Molecular Insights into Glyphosate Adsