of Al2O3 to form soluble Al ions, we had to examine ternary mixtures

247

of MnO2, Al ion, and NOM to fully understand the role of Al ions in ternary mixtures of MnO2, Al2O3,

248

and NOM. This included examining how soluble Al is adsorbed by MnO2 in the absence and presence of

249

NOM, how Al ions affect NOM adsorption, and how Al ions affect oxidative reactivity of MnO2, as

250

shown in the following three sections.

251

Our previous work showed that Al ions can sorb strongly to the MnO2 surface which significantly

252

lowered the oxidative reactivity of MnO2.[8] The sorption of a number of divalent metal cations on a

253

hydrous Mn oxide has been modeled based on a 2-site diffuse double layer model.[31] We adopted the

254

same modeling approach to provide macroscopic information on the formed surface complexes. As

255

details shown in the SI Table S1 and Figure S6, there are mainly two types of surface complexes formed:

256 257

𝐴𝐴 3+ + ≡ 𝑋𝑋𝑋 ⇌ ≡ 𝑋𝑋𝐴𝑙 2+ + 𝐻 +

𝐴𝐴 3+ + ≡ 𝑌𝑌𝑌 ⇌ ≡ 𝑌𝑌𝐴𝐴 2+ + 𝐻 +

log K1 = 2.58

(1)

log K2 = 1.44

(2)

258

where ≡XOH and ≡YOH are the strong and weak surface sites, and ≡XOAl2+ and ≡YOAl2+ are the two

259

types of inner-sphere surface complexes formed. Based on the obtained complexation constants, Al ions 11 ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 21

260

can be strongly sorbed by MnO2.[31] Al ions have also been reported to be able to form strong complexes

261

with NOM and NOM model compounds.[16, 32]

262

At both 0.009 and 0.03 mM Al ions, the adsorption of Al ions by MnO2 decreased with the

263

increase in AHA or LHA concentration (Figure 4). This is likely due to (1) competition of the HAs and Al

264

ions for the limited number of surface sites and (2) formation of soluble Al-HA complexes that resist

265

adsorption by MnO2. As reported, the enhanced oxidative reactivity of MnO2 in the presence of metal

266

ions and 0.1 – 1 mg/L of HA was attributed to the strong binding ability of HA for the ions and thus less

267

“occupied” surfaces of MnO2 by the ions.[4] In addition, in the presence of FA, sorption of Cu2+ by

268

hematite decreased at pH >6 due to an increasing concentration of soluble Cu-FA complexes.[19] The

269

decrease in the amount of Al adsorbed with an increasing amount of LHA and AHA is thus at least partly

270

due to soluble Al-HA complexes formed. Because the effect of LHA on the amount of Al sorbed is more

271

significant than AHA, LHA might have formed stronger complexes with Al ions than AHA.

272

Adsorption of Al ions by MnO2 was only slightly lowered by the presence of a large amount of

273

alginate or PA (Figure 4). This could be mainly due to the poor ability of alginate and PA to form soluble

274

complexes with Al ions. Indeed, PA was reported to only form outer-sphere complexes with Al and Fe

275

oxides.[28, 30] The less adsorption of alginate and PA by MnO2 when their concentrations are not too high

276

(Figure 1b) could be another reason, which allows little competition for surface sites.

0.010

0

10

Alginate/PA (mg/L) 20 30 40

50

3+

60

0.035

0

10

a)

0.009 mM Al

Alginate/PA (mg/L) 20 30 40 3+ 0.03 mM Al

50

60 b)

0.030

0.008

[Al3+] sorbed

0.025 0.006

0.020 0.015

0.004

AHA LHA Alginate PA

0.002 0.000

277

0

2

0.010

AHA LHA Alginate PA

0.005

4 6 8 10 AHA/LHA (mg-C/L)

12

14

0.000

0

2

4 6 8 10 AHA/LHA (mg-C/L)

12

14

12 ACS Paragon Plus Environment

Page 13 of 21

278 279 280 281 282

Environmental Science & Technology

Figure 4: Sorption of a) 0.009 mM and b) 0.03 mM of Al ions by MnO2 in the presence of different amounts of NOM. Experimental conditions: 5.0 mg/L MnO2, 0-50 mg/L NOM, 0.01 M NaCl, pH 5. Effect of Al ions on NOM adsorption by MnO2 The presence of Al ions could also affect the extent of NOM adsorbed by MnO2. As the results

283

show in Figure 5a-b, the adsorption of AHA and LHA by MnO2 increased when the concentration of

284

soluble Al was increased from 0 to 0.009 mM. There was an even higher increase in the adsorption of

285

both AHA and LHA by MnO2 when the concentration of soluble Al was increase to 0.03 mM. The higher

286

adsorption of both HAs by MnO2 in the presence of Al ions is likely due to the neutralization of MnO2

287

negative surface charge upon adsorption of the positively charged Al ions. This results in more adsorption

288

of the negatively charged HAs.[16-18] Addition of Ca2+ and Cu2+ was found to increase the amount of FA

289

adsorbed by goethite.[14, 21] There is also a likelihood of surface ternary complex formation. Given the fact

290

that Al ions can form strong complexes with both MnO2 and HA, and HA is typically only adsorbed as

291

outer-sphere complexes,[20, 28-30] the observation that Al ions slighted enhanced HA sorption is likely

292

partly due to the formation of ternary A complex (>Mn-Al-HA).[33]

Qe (mg-C/g)

8

AHA

a)

6

6

4

4

2

0

2

Qe (mg/g)

8

4

6 8 10 Ce (mg-C/L)

14 c)

0

3

4

2

2

0

10

20

30 40 Ce (mg/L)

50

10 15 Ce (mg-C/L)

25 d)

0 mM 0.009 mM 0.03 mM

0 60

20

PA

1

0 mM 0.009 mM 0.03 mM

0

5

4

6

b)

0 mM 0.009 mM 0.03 mM

0

12

Alginate

LHA

2

0 mM 0.009 mM 0.03 mM

0

293

8

0

10

20

30 40 Ce (mg/L)

50

60

13 ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 21

294 295 296

Figure 5: Effect of Al ions on the adsorption of a) AHA, b) LHA, c) alginate, and d) PA by MnO2. Conditions: 5 mg/L MnO2, 0–0.03 mM of Al ions, 0–12.5 mg-C/L AHA and LHA, 0–50 mg/L alginate and PA, 0.01 M NaCl, pH 5.0

297

Adsorption of alginate and PA by MnO2 was not affected or slightly lower upon the addition of

298

0.009 mM Al ions, but increased in the presence of 0.03 mM Al ions than in the absence of the Al ions

299

(Figure 5c-d). Together with the observation that alginate and PA did not affect Al adsorption (Figure 4),

300

there seemed to be only minor electrostatic effects between alginate/PA and Al ions in affecting each

301

other’s adsorption by MnO2.

302

Reactivity of MnO2 with Al ions and NOM

303

Rate constants (k) for triclosan oxidation by MnO2 were calculated using pseudo first-order

304

kinetics for the initial reaction period (typically < 1 h at pH 5, examples in Figure S7).[1] k will be used as

305

an indicator for MnO2 oxidative reactivity throughout this work.

306

Similar to our previous work, [23] the obtained k values were converted to term “P”. P values will

307

be used to distinguish the effects of NOM on MnO2 reactivity from those of Al2O3 or Al ions. This is

308

done by normalizing the reactivity of MnO2 in the presence of NOM (kwith_NOM) by that in the absence of

309

NOM (kwithout_NOM):

310 = P

k withNOM × 100% ………………………………………………………………………..……..(3) k withoutNOM

311

For example, kwith_NOM of MnO2+Al2O3+NOM vs kwithout_NOM of MnO2+Al2O3. P values can suggest how

312

the addition of NOM affects the reactivity of MnO2. A P value of greater than 100% would indicate an

313

enhancing effect of NOM while a P value of lower than 100% would indicate an inhibition effect of

314

NOM. Moreover, we can compare the P values of ternary mixtures to those of the respective binary

315

mixture of MnO2+NOM to see the effect of Al2O3/Al ions. For instance, if the P value of

316

MnO2+Al2O3+AHA is greater than that of MnO2+AHA under otherwise identical conditions, it suggests

317

there is less inhibition on the MnO2 reactivity in the ternary mixture than the sum of two binary mixtures:

318

MnO2+AHA and MnO2+Al2O3. In other words, there are additional interactions in the ternary mixture to

319

alleviate the inhibition effects in the binary mixtures.

14 ACS Paragon Plus Environment

Page 15 of 21

320

Environmental Science & Technology

Similar to our previous work, [23] the reactivity of MnO2 is slightly higher in the presence of 0.2

321

mg-C/L of AHA or 0.5 mg-C/L of LHA (Figures 6a-b). This is mostly due to the formation of soluble

322

HA-Mn complexes which inhibits the adsorption of the produced Mn2+ during MnO2 reaction.[4] Addition

323

of more than 2 mg-C/L of AHA and LHA increasingly inhibited the MnO2 reactivity, mostly due to more

324

HAs adsorbed by MnO2. [23] Some previous studies have made similar observations which have been

325

attributed to the blockage of MnO2 surface sites.[1] As presented in our previous paper, [23] both alginate

326

and PA did not affect MnO2 reactivity, likely because of their poor adsorption by MnO2 under the

327

experimental conditions. The effect of alginate was only seen at a very high concentration of 25 mg/L

328

when a significant amount of alginate was adsorbed (Figure 1b). 250

a)

0 0.2 0.4 2 4 10

AHA

P%

200 150

b)

0 0.5 2 4.6 12.5

LHA

100 50 0

A AH O Al 2

A AH + 3

250 c)

M 9m 0 0 0. HA +A Alginate

Al

M 3m 0 . 0 HA +A

Al

A LH

0 1 5 10 25

200

O Al 2

A LH + 3

d)

M 9m 0 0.0 A H +L PA

Al

M 3m 0 . 0 HA +L

Al

P%

150

0 1 5 10 25 50

100 50 0

329 330 331 332

. Alg

O Al 2

. Alg + 3

mM 09 0 . 0 lg. +A

Al

mM 3 0 0. lg. +A

Al

PA

A Al Al +P M M m 3 m O 3 09 Al 2 0.0 A 0.0 A +P +P

Figure 6: Effects of a) AHA, b) LHA, c) alginate, and d) PA on the oxidative reactivity of MnO2 with or without Al2O3 or Al ion. Reaction conditions: 0.1 g/L Al2O3, 0.009 or 0.03 mM Al ions, 10 µM triclosan, 5.0 mg/L MnO2, 0.01 M NaCl, 25 mM acetic buffer, and pH 5.0. The legends indicate the concentrations 15 ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 21

333 334 335 336

of NOM in mg-C/L for AHA and LHA and mg/L for alginate and PA. Error bars indicate standard deviations of duplicate experiments.

337

added to the mixture of Al ions and MnO2, MnO2 reactivity was mostly comparable to that of the

338

respective binary mixture of MnO2+HA (Figure 6a-b). This correlates well with the negligible effect of

339

HAs on the amount of Al adsorbed when HA concentrations are low (Figure 4). The addition of 2.0 to

340

12.5 mg-C/L of AHA or LHA to the mixture of Al ion (0.009 mM) and MnO2, however, yielded generally

341

greater P values than the respective mixture of MnO2+HA (Figure 6a-b). The effect was more significant

342

when Al ion concentration reached 0.03 mM. This again agrees well with the decreasing amount of Al

343

sorbed by MnO2 with increasing HA loading (Figure 4). Although the higher amount of HA adsorbed in

344

the presence of Al ions (Figure 5a-b) would likely inhibit MnO2 reactivity, it did not seem to compete

345

well with the above enhancing effect.

When only a small amount of HA (i.e., 0.2-0.4 mg-C/L of AHA or 0.5 mg-C/L of LHA) was

346

The presence of a wide range of alginate (1 to 25 mg/L) or PA (1-50 mg/L) in ternary mixtures of

347

MnO2, NOM, and Al ion, surprisingly, yielded smaller P values than those of the respective MnO2+NOM

348

(Figure 6c-d). The effect was more drastic when soluble Al concentration reached 0.03 mM. Under these

349

conditions, similar amounts of Al were sorbed in the binary vs ternary systems (Figure 4). Our previous

350

work also showed poor aggregation within MnO2, [23] so the most likely reason for the observed lower P

351

values is the enhanced adsorption of the NOMs once the oxide surfaces partially complex with Al ions

352

(Figure 5c-d). Once Al ions complex with MnO2, the surface becomes less negatively charged. This

353

would facilitate further adsorption of the negatively charged NOM and hence more inhibition of the

354

MnO2 reactivity. The presence of Cu2+ has also been reported to markedly increase FA adsorption by

355

goethite.[21]

356

Upon understanding the dominant interactions in the ternary MnO2+Al+NOM mixtures and how

357

the interactions affect MnO2 reactivity, we are now ready to examine ternary MnO2+Al2O3+NOM

358

mixtures to see how important Al ions and heteroaggregation are in affecting MnO2 reactivity in the later

359

mixtures.

360 16 ACS Paragon Plus Environment

Page 17 of 21

361

Environmental Science & Technology

Reactivity of MnO2 with Al2O3 and NOM

362

Comparison of the P values of ternary mixtures of MnO2+Al2O3+HA to those of binary mixtures

363

of MnO2+HA show comparable to slightly lower P values at 0.2 to 0.4 mg-C/L of AHA or 0.5 mg-C/L of

364

LHA (Figure 6a-b). Higher P values for the ternary mixtures, however, were observed at 2.0 to 10 mg-

365

C/L of AHA or 2.0 to 12.5 mg-C/L of LHA.

366

The effects of HAs on MnO2 reactivity in ternary mixtures with Al2O3 are very different from

367

those observed for ternary mixtures of MnO2+Al+HA (Figure 6). If complexation of Al ions plays an

368

important role in MnO2+Al2O3+HA, the effect of HAs should be comparable to that of MnO2+Al+HA.

369

The different trends observed in Figure 6 thus suggest that Al ions play a much less significant role in

370

affecting MnO2 reactivity in ternary MnO2+Al2O3+HA mixtures than in binary MnO2+Al2O3 mixtures.

371

Indeed, the trend can be well explained by the inhibited dissolution of Al2O3 in the presence of HAs

372

(Figure 3) so that only a small amount of Al ions is now available to complex with MnO2. As a result,

373

heteroaggregation between MnO2 and Al2O3 has become the dominant interaction mechanism in

374

MnO2+Al2O3+HA such that increasing the concentration of HAs to ≥ 2 mg-C/L increased the stability of

375

the oxide mixtures (or decreased the extent of heteroaggregation between the two metal oxides), as shown

376

in Figures 2 and S4. The end result is enhanced MnO2 reactivity at higher concentrations of HAs. When

377

the concentration of HA was low (0.2-0.4 mg-C/L of AHA or 0.5 mg-C/L of LHA), the enhanced extent

378

of heteroaggregation as observed in Figures 2 and S4 can also explain the slightly inhibited MnO2

379

reactivity in Figure 6a-b. Similar effects of the extent of heteroaggregation on MnO2 reactivity due to

380

NOM adsorption have been observed for MnO2 and iron oxides when heteroaggregation was the major

381

interaction mechanism. [23]

382

In the presence of 1-25 mg/L of alginate or 1-50 mg/L PA, higher P values were observed for all

383

ternary mixtures with Al2O3 than the respective binary MnO2+NOM mixture. This trend, again, is

384

opposite to what was observed for ternary MnO2+Al+NOM, suggesting the effect of Al complexation

385

with MnO2 was minor at most. Similar to the HAs, alginate and PA mainly affected the extent of

386

heteroaggregation between MnO2 and Al2O3 to affect MnO2 reactivity.

387 17 ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 21

388

Environmental Significance. Currently, the literature on the oxidative transformation of organic

389

contaminants (OCs) focuses on model systems where only pure oxidants are present. However, complex

390

environmental conditions prevent us from quantitatively extrapolating the results obtained from model

391

systems to environmental systems. As a result, we are far from developing environmental fate models

392

that can be used to predict the redox behavior of OCs in the environment. Given the fact that (1)

393

mechanistic understanding of the redox transformation of many OCs by single Mn oxides has been

394

achieved, yet (2) little information is available regarding the redox transformation of OCs in soil-water

395

environments, it is important to investigate more complex model systems that are based on the existing

396

single oxide systems but are more representative of the environment, so that further examination of the

397

environment is feasible. Mixtures of metal oxides are ubiquitous in soils and aquatic environments,

398

among which Al, Fe, and Si oxides are among the most important components mixing with Mn oxides.[34-

399

37]

400

how the redox reactivity of MnO2 was affected by the interactions. Together with our recent work on

401

binary Fe/Al/Si and Mn oxide mixtures[8] and ternary MnO2, Fe oxides, and NOM mixtures, [23] the gap

402

between single oxide systems and complex environmental systems is narrower. This makes it more

403

realistic to incorporate redox transformation, which is lacking in the existing models, into a new

404

generation of environmental fate models.

Our current work examined how an Al oxide interacted with a Mn oxide in the presence of NOM and

405

In addition to being important in redox transformation of OCs, Mn oxides, the most abundant Mn

406

form in many sedimentary environments, affect a wide range of geological processes, including anaerobic

407

degradation of organic matter, biogeochemical cycling of trace elements, and the development of anoxic

408

conditions in soils.[38-43] Our findings will also allow a more accurate modeling of the redox activity of

409

Mn oxides in geological environments, which in turn will enable better modeling of the redox cycling of

410

Mn oxides and the corresponding cycling of many other elements under geological conditions.

411 412

Acknowledgements

413

This material is based upon work supported by the National Science Foundation under Grant CBET-

414

1236517 and Instrument Grant CHE-0923077. 18 ACS Paragon Plus Environment

Page 19 of 21

Environmental Science & Technology

415 416

Supporting Information Available

417

Supporting information including Texts S1 – S5, seven figures and one table is available.

418

information is available free of charge via the Internet at http://pubs.acs.org/.

This

419 420

References

421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461

1. Zhang, H.; Huang, C.-H., Oxidative transformation of triclosan and chlorophene by manganese oxides. Environ. Sci. Technol. 2003, 37 (11), 2421-2430. 2. Zhang, H.; Huang, C.-H., Oxidative transformation of fluoroquinolone antibacterial agents and structurally related amines by manganese oxide. Environ. Sci. Technol. 2005, 39 (12), 4474-4483. 3. Zhang, H.; Huang, C.-H., Reactivity and transformation of antibacterial N-oxides in the presence of manganese oxide. Environ. Sci. Technol. 2005, 39 (2), 593-601. 4. Xu, L.; Xu, C.; Zhao, M.; Qiu, Y.; Sheng, G. D., Oxidative removal of aqueous steroid estrogens by manganese oxides. Water Res. 2008, 42 (20), 5038-5044. 5. Laha, S.; Luthy, R. G., Oxidation of aniline and other primary aromatic amines by manganese dioxide. Environ. Sci. Technol. 1990, 24 (3), 363-373. 6. Murray, J. W., The surface chemistry of hydrous manganese dioxide. J. Colloid Interf. Sci. 1974, 46 (3), 357-371. 7. Stone, A. T., Reductive dissolution of manganese(III/IV) oxides by substituted phenols. Environ. Sci. Technol. 1987, 21 (10), 979-988. 8. Taujale, S.; Zhang, H., Impact of interactions between metal oxides to oxidative reactivity of manganese dioxide. Environ. Sci. Technol. 2012, 46 (5), 2764-2771. 9. Chen, K. L.; Elimelech, M., Interaction of fullerene (C60) nanoparticles with humic acid and alginate coated silica surfaces: measurements, mechanisms, and environmental implications. Environ. Sci. Technol. 2008, 42 (20), 7607-7614. 10. Chen, K. L.; Mylon, S. E.; Elimelech, M., Aggregation kinetics of alginate-coated hematite nanoparticles in monovalent and divalent electrolytes. Environ. Sci. Technol. 2006, 40 (5), 15161523. 11. Davis, J. A., Adsorption of natural dissolved organic matter at the oxide/water interface Geochim. Cosmochim. Acta 1982, 46 (11), 2381–2393. 12. Davis, J. A., Complexation of trace metals by adsorbed natural organic matter. Geochim. Cosmochim. Acta 1984, 48 (4), 679-691. 13. 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. 14. Weng, L. P.; Koopal, L. K.; Hiemstra, T.; Meeussen, J. C. L.; Van Riemsdijk, W. H., Interactions of calcium and fulvic acid at the goethite-water interface. Geochim. Cosmochim. Acta 2005, 69 (2), 325-339. 15. Lee, Y. J.; Elzinga, E. J.; Reeder, R. J., Cu(II) adsorption at the calcite–water interface in the presence of natural organic matter: Kinetic studies and molecular-scale characterization. Geochim. Cosmochim. Acta 2005, 69 (1), 49-61. 16. Persson, P.; Karlsson, M.; Öhman, L.-O., Coordination of acetate to Al(III) in aqueous solution and at the water-aluminum hydroxide interface: a potentiometric and attenuated total reflectance FTIR study. Geochim. Cosmochim. Acta 1998, 62 (23–24), 3657-3668. 17. Sharma, P.; Ofner, J.; Kappler, A., Formation of binary and ternary colloids and dissolved complexes of organic matter, Fe and As. Environ. Sci. Technol. 2010, 44 (12), 4479-4485. 18. Strathmann, T. J.; Myneni, S. C. B., Effect of soil fulvic acid on nickel(II) sorption and bonding at the aqueous-boehmite (γ-AlOOH) interface. Environ. Sci. Technol. 2005, 39 (11), 4027-4034. 19 ACS Paragon Plus Environment

