Molecular-Scale Study of Aspartate Adsorption on Goethite and

Feb 12, 2016 - Abstract Image. Knowledge of the interfacial interactions between aspartate and minerals, especially its competition with phosphate, is...
0 downloads 10 Views 746KB Size
Subscriber access provided by UNIV OF CALIFORNIA SAN DIEGO LIBRARIES

Article

Molecular-Scale Study of Aspartate Adsorption on Goethite and Competition with Phosphate Yanli Yang, Shengrui Wang, Yisheng Xu, Binghui Zheng, and Jingyang Liu Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.5b05450 • Publication Date (Web): 12 Feb 2016 Downloaded from http://pubs.acs.org on February 13, 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 32

Environmental Science & Technology



Molecular-Scale Study of Aspartate Adsorption on Goethite and Competition



with Phosphate



Yanli Yang,†, ‡ Shengrui Wang,*, †, ‡ Yisheng Xu,† Binghui Zheng,† and Jingyang Liu§

4  †



State Key Laboratory of Environmental Criteria and Risk Assessment, Chinese Research Academy of Environmental Sciences, Beijing 100012, China

6  7  ‡



State Environmental Protection Key Laboratory for Lake Pollution Control,



Research Center of Lake Eco-environment, Chinese Research Academy of

10 

Environmental Sciences, Beijing 100012, China

11  §

12 

State Key Laboratory of Environmental Protection Ecology Industry,

Chinese Research Academy of Environmental Sciences, Beijing 100012, China

13  14  15  16  17  18  19  20  21 

1   

ACS Paragon Plus Environment

Environmental Science & Technology

22 

ABSTRACT

23 

Knowledge of the interfacial interactions between aspartate and minerals, especially its

24 

competition with phosphate, is critical to understanding the fate and transport of amino

25 

acids in the environment. Adsorption reactions play important roles in the mobility,

26 

bioavailability, and degradation of aspartate and phosphate. Attenuated total reflectance

27 

Fourier-transform infrared (ATR-FTIR) measurements and density functional theory

28 

(DFT) calculations were used to investigate the interfacial structures and their relative

29 

contributions in single-adsorbate and competition systems. Our results suggest three

30 

dominant mechanisms for aspartate: bidentate inner-sphere coordination involving both

31 

α- and γ-COO−, outer-sphere complexation via electrostatic attraction and H-bonding

32 

between aspartate NH2 and goethite surface hydroxyls. The interfacial aspartate is

33 

mainly governed by pH and is less sensitive to changes of ionic strength and aspartate

34 

concentration. The phosphate competition significantly reduces the adsorption capacity

35 

of aspartate on goethite. Whereas phosphate adsorption is less affected by the presence

36 

of aspartate, including the relative contributions of diprotonated monodentate,

37 

monoprotonated bidentate, and nonprotonated bidentate structures. The adsorption

38 

process facilitates the removal of bioavailable aspartate and phosphate from the soil

39 

solution as well as from the sediment pore water and the overlying water.  

40 

 

41 

 

42 

 

2   

ACS Paragon Plus Environment

Page 2 of 32

Page 3 of 32

Environmental Science & Technology

43 

 

TOC Art

44 

45 

 

46 

 

47 

 

48 

 

49 

 

50 

 

51 

 

52 

 

53 

 

54 

 

55 

 

56 

 

57 

 

58 

 

59 

 

60 

  3   

ACS Paragon Plus Environment

Environmental Science & Technology

61 

INTRODUCTION

62 

Amino acids are an important class of chemicals in the environment, which are

63 

widely spread in soils, sediments, and natural waters. The net charge of amino acids

64 

varies with pH because of the deprotonation of NH3+ and COOH, which leads to a series

65 

of interesting adsorption behaviors.1 The adsorption of amino acids on minerals can

66 

retard their migration and alter their bioavailability in soils and sediments, and the

67 

interfacial structures may determine their mechanism and difficulty of degradation.

68 

Therefore, knowledge of amino acids’ molecular-level interactions is critical to

69 

understanding their fate and transport in the environment as well as the geochemical

70 

cycle of nitrogen. Notably, amino acids are considered to be the fundamental foundation

71 

for life. In-depth studies of their adsorption on minerals are needed to evaluate

72 

biomineralization, the viability of metal implants in the human body, and the origin and

73 

early evolution of life on earth.1-3

74 

Aspartate is the simplest acidic amino acid containing one NH3+ and two COOH

75 

groups, which is frequently detected at high levels in natural waters.4 In contrast to the

76 

extensive reports of dicarboxylate at mineral/aqueous interfaces, only few studies have

77 

investigated the adsorption of aspartate on TiO2,5-10 Al2O3,3 and clay minerals.11, 12

78 

Interestingly, wide diversities have been observed in the mechanisms and interfacial

79 

configurations depending on the mineral phase. Recent research on aspartate adsorption

80 

has mainly focused on TiO2, although it usually exists as a minor accessory phase in

81 

most rocks and soils.3 Parikh et al. observed that aspartate reacted with TiO2 via outer-

82 

sphere coordination, without inducing peak shifts indicative of covalent bonding.5 In 4   

ACS Paragon Plus Environment

Page 4 of 32

Page 5 of 32

Environmental Science & Technology

83 

contrast, other reports have indicated inner-sphere complexation.6-9 Jonsson et al.

84 

resolved a bridging bidentate structure through both aspartate COO− and an outer-

85 

sphere (or H-bonded) species.10 Notably, adsorption mechanisms are largely

86 

determined by the adsorbate structure and mineral phase. Therefore, the adsorption of

87 

aspartate on different minerals must still be investigated. Aluminum oxides and iron

88 

oxides are much more relevant than TiO2 to the fields of earth and environmental

89 

science. Greiner et al. observed that aspartate attached to the γ-Al2O3/D2O interface via

90 

bidentate and tetradentate coordination and outer-sphere complexation.3 Because of its

91 

high specific surface area, goethite has strong affinities for small organic acids,

92 

oxyanions, and heavy metals.13-16 However, very little is known about the adsorption

93 

behaviors of amino acids on this model mineral.17 Insight into the molecular-level

94 

process of aspartate on goethite motivates our experimental and theoretical research.

95 

Unlike the simple systems with one amino acid and one mineral phase, the natural

96 

environment is more complicated involving abundant oxyanions, metal ions, dissolved

97 

organic matter, etc. The importance of competitive adsorption (or coadsorption) with

98 

coexisting substance cannot be overstated.18 The adsorption of glutamate, lysine, and

99 

aspartate has recently been studied in the presence of Ca2+ (and Mg2+) via batch

100 

adsorption experiments and surface complexation modeling.1, 19 Limited knowledge is

101 

available for the molecular-level mechanisms of amino acids on minerals in

102 

competition with phosphate, particularly the in-situ adsorption process. The insufficient

103 

study inspires our research to reveal the interfacial structures, the adsorption capacity,

104 

and relative content of each surface complex of aspartate and phosphate in binary5   

ACS Paragon Plus Environment

Environmental Science & Technology

105 

adsorbate systems. The interfacial configurations of phosphate on goethite have long

106 

been sought but without consistent conclusions.20-23 Most reports supported the inner-

107 

sphere coordination in different structures, while Kubicki proposed the coexistence of

108 

inner- and outer-sphere complexation.23 Furthermore, there have been a few studies

109 

with respect to the phosphate adsorption in the presence of arsenate, U(VI), Cr(VI),

110 

humic acid, and organic pollutants.24-28 The existing reports provide valuable

111 

information for reference and comparison.

112 

The aspartate adsorption on goethite was explored under the effects of pH, ionic

113 

strength, aspartate concentration, and phosphate competition. The objective of this

114 

study was to reveal the interfacial structures and relative content of each complex at the

115 

molecular level, which have great impact on the mobility, bioavailability, and

116 

degradation of the adsorbates. Besides the ATR-FTIR measurements, theoretical

117 

calculations were used to determine the configurations of adsorbed aspartate and

118 

phosphate. The two-dimensional correlation spectroscopy (2D-COS) was employed to

119 

identify the number of interfacial phosphate structures and peaks belonging to each.

120 

The relative contributions of each surface complex were analyzed because of their

121 

sensitive responses to spectral changes in different experimental conditions. This

122 

research provides new and complementary information for understanding the

123 

adsorption of aspartate and phosphate on metal oxides and can be used to predict and

124 

describe the behaviors of amino acids in the environment.

6   

ACS Paragon Plus Environment

Page 6 of 32

Page 7 of 32

125 

Environmental Science & Technology

EXPERIMENTAL SECTION

126 

1. Materials. L-Aspartic acid and NaH2PO4 were purchased from Sigma-Aldrich.

127 

All of the chemicals were of analytical or guaranteed reagent grade and were used as

128 

received. The samples were prepared in Milli-Q water that was boiled for 60 min and

129 

cooled with N2 purging to remove CO2. Goethite was prepared according to the

130 

procedure described elsewhere.29 The point of zero charge (PZC) and Brunauer-

131 

Emmett-Teller (BET) surface area of goethite were determined as pH 8.9 and 84.7 m2/g,

132 

respectively.18, 29 The particles are well-crystallized needles with a length of 100−200

133 

nm, which are mainly terminated by (110) and (100) planes.29, 30

134 

2. ATR-FTIR Spectroscopy Study. The ATR spectra were recorded with a Perkin-

135 

Elmer Spectrum 100 instrument that was equipped with an MCT detector, a constant

136 

flow pump, and a horizontal ATR accessory (PIKE Technologies, USA). The flow cell

137 

was fitted with a 45° ZnSe or Ge crystal. The usable pH for the ZnSe crystal was chosen

138 

within the range of 5−9 to avoid etching. The Ge crystal was used with pH levels less

139 

than 5 or greater than 9. Spectra acquisition, subtraction, normalization, and baseline

140 

correction were performed using the Spectrum software (Perkin-Elmer, Inc., USA). All

141 

of the spectra were collected with 256 scans at 4 cm−1 resolution to reduce noise. The

142 

peak numbers and positions were justified using the second derivative. Curve-fitting

143 

analysis of the overlapping peaks was conducted using Gaussian line shapes.31, 32

144 

The spectra of aspartate and phosphate solutions were obtained by subtracting the

145 

spectrum of background electrolyte at the same pH from the sample spectrum. The

146 

interfacial aspartate and phosphate were measured with a goethite film on the ATR 7   

ACS Paragon Plus Environment

Environmental Science & Technology

147 

crystal. The film was coated onto the crystal by applying 1 mL of 1 g/L goethite

148 

suspension and drying it in an oven at 50 °C for 1 h. Prior to use, the film was gently

149 

rinsed with Milli-Q water to remove loosely deposited particles. The background

150 

electrolyte at a predetermined pH was passed through the flow cell at a rate of 0.25

151 

mL/min until equilibrium was established, namely there was no further change in the

152 

spectra. The background spectrum was collected, and the solution was then changed to

153 

the sample with the same pH and ionic strength. The interfacial spectra were recorded

154 

as a function of time for 60 min. In single-adsorbate systems, the total concentrations

155 

of aspartate solutions ranged from 1 to 10 mM to obtain high-quality spectra. In

156 

comparison with the absorbance of interfacial aspartate, contributions from the

157 

dissolved species are negligible. The dynamic process at pH 6 was also investigated by

158 

flowing 0.1 M NaCl solution after 1 mM aspartate adsorption for ~30 min. In

159 

competition systems, aspartate and phosphate at the concentration of 1 mM was

160 

analyzed in 0.1 M NaCl solution. The slight dissolution of goethite had no detectable

161 

effect during the flow-cell measurements.

162 

3. Quantum Chemical Calculations. Geometry optimization and frequency

163 

calculations were performed using the Gaussian 03 program33 with the hybrid DFT

164 

B3LYP method. The clusters consisting of one single, two or three edge-sharing

165 

Fe(III)-octahedra were used to model the optimized geometries of aspartate and

166 

phosphate on goethite surface. The cluster size was selected to satisfy the basic

167 

demands for structure calculations of the interfacial configurations. For example, the

168 

aspartate bidentate structure involving both COO− was simulated with a three edge8   

ACS Paragon Plus Environment

Page 8 of 32

Page 9 of 32

Environmental Science & Technology

169 

sharing octahedra. The distance between the neighbouring surface Fe atoms is similar

170 

to that between the two O atoms in one aspartate COO−, but is much smaller than that

171 

from different COO−. These small cluster models are efficient for frequency

172 

calculations of small organic compounds and oxyanions on minerals.29, 31, 34-36 The

173 

clusters can reproduce main features of the optical response at the interface, though the

174 

small unit cells do not provide complete representations of the mineral surfaces.37, 38

175 

Moreover, this technique is computationally much more tractable than the periodic slab

176 

models which have not been extensively tested for the ability to calculate the vibrational

177 

spectra of surface complexes on minerals.23 The solvation effect was considered by

178 

placing explicit H2O around the dissolved and adsorbed species in a gas phase.31, 34, 36

179 

The number of H2O molecules for inner-sphere coordination was investigated to

180 

confirm the accuracy of frequency calculations and configuration determinations. The

181 

frequencies were calculated using the 6-31+G(d, p) basis set on C, H, O, N, and P with

182 

a scale factor of 0.964, coupled to a LanL2DZ basis set on Fe with a scale factor of

183 

0.961.39

184 

4. Two-Dimensional Correlation Spectroscopy. The 2D-COS can greatly enhance

185 

the resolution of highly overlapping peaks and facilitate the assignment of peaks

186 

belonging to each complex, which has successfully been applied to the adsorption

187 

systems of small organic compounds on goethite.14, 29, 40 The dynamic spectra in the

188 

region 1300−900 cm−1 from 10 to 60 min were baseline-corrected and smoothed to

189 

calculate the synchronous and asynchronous plots with the 2D Shige program (Shigeaki

190 

Morita, Kwansei-Gakuin University, 2004−2005). The averaged spectrum was set as a 9   

ACS Paragon Plus Environment

Environmental Science & Technology

191 

reference.29, 41 In the synchronous spectra, an auto peak is responsible for the changes

192 

of peak intensity over time, while a cross peak provides the response to the time

193 

perturbation at two different bands. A positive cross peak arises when the two bands

194 

increase or decrease simultaneously, whereas a negative cross peak demonstrates the

195 

opposite change in their peak intensities. The asynchronous spectra do not have auto

196 

peaks. An asynchronous cross peak indicates the uncorrelated response of two bands,

197 

which originate from different surface complexes or moieties of the same complex in

198 

different molecular environments.29

199 

RESULTS AND DISCUSSION

200 

1. ATR Spectra and DFT Calculations of Dissolved Aspartate. Aspartic acid

201 

contains one NH3+ group with pKa3 = 9.8 and two COOH groups with pKa1 = 2.1 and

202 

pKa2 = 3.9.5,

203 

monoanionic, and dianionic species, respectively (Figure 1 and Figure S1 in the

204 

Supporting Information, SI). On the basis of the second derivative, more frequencies

205 

were observed within the 1700−1500 cm−1 range than in previous studies. Curve-fitting

206 

results of the overlapping peaks are of high quality with r2 more than 0.99. The peak

207 

assignments are on the basis of DFT calculations (Table S1). Our correlation analysis

208 

shows a reasonable agreement between the theoretical frequencies and the experimental

209 

data (SI Tables S2−S3 and Figure S2).

6

The solution spectra at pH 3, 6, and 11 represent the zwitterionic,

10   

ACS Paragon Plus Environment

Page 10 of 32

Page 11 of 32

Environmental Science & Technology

210  211  212  213 

Figure 1. (A) Optimized structures of the dissolved aspartate. Eleven explicit H2O molecules are not

214 

At pH 3, aspartate is in the zwitterionic form with protonated γ-COOH and NH3+.

215 

The γ-COOH vibrations contribute to a band at 1724 cm−1 associated with ν(C=O),5, 42

216 

a band at 1243 cm−1 associated with coupled ν(C−O) and δ(O−H),42 and a band at 1351

217 

cm−1 associated with νs(γ-COO−). The δas(NH3+), δs(NH3+),42 and ρr(NH3+) modes are

218 

observed at 1637, 1511, and 1272 cm−1, respectively. The two bands at 16015, 42 and

219 

1396 cm−1 are attributed to the asymmetric and symmetric ν(α-COO−), respectively.

220 

The δ(CH2) scissor band appears at 1416 cm−1, and the ρ(CH2) wag appears near 1213

221 

cm−1. The 1305 cm−1 peak is attributed to δ(CH).42

shown for clarity. (B) Curve-fitting analysis of the solution spectra for 0.1 M aspartate in 0.1 M NaCl. The spectra were normalized to the most intense peak.

The solution spectra exhibit significant changes with increasing pH. Deprotonation

222 

11   

ACS Paragon Plus Environment

Environmental Science & Technology

223 

of γ-COOH at pH 6 increases the symmetry in the molecular structure, which results in

224 

downward shifts of νas(α-COO−) from 1601 to 1572 cm−1 and δs(NH3+) from 1511 to

225 

1476 cm−1. Furthermore, the νas(γ-COO−) mode begins to appear at 1596 cm−1.

226 

Interestingly, the δas(NH3+), νs(α-COO−), and νs(γ-COO−) vibrations are affected less

227 

by COOH deprotonation. The two bands at 1417 and 1308 cm−1 are assigned to CH2

228 

scissoring and wagging, respectively. The band at 1327 cm−1 is due to δ(CH), and the

229 

1233 cm−1 peak is due to NH3+ rocking.

230 

For the aspartate dianion at pH 11, the NH3+ deprotonation further increases the

231 

similarity between α- and γ-COO−. Accordingly, their asymmetric stretches shift

232 

downward to 1550 and 1568 cm−1, respectively. NH2 scissoring occurs at 1598 cm−1,

233 

and the rocking mode is centered at 1325 cm−1. The vibrations associated with νs(α-

234 

COO−) at 1395 cm−1, νs(γ-COO−) at 1359 cm−1, δ(CH2) scissoring at 1419 cm−1, and

235 

ρω(CH2) at 1307 cm−1 are not sensitive to NH3+ deprotonation.

236 

2. ATR Spectra of Adsorbed Aspartate at the Goethite/Aqueous Interface. The

237 

interfacial spectra with 1 mM aspartate in 0.1 M NaCl exhibit different characteristics

238 

in peak numbers, locations, and profiles at pH 3, 6, and 11, which suggests disparate

239 

complexation mechanisms at different pH (Figure 2). Notably, the interfacial aspartate

240 

at a certain pH remains unchanged during adsorption, as indicated by the same peak

241 

positions in the dynamic spectra (SI Figure S3).

12   

ACS Paragon Plus Environment

Page 12 of 32

Page 13 of 32

Environmental Science & Technology

242  243  244 

Figure 2. Curve-fitting analysis of the interfacial spectra collected at 20 min for 1 mM aspartate in 0.1

245 

Compared with the spectrum of aspartate zwitterion, the interfacial spectrum at pH

246 

3 exhibits significant changes in peak locations and shapes, which indicates inner-

247 

sphere coordination.31 The absence of γ-COOH vibrations at 1724 and 1243 cm−1

248 

suggests the involvement of distal carboxyl upon adsorption. The surface complex with

249 

deprotonated COO− has also been proposed for glutamate and carboxylic acids on metal

250 

oxides.3, 5, 6, 43, 44 The dianionic-type configurations could be excluded because of the

251 

persistent existence of δs(NH3+) at 1518 cm−1 after adsorption. The nonparticipation of

252 

the NH3+ group in surface complexation is supported by most previous studies of amino

253 

acid adsorption.3,

254 

mostly positive sites of goethite at pH 3 (below pHpzc 8.9). The strong adsorption of

255 

aspartate on goethite is primarily attributed to the γ-COOH group because little

256 

adsorption on metal oxides is observed for amino acids with neutrally charged side

M NaCl. The spectra were normalized to the most intense peak.

5, 6

The NH3+ group experiences electrostatic repulsion from the

13   

ACS Paragon Plus Environment

Environmental Science & Technology

257 

chains or for carboxylic acids with a single COOH and no other participating groups.3,

258 

6, 17

259 

could be ambiguous when characterized on the sole basis of ATR-FTIR analysis,

260 

thereby necessitating in-depth DFT calculations for theoretical confirmation. Notably,

261 

no apparent outer-sphere adsorption is detected with 1 mM aspartate in 0.1 M NaCl at

262 

pH 3. One reason for this lack of adsorption is that the protonated γ-COOH is not

263 

electrostatically attracted to the positively charged goethite surface. Another reason is

264 

that the deprotonated α-COO− and adjacent NH3+ can preclude electrostatic binding to

265 

the mineral surface.6

However, further differentiation between the bidentate and monodentate structures

266 

The peak numbers and positions of the interfacial spectrum at pH 6 resemble those

267 

of the solution spectrum of aspartate monoanion, except for the two additional bands at

268 

1517 and 1285 cm−1. The slight shifts of main peaks are due to the smaller distortions

269 

of aspartate upon adsorption. The nearly unchanged COO− stretches (1600, 1564, 1389,

270 

and 1356 cm−1) suggest the weak outer-sphere coordination5, 31, 43 via electrostatic

271 

interactions that occur between the monoanionic aspartate and the positive sites of

272 

goethite. The goethite surface herein still has a net amount of positive charge (below

273 

pHpzc), though the elevated pH reduces the relative concentration of positive to neutral

274 

to negative sites. The significant adsorption is closely associated with the electrostatic

275 

attraction via aspartate γ-COO−, which facilitates the further attachment of α-COO−, as

276 

evidenced by distinct variations in the relative intensities of νs(α-COO−) at 1389 cm−1.

277 

The charge distribution of the monoanionic aspartate could be rearranged when it is

278 

attracted to the goethite surface as a result of the different molecular environment from 14   

ACS Paragon Plus Environment

Page 14 of 32

Page 15 of 32

Environmental Science & Technology

279 

the solution phase. Notably, the bands at ~1517 and 1285 cm−1 were also detected in

280 

the interfacial aspartate at pH 3, which indicates the same origin of inner-sphere

281 

coordination. The dominant outer-sphere complexation is further confirmed by the

282 

dramatically decreased peak intensities along with the flowing of back ground

283 

electrolyte (SI Figure S4).

284 

The weak absorbance in the poor spectrum at pH 11 indicates unfavorable adsorption

285 

capacity, which has also been observed for aspartate on TiO2 and Al2O3 under strong

286 

basic conditions.3, 6, 10 The bands at 1600−1300 cm−1 remain nearly unchanged upon

287 

adsorption, which rules out inner-sphere complexation. The electrostatic outer-sphere

288 

coordination is also impossible due to the repulsion between the COO− groups in

289 

aspartate dianion and goethite surface with net negative charge (above pHpzc 8.9). The

290 

slight adsorption can be attributed to the H-bonding between aspartate NH2 and goethite

291 

surface hydroxyl groups. Furthermore, these H-bonds are weakened by the electrostatic

292 

repulsion from aspartate COO−. The weak adsorption and strong interference from

293 

H−O−H bending of water result in the poor-quality spectra, especially in the range of

294 

1650−1500 cm−1 (SI Figure S3). Consequently, we focused on investigating aspartate

295 

adsorption at pH 3 and 6 in this study.

296 

3. Effects of Ionic Strength and Aspartate Concentration on Adsorption. The

297 

interfacial spectra with 1 mM aspartate in Figure 3 exhibit sensitive responses to the

298 

decreased ionic strength (0.01 M NaCl) at pH 3 and 6 (SI Figure S5). The spectra at pH

299 

3 in 0.01 and 0.2 M NaCl both resolve the same inner-sphere complex as that in 0.1 M

300 

NaCl, as demonstrated by the main peak positions in both spectra being identical to 15   

ACS Paragon Plus Environment

Environmental Science & Technology

301 

those of the spectrum in 0.1 M NaCl solution. The increased NaCl concentration (0.2

302 

M) has no detectable effect on the peak locations and profiles of the adsorbed spectrum.

303 

In contrast, the overlapping peaks from 1700 to 1550 cm−1 become broader at the lower

304 

ionic strength (0.01 M NaCl) because of the appearance of two bands at 1598 and 1567

305 

cm−1. Notably, these two peaks are similar to those of the monoanionic aspartate, which

306 

suggests electrostatic outer-sphere coordination between the positively charged

307 

goethite and the small amount of aspartate monoanion (SI Figure S1).

308  309  310 

Figure 3. Curve-fitting analysis of the interfacial spectra collected at 20 min with 1 mM aspartate in 0.01

311 

At pH 6, aspartate adsorption is governed by the dominant electrostatic outer-sphere

312 

complexation and slight inner-sphere coordination. The peak locations remain nearly

and 0.2 M NaCl. The spectra were normalized to the most intense peak.

16   

ACS Paragon Plus Environment

Page 16 of 32

Page 17 of 32

Environmental Science & Technology

313 

the same at different NaCl concentrations. Interestingly, the contribution of the ~1518

314 

cm−1 peak to the interfacial spectrum increases along with elevated ionic strength,

315 

whereas the other frequencies are much more insensitive. The peak-area ratio can be

316 

used to reflect the relative adsorption capacity between outer-sphere and inner-sphere

317 

coordination.29 The ratios in the peak area of ~1484 cm−1 (outer-sphere complex) to

318 

~1518 cm−1 (inner-sphere complex) are 4.23, 2.22, and 1.54 for 0.01, 0.1, and 0.2 M

319 

NaCl, respectively. At higher ionic strengths, the reduced contribution of outer-sphere

320 

complexation to the total aspartate adsorption is mainly due to the weakened

321 

electrostatic attraction between the dissolved aspartate and the goethite surface. The

322 

diffuse layer thickness decreases with increasing electrolyte ion density.45

323 

In this study, we also investigated the effect of aspartate concentration on the

324 

interfacial structures and their relative contributions (SI Figure S6). The peak locations

325 

and profiles of the spectra for 5 and 10 mM aspartate at pH 3 resemble those for 1 mM

326 

aspartate, which indicates the same inner-sphere complexation (Figures 2 and 4).

327 

Furthermore, no evidence of an additional surface complex (appearance of new bands)

328 

is detected. At pH 6, the dominant outer-sphere complexation and unfavorable inner-

329 

sphere coordination are also resolved in the spectra for 5 and 10 mM aspartate because

330 

the frequencies are identical to those for 1 mM aspartate. Interestingly, the elevated

331 

aspartate concentration slightly enhances the inner-sphere contribution to the overall

332 

adsorption. The peak-area ratios between ~1518 and ~1484 cm−1 are 0.45, 0.65, and

333 

0.72 for 1, 5, and 10 mM aspartate, respectively.

17   

ACS Paragon Plus Environment

Environmental Science & Technology

334  335  336 

Figure 4. Curve-fitting analysis of the interfacial spectra collected at 20 min with 5 and 10 mM aspartate

337 

4. DFT Calculations of Adsorbed Aspartate on Goethite. Seven possible

338 

interfacial configurations, including two mononuclear monodentate (M-M), two

339 

mononuclear bidentate (M-B), and three binuclear bidentate (B-B) structures, were

340 

calculated to confirm the proposed inner-sphere complex (Figure 5). The M-M, M-B,

341 

B-B I, and B-B II structures coordinate to the surface iron via single aspartate α- or γ-

342 

COO−. The B-B III complex binds to two Fe atoms while involving both α- and γ-COO−.

343 

As suggested by Ha and Hwang et al., the optimized configuration can be determined

344 

by the agreement between the experimental and theoretical vibrations.31, 44 The results

345 

indicate that the B-B III structure fits much better with the observed inner-sphere

346 

complex than the other configurations (SI Tables S4−S6 and Figures S7−S8). Peak

in 0.1 M NaCl. The spectra were normalized to the most intense peak.

18   

ACS Paragon Plus Environment

Page 18 of 32

Page 19 of 32

Environmental Science & Technology

347 

assignments of the B-B III complex are on the basis of DFT calculations. The δas(NH3+)

348 

and δs(NH3+) deformations contribute to the bands at 1649 and 1518 cm−1, respectively.

349 

The νas(α-COO−) vibrations occur at 1608 cm−1, and the νas(γ-COO−) vibrations occur

350 

near 1578 cm−1. The two peaks at 1421 and 1394 cm−1 are both ascribed to the mixed

351 

symmetric stretches of α- and γ-COO−. The bands at 1448 and 1310 cm−1 correspond

352 

to CH2 scissoring and wagging, respectively. The band at 1349 cm−1 is caused by δ(CH),

353 

and the 1285 cm−1 peak is caused by ρr(NH3+). The DFT calculations of outer-sphere

354 

coordination were also considered in this study. Our results accurately simulate the

355 

observed frequencies of electrostatic and H-bonded outer-sphere complexes (SI Tables

356 

S5−S6 and Figure S8). Because of the slight distortions upon adsorption, peak

357 

assignments of the two complexes are consistent with those of the corresponding

358 

dissolved species.

359  360  361 

Figure 5. Optimized configurations of interfacial aspartate and phosphate in mononuclear monodentate (M-M), mononuclear bidentate (M-B), and binuclear bidentate (B-B) structures. Explicit H2O molecules 19   

ACS Paragon Plus Environment

Environmental Science & Technology

362 

are not shown for clarity.

363 

5. Phosphate Only Adsorption on Goethite. The interfacial spectra with 1 mM

364 

phosphate are of the same peak numbers and locations at pH 3 and 6, which

365 

demonstrates the identical surface complexes (Figure 6 and SI Figure S9). The dramatic

366 

shape differences at these two pH indicate the uneven content of each configuration,

367 

which merits further analysis on the overlapping peaks with 2D-COS. As detailedly

368 

discussed in SI (Figures S10−S11 and Table S7), three frequency groups are resolved

369 

at both pH 3 and 6, namely (A) ~1253, 1222, 1128, and 1008 cm−1; (B) ~1172, 1073,

370 

and 969 cm−1; and (C) ~1096, 1046, and 939 cm−1. In comparison with those of the

371 

dissolved H2PO4− and HPO42− (SI Figures S12−S13), the interfacial spectra experience

372 

distinct peak shifts upon adsorption, which indicates the inner-sphere coordination.31

373 

Our DFT results suggest that the combination of diprotonated M-M (group A),

374 

monoprotonated B-B (group B), and nonprotonated B-B (group C) structures can well

375 

describe the interfacial phosphate on goethite (SI Tables S8−S10 and Figures S14−S15).

376 

The optimized configurations are different from those reported in previous studies,

377 

which provide complementary insight into the inconsistent mechanisms of phosphate

378 

adsorption (SI Table S11).20-23 For example, Luengo proposed that the adsorbed

379 

phosphate on goethite were in the monoprotonated and nonprotonated bidentate forms,

380 

whereas Kubicki supported the monodentate and bidentate HPO42− complexes as well

381 

as the outer-sphere coordination.21, 23

20   

ACS Paragon Plus Environment

Page 20 of 32

Page 21 of 32

382  383  384  385  386 

Environmental Science & Technology

Figure 6. (A–D) Curve-fitting analysis of the interfacial spectra collected at 20 min with 1 mM phosphate in single- and binary-adsorbate systems (I = 0.1 M NaCl). (E) The interfacial spectra collected at 10 min with 1 mM aspartate in competition with 1 mM phosphate in 0.1 M NaCl. The spectra were normalized to the most intense peak.

387 

The significant spectral changes motivate our further research on the effect of pH on

388 

the relative contents of each complex. The peak-area ratios of the bands at about 1008,

389 

1073, and 1046 cm–1 are 5.69:1:1.06 and 4.13:1:4.24 at pH 3 and 6, respectively. The

390 

nonprotonated B-B complexation (1046 cm–1) becomes more important at pH 6. The

391 

P–OH deprotonation is easier at high pH, which facilitates the binding of O atom in P–

392 

OH to the surface iron. The relative contribution of diprotonated M-M structure (1008

393 

cm–1) decreases markedly with the increasing pH. However, it is still considerable at

394 

pH 6 where H2PO4− is not the dominant species in solution phase (90% at pH 3 and 10%

395 

at pH 6, SI Figure S13). In one, more phosphate can be adsorbed on the limited goethite 21   

ACS Paragon Plus Environment

Environmental Science & Technology

396 

surface sites via M-M coordination. In another, the diprotonated complex is stabilized

397 

by the monoprotonated and nonprotonated B-B structures via H-bonding.

398 

6. Competitive Adsorption between Aspartate and Phosphate. When competition

399 

with each other, the structures of adsorbed aspartate and phosphate are identical to those

400 

in their single-adsorbate systems (Figure 6). Within the range of 1300−900 cm−1, the

401 

interfacial spectra with 1 mM aspartate and phosphate are of exactly similar peak

402 

numbers and locations to those of the individual phosphate adsorption on goethite.

403 

Interestingly, poor spectral quality is observed in the 1800−1300 cm−1 region, which in

404 

general resembles the features of aspartate only adsorption at the same pH.

405 

As for aspartate, the rather low absorbance indicates the unfavorable adsorption as a

406 

result of phosphate competition. Besides the dramatically reduced capacity in the initial

407 

adsorption phase, the weaker affinity of aspartate on goethite is demonstrated by the

408 

decreasing peak intensities along with the reaction time (SI Figure S16). Notably, the

409 

influencing mechanisms of phosphate competition are different at pH 3 and 6. At pH 3,

410 

most of the goethite surface sites are occupied by phosphate, and the adsorbed aspartate

411 

B-B III complex is partly removed by the phosphate inner-sphere coordination. The

412 

slight adsorption at pH 6 is attributed to the weakened electrostatic attractions between

413 

aspartate monoanion and goethite surface. The net positive charge of goethite surface

414 

decreases significantly with the attachment of phosphate (H2PO4− and HPO42−). The

415 

more phosphate adsorption, the less aspartate outer-sphere complexation.

416 

In contrast to the partial removal of adsorbed aspartate in the initial phase, phosphate

417 

experiences fast and considerable adsorption within 30 min (SI Figure S17). Because 22   

ACS Paragon Plus Environment

Page 22 of 32

Page 23 of 32

Environmental Science & Technology

418 

of the much stronger adsorption affinity, the interfacial phosphate on goethite is hardly

419 

affected by the presence of aspartate (Figure 6). Moreover, no significant variations are

420 

observed from the contributions of each complex to the overall phosphate adsorption

421 

when competing with aspartate. The peak-area ratios of the ~1008, 1073, and 1046 cm–

422 

1

peaks herein are 5.41:1:1.11 and 4.18:1:4.39 at pH 3 and 6, respectively.

423 

7. Environmental Significance. The molecular-level adsorption mechanisms of

424 

aspartate and phosphate on goethite were investigated both in their individual and

425 

competition systems. As for aspartate only adsorption, we resolve one inner-sphere and

426 

two outer-sphere complexes that differ from those on Al2O3 and TiO2, which further

427 

demonstrates the critical influence of mineral phase on aspartate adsorption.3, 5 The B-

428 

B III inner-sphere coordination and electrostatic attraction result in favorable

429 

adsorption under acidic and near-neutral conditions. The markedly decreased

430 

adsorption at pH 11 is mainly due to the weakened H-bonding (by electrostatic

431 

repulsion) and the small number of available adsorption sites. The adsorption of amino

432 

acids on minerals can reduce their biological degradation rates in the environment,

433 

which has important impact on the geochemical cycle of nitrogen.46 Moreover, the

434 

adsorption modes significantly affect the stability of interfacial aspartate. In inner-

435 

sphere and electrostatic outer-sphere coordination, the aspartate NH3+ remains

436 

relatively free of surface complexation, which leaves it potentially subject to

437 

degradation.47, 48

438 

The phosphate competition dramatically changes the adsorption behaviors of

439 

aspartate on goethite. The unfavorable aspartate adsorption and partial removal of the 23   

ACS Paragon Plus Environment

Environmental Science & Technology

440 

adsorbed complex are attributed to the much weaker adsorption affinity than that of

441 

phosphate. The abundant phosphate in natural waters can inhibit the aspartate

442 

adsorption and accelerate its migration in the environment. On the contrary, the

443 

phosphate adsorption is less affected by the presence of aspartate, including the

444 

interfacial configurations (diprotonated M-M, monoprotonated B-B, and nonprotonated

445 

B-B) and the relative contributions of each complex. The favorable adsorption

446 

promotes the removal of bioavailable phosphate from natural solutions, i.e. the soil

447 

solution, the sediment pore water, and the overlying water. These results have important

448 

implications for understanding the fate and transport of amino acids and phosphate.

449 

ASSOCIATED CONTENT

450 

Supporting Information

451 

Speciation distribution of dissolved aspartate and phosphate as a function of pH.

452 

Dynamic spectra of aspartate and phosphate on goethite in single-adsorbate systems.

453 

DFT calculations of dissolved aspartate in solution and interfacial configurations on

454 

goethite. 2D correlation analysis of phosphate only adsorption. ATR spectra of

455 

dissolved phosphate and DFT calculations of interfacial phosphate on goethite.

456 

Interfacial phosphate on goethite reported by previous studies. Dynamic spectra of

457 

competitive adsorption between aspartate and phosphate. This material is available free

458 

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

459 

AUTHOR INFORMATION

460 

Corresponding Author 24   

ACS Paragon Plus Environment

Page 24 of 32

Page 25 of 32

Environmental Science & Technology

461 

Shengrui Wang*

462 

Address: Chinese Research Academy of Environmental Sciences, 8 Dayangfang,

463 

Beiyuan, Beijing 100012, China

464 

Tel.: +86-10-84915277. Fax: +86-10-84915190. E-mail: [email protected].

465 

Notes

466 

The authors declare no competing financial interest.

467 

ACKNOWLEDGEMENTS

468 

This research was supported by the National Natural Science Foundation of China

469 

(U1202235), the National High-Level Talents Special Support Plan (for Science and

470 

Technology Innovation Talents to Special Support Plan), the National Key Science and

471 

Technology Special Program “Water Pollution Control and Treatment” (2012ZX07102-

472 

004), and the China Postdoctoral Science Foundation (2014M561024).

473 

REFERENCES

474 

(1) Lee, N.; Sverjensky, D. A.; Hazen, R. M. Cooperative and competitive adsorption

475 

of amino acids with Ca2+ on rutile (α-TiO2). Environ. Sci. Technol. 2014, 48 (16),

476 

9358−9365.

477 

(2) Sverjensky, D. A.; Jonsson, C. M.; Jonsson, C. L.; Cleaves, H. J.; Hazen, R. M.

478 

Glutamate surface speciation on amorphous titanium dioxide and hydrous ferric oxide.

479 

Environ. Sci. Technol. 2008, 42 (16), 6034−6039.

480 

(3) Greiner, E.; Kumar, K.; Sumit, M.; Giuffre, A.; Zhao, W.; Pedersen, J.; Sahai, N.

481 

Adsorption of L-glutamic acid and L-aspartic acid to γ-Al2O3. Geochim. Cosmochim. 25   

ACS Paragon Plus Environment

Environmental Science & Technology

482 

Acta 2014, 133, 142−155.

483 

(4) Chu, W.-H.; Gao, N.-Y.; Deng, Y.; Dong, B.-Z. Formation of chloroform during

484 

chlorination of alanine in drinking water. Chemosphere 2009, 77 (10), 1346−1351.

485 

(5) Parikh, S. J.; Kubicki, J. D.; Jonsson, C. M.; Jonsson, C. L.; Hazen, R. M.;

486 

Sverjensky, D. A.; Sparks, D. L. Evaluating glutamate and aspartate binding

487 

mechanisms to rutile (α-TiO2) via ATR-FTIR spectroscopy and quantum chemical

488 

calculations. Langmuir 2011, 27 (5), 1778−1787.

489 

(6) Roddick-Lanzilotta, A. D.; McQuillan, A. J. An in situ infrared spectroscopic study

490 

of glutamic acid and of aspartic acid adsorbed on TiO2: Implications for the

491 

biocompatibility of titanium. J. Colloid Interface Sci. 2000, 227 (1), 48−54.

492 

(7) Pászti, Z.; Guczi, L. Amino acid adsorption on hydrophilic TiO2: A sum frequency

493 

generation vibrational spectroscopy study. Vib. Spectrosc 2009, 50 (1), 48−56.

494 

(8) Giacomelli, C. E.; Avena, M. J.; De Pauli, C. P. Aspartic acid adsorption onto TiO2

495 

particles surface. Experimental data and model calculations. Langmuir 1995, 11 (9),

496 

3483−3490.

497 

(9) Guo, Y.-n.; Lu, X.; Zhang, H.-p.; Weng, J.; Watari, F.; Leng, Y. DFT Study of the

498 

adsorption of aspartic acid on pure, N-doped, and Ca-doped rutile (110) surfaces. J.

499 

Phys. Chem. C 2011, 115 (38), 18572−18581.

500 

(10) Jonsson, C. M.; Jonsson, C. L.; Estrada, C.; Sverjensky, D. A.; Cleaves Ii, H. J.;

501 

Hazen, R. M. Adsorption of L-aspartate to rutile (α-TiO2): Experimental and theoretical

502 

surface complexation studies. Geochim. Cosmochim. Acta 2010, 74 (8), 2356−2367.

503 

(11) Ikhsan, J.; Johnson, B. B.; Wells, J. D.; Angove, M. J. Adsorption of aspartic acid 26   

ACS Paragon Plus Environment

Page 26 of 32

Page 27 of 32

Environmental Science & Technology

504 

on kaolinite. J. Colloid Interface Sci. 2004, 273 (1), 1−5.

505 

(12) Naidja, A.; Huang, P. M. Aspartic acid interaction with Ca-montmorillonite:

506 

adsorption, desorption and thermal stability. Appl. Clay Sci. 1994, 9 (4), 265−281.

507 

(13) Hanna, K.; Martin, S.; Quilès, F.; Boily, J. F. Sorption of phthalic acid at goethite

508 

surfaces under flow-through conditions. Langmuir 2014, 30 (23), 6800−6807.

509 

(14) Lindegren, M.; Loring, J. S.; Persson, P. Molecular structures of citrate and

510 

tricarballylate adsorbed on α-FeOOH particles in aqueous suspensions. Langmuir 2009,

511 

25 (18), 10639−10647.

512 

(15) Hongshao, Z.; Stanforth, R. Competitive adsorption of phosphate and arsenate on

513 

goethite. Environ. Sci. Technol. 2001, 35 (24), 4753−4757.

514 

(16) Villalobos, M.; Trotz, M. A.; Leckie, J. O. Surface complexation modeling of

515 

carbonate effects on the adsorption of Cr(VI), Pb(II), and U(VI) on goethite. Environ.

516 

Sci. Technol. 2001, 35 (19), 3849−3856.

517 

(17) Norén, K.; Loring, J. S.; Persson, P. Adsorption of alpha amino acids at the

518 

water/goethite interface. J. Colloid Interface Sci. 2008, 319 (2), 416−428.

519 

(18) Yang, Y.; Duan, J.; Jing, C. Molecular-scale study of salicylate adsorption and

520 

competition with catechol at goethite/aqueous solution interface. J. Phys. Chem. C 2013,

521 

117 (20), 10597−10606.

522 

(19) Estrada, C. F.; Sverjensky, D. A.; Pelletier, M.; Razafitianamaharavo, A.; Hazen,

523 

R. M. Interaction between L-aspartate and the brucite [Mg(OH)2]–water interface.

524 

Geochim. Cosmochim. Acta 2015, 155, 172−186.

525 

(20) Tejedor-Tejedor, M. I.; Anderson, M. A. The protonation of phosphate on the 27   

ACS Paragon Plus Environment

Environmental Science & Technology

526 

surface of goethite as studied by CIR-FTIR and electrophoretic mobility. Langmuir

527 

1990, 6 (3), 602−611.

528 

(21) Luengo, C.; Brigante, M.; Antelo, J.; Avena, M. Kinetics of phosphate adsorption

529 

on goethite: Comparing batch adsorption and ATR-IR measurements. J. Colloid

530 

Interface Sci. 2006, 300 (2), 511−518.

531 

(22) Rahnemaie, R.; Hiemstra, T.; van Riemsdijk, W. H. Geometry, charge distribution,

532 

and surface speciation of phosphate on goethite. Langmuir 2007, 23 (7), 3680−3689.

533 

(23) Kubicki, J. D.; Paul, K. W.; Kabalan, L.; Zhu, Q.; Mrozik, M. K.; Aryanpour, M.;

534 

Pierre-Louis, A.-M.; Strongin, D. R. ATR-FTIR and density functional theory study of

535 

the structures, energetics, and vibrational spectra of phosphate adsorbed onto goethite.

536 

Langmuir 2012, 28 (41), 14573−14587.

537 

(24) Zeng, H.; Fisher, B.; Giammar, D. E. Individual and competitive adsorption of

538 

arsenate and phosphate to a high-surface-area iron oxide-based sorbent. Environ. Sci.

539 

Technol. 2008, 42 (1), 147−152.

540 

(25) Cheng, T.; Barnett, M. O.; Roden, E. E.; Zhuang, J. Effects of phosphate on

541 

uranium(VI) adsorption to goethite-coated sand. Environ. Sci. Technol. 2004, 38 (22),

542 

6059−6065.

543 

(26) Wang, X.-H.; Liu, F.-F.; Lu, L.; Yang, S.; Zhao, Y.; Sun, L.-B.; Wang, S.-G.

544 

Individual and competitive adsorption of Cr(VI) and phosphate onto synthetic Fe–Al

545 

hydroxides. Colloids Surf., A 2013, 423, 42−49.

546 

(27) Hiemstra, T.; Mia, S.; Duhaut, P.-B.; Molleman, B. Natural and pyrogenic humic

547 

acids at goethite and natural oxide surfaces interacting with phosphate. Environ. Sci. 28   

ACS Paragon Plus Environment

Page 28 of 32

Page 29 of 32

Environmental Science & Technology

548 

Technol. 2013, 47 (16), 9182−9189.