Environmental Science & Technology

462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514

Page 20 of 21

19. Christl, I.; Kretzschmar, R., Interaction of copper and fulvic acid at the hematite-water interface. Geochim. Cosmochim. Acta 2001, 65 (20), 3435-3442. 20. Johnson, S. B.; Yoon, T. H.; Slowey, A. J.; Brown, G. E., Adsorption of organic matter at mineral/water interfaces: 3. Implications of surface dissolution for adsorption of oxalate. Langmuir 2004, 20 (26), 11480-11492. 21. Weng, L.; Van Riemsdijk, W. H.; Hiemstra, T., Cu2+ and Ca2+adsorption to goethite in the presence of fulvic acids. Geochim. Cosmochim. Acta 2008, 72 (24), 5857-5870. 22. Chen, K. L.; Mylon, S. E.; Elimelech, M., Enhanced aggregation of alginate-coated iron oxide (hematite) nanoparticles in the presence of calcium, strontium, and barium cations. Langmuir 2007, 23 (11), 5920-5928. 23. Zhang, H.; Taujale, S.; Huang, J.; Lee, G.-J., Effects of NOM on oxidative reactivity of manganese dioxide in binary oxide mixtures with goethite or hematite. Langmuir 2015, 31 (9), 2790-2799. 24. Evanko, C. R.; Dzombak, D. A., Influence of structural features on sorption of NOM-analogue organic acids to goethite. Environ. Sci. Technol. 1998, 32 (19), 2846-2855. 25. Phenrat, T.; Saleh, N.; Sirk, K.; Tilton, R. D.; Lowry, G. V., Aggregation and sedimentation of aqueous nanoscale zerovalent iron dispersions. Environ. Sci. Technol. 2006, 41 (1), 284-290. 26. Phenrat, T.; Song, J. E.; Cisneros, C. M.; Schoenfelder, D. P.; Tilton, R. D.; Lowry, G. V., Estimating attachment of nano- and submicrometer-particles coated with organic macromolecules in porous media: Development of an empirical model. Environ. Sci. Technol. 2010, 44 (12), 4531-4538. 27. Chen, G.; Liu, X.; Su, C., Distinct effects of humic acid on transport and retention of TiO2 rutile nanoparticles in saturated sand columns. Environ. Sci. Technol. 2012, 46 (13), 7142-7150. 28. Johnson, S. B.; Yoon, T. H.; Brown, G. E., Adsorption of organic matter at mineral/water interfaces: 5. Effects of adsorbed natural organic matter analogues on mineral dissolution. Langmuir 2005, 21 (7), 2811-2821. 29. Johnson, S. B.; Yoon, T. H.; Kocar, B. D.; Brown, G. E., Adsorption of organic matter at mineral/water interfaces. 2. Outer-sphere adsorption of maleate and implications for dissolution processes. Langmuir 2004, 20 (12), 4996-5006. 30. Yoon, T. H.; Johnson, S. B.; Brown, G. E., Adsorption of organic matter at mineral/water interfaces. IV. Adsorption of humic substances at boehmite/water interfaces and impact on boehmite dissolution. Langmuir 2005, 21 (11), 5002-5012. 31. Tonkin, J. W.; Balistrieri, L. S.; Murray, J. W., Modeling sorption of divalent metal cations on hydrous manganese oxide using the diffuse double layer model. Appl. Geochem. 2004, 19 (1), 29-53. 32. Yamamoto, Y.; Takahashi, Y.; Shimizu, H., Speciation of iron in humic substances by X-ray absorption fine structure and its effect on the complexation between humic substances and trace metal ions. Chem. Lett. 2009, 38 (3), 278-279. 33. Huang, C. P.; Rhoads, E. A.; Hao, O. J., Adsorption of Zn(II) onto hydrous aluminosilicates in the presence of EDTA. Water Res. 1988, 22 (8), 1001-1009. 34. Anderson, P. R.; Benjamin, M. M., Surface and bulk characteristics of binary oxide suspensions. Environ. Sci. Technol. 1990, 24 (5), 692-698. 35. Harvey, O. R.; Rhue, R. D., Kinetics and energetics of phosphate sorption in a multi-component Al(III)-Fe(III) hydr(oxide) sorbent system. J. Colloid Interf. Sci. 2008, 322 (2), 384-393. 36. Meng, X.; Letterman, R. D., Effect of component oxide interaction on the adsorption properties of mixed oxides. Environ. Sci. Technol. 1993, 27 (5), 970-975. 37. Nano, G. V.; Strathmann, T. J., Ferrous iron sorption by hydrous metal oxides. J. Colloid Interf. Sci. 2006, 297 (2), 443-454. 38. Banwart, S. A., Reduction of iron(III) minerals by natural organic matter in groundwater. Geochim. Cosmochim. Acta 1999, 63 (19/20), 2919-2928. 39. Bond, D. R.; Lovley, D. R., Reduction of Fe(III) oxide by methanogens in the presence and absence of extracellular quinones. Environ. Microbiol. 2002, 4 (2), 115-124. 40. Lovley, D. R., Microbial reduction of iron, manganese, and other metals. Adv. Agron. 1995, 54, 175231. 41. Lovley, D. R.; Holmes, D. E.; Nevin, K. P., Dissimilatory Fe(III) and Mn(IV) reduction. Adv. Microb. Physiol. 2004, 49, 219-286. 20 ACS Paragon Plus Environment

Page 21 of 21

515 516 517 518 519 520

Environmental Science & Technology

42. Lower, S. K.; Hochella, M. F., Jr.; Beveridge, T. J., Bacterial recognition of mineral surfaces: Nanoscale interactions between Shewanella and alpha-FeOOH. Science 2001, 292 (5520), 13601363. 43. Van Cappellen, P.; Wang, Y., Cycling of iron and manganese in surface sediments: A general theory for the coupled transport and reaction of carbon, oxygen, nitrogen, sulfur, iron, and manganese. Am. J. Sci 1996, 296 (3), 197-243.

521 522

21 ACS Paragon Plus Environment