549 

(28) Qin, X.; Liu, F.; Wang, G.; Li, L.; Wang, Y.; Weng, L. Modeling of levofloxacin

550 

adsorption to goethite and the competition with phosphate. Chemosphere 2014, 111,

551 

283−290.

552 

(29) Yang, Y.; Yan, W.; Jing, C. Dynamic adsorption of catechol at the goethite/aqueous

553 

solution interface: A molecular-scale study. Langmuir 2012, 28 (41), 14588−14597.

554 

(30) Ding, X.; Song, X.; Boily, J.-F. Identification of fluoride and phosphate binding

555 

sites at FeOOH surfaces. J. Phys. Chem. C 2012, 116 (41), 21939−21947.

556 

(31) Ha, J.; Hyun Yoon, T.; Wang, Y.; Musgrave, C. B.; Brown, J. G. E. Adsorption of

557 

organic matter at mineral/water interfaces: 7. ATR-FTIR and quantum chemical study

558 

of lactate interactions with hematite nanoparticles. Langmuir 2008, 24 (13), 6683−6692.

559 

(32) Foerstendorf, H.; Jordan, N.; Heim, K. Probing the surface speciation of uranium

560 

(VI) on iron (hydr)oxides by in situ ATR FT-IR spectroscopy. J. Colloid Interface Sci.

561 

2014, 416, 133−138.

562 

(33) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.;

563 

Cheeseman, J. R.; Montgomery, Jr., J. A.; Vreven, T.; Kudin, K. N.; Burant, J. C.;

564 

Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani,

565 

G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda,

566 

R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.;

567 

Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.;

568 

Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.;

569 

Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. 29   

ACS Paragon Plus Environment

Environmental Science & Technology

570 

J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick,

571 

D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul,

572 

A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.;

573 

Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.;

574 

Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M.

575 

W.; Gonzalez, C.; and Pople, J. A. Gaussian 03, Revision E.01, Gaussian, Inc.,

576 

Wallingford CT, 2004..

577 

(34) Bhandari, N.; Hausner, D. B.; Kubicki, J. D.; Strongin, D. R. Photodissolution of

578 

ferrihydrite in the presence of oxalic acid: An in situ ATR-FTIR/DFT study. Langmuir

579 

2010, 26 (21), 16246−16253.

580 

(35) Johnston, C. P.; Chrysochoou, M. Investigation of chromate coordination on

581 

ferrihydrite by in situ ATR-FTIR spectroscopy and theoretical frequency calculations.

582 

Environ. Sci. Technol. 2012, 46 (11), 5851−5858.

583 

(36) Li, W.; Pierre-Louis, A.-M.; Kwon, K. D.; Kubicki, J. D.; Strongin, D. R.; Phillips,

584 

B. L. Molecular level investigations of phosphate sorption on corundum (α-Al2O3) by

585 

31

586 

Cosmochim. Acta 2013, 107, 252−266.

587 

(37) Pierre-Louis, A.-M.; Hausner, D. B.; Bhandari, N.; Li, W.; Kim, J.; Kubicki, J. D.;

588 

Strongin, D. Adsorption of carbon dioxide on Al/Fe oxyhydroxide. J. Colloid Interface

589 

Sci. 2013, 400, 1−10.

590 

(38) Sanchez-de-Armas, R.; San-Miguel, M. A.; Oviedo, J.; Marquez, A.; Sanz, J. F.

591 

Electronic structure and optical spectra of catechol on TiO2 nanoparticles from real time

P solid state NMR, ATR-FTIR and quantum chemical calculation. Geochim.

30   

ACS Paragon Plus Environment

Page 30 of 32

Page 31 of 32

Environmental Science & Technology

592 

TD-DFT simulations. Phys. Chem. Chem. Phys. 2011, 13 (4), 1506−1514.

593 

(39) NIST Computational Chemistry Comparison and Benchmark Database; NIST

594 

Standard Reference Database Number 101; Johnson, R. D., III, Ed.; Release 16a,

595 

August 2013, http://cccbdb.nist.gov/.

596 

(40) Norén, K.; Persson, P. Adsorption of monocarboxylates at the water/goethite

597 

interface: The importance of hydrogen bonding. Geochim. Cosmochim. Acta 2007, 71

598 

(23), 5717−5730.

599 

(41) Noda, I.; Ozaki, Y. Two-Dimensional Correlation Spectroscopy: Applications in

600 

Vibrational and Optical Spectroscopy; John Wiley & Sons: Chichester, England, 2004.

601 

(42) Pearson, J. F.; Slifkin, M. A. The infrared spectra of amino acids and dipeptides.

602 

Spectrochim. Acta 1972, 28A (12), 2403−2417.

603 

(43) Johnson, S. B.; Yoon, T. H.; Kocar, B. D.; Brown, G. E. Adsorption of organic

604 

matter at mineral/water interfaces. 2. Outer-sphere adsorption of maleate and

605 

implications for dissolution processes. Langmuir 2004, 20 (12), 4996−5006.

606 

(44) Hwang, Y. S.; Liu, J.; Lenhart, J. J.; Hadad, C. M. Surface complexes of phthalic

607 

acid at the hematite/water interface. J. Colloid Interface Sci. 2007, 307 (1), 124−134.

608 

(45) Gulley-Stahl, H.; Hogan, P. A.; Schmidt, W. L.; Wall, S. J.; Buhrlage, A.; Bullen,

609 

H. A. Surface complexation of catechol to metal oxides: An ATR-FTIR, adsorption, and

610 

dissolution study. Environ. Sci. Technol. 2010, 44 (11), 4116−4121.

611 

(46) Kitadai, N.; Yokoyama, T.; Nakashima, S. In situ ATR-IR investigation of L-lysine

612 

adsorption on montmorillonite. J. Colloid Interface Sci. 2009, 338 (2), 395−401.

613 

(47) Sheals, J.; Sjöberg, S.; Persson, P. Adsorption of glyphosate on goethite:  31   

ACS Paragon Plus Environment

Environmental Science & Technology

614 

Molecular characterization of surface complexes. Environ. Sci. Technol. 2002, 36 (14),

615 

3090−3095.

616 

(48) Jonsson, C. M.; Persson, P.; Sjöberg, S.; Loring, J. S. Adsorption of glyphosate on

617 

goethite (α-FeOOH): Surface complexation modeling combining spectroscopic and

618 

adsorption data. Environ. Sci. Technol. 2008, 42 (7), 2464−2469. 

32   

ACS Paragon Plus Environment

Page 32 of 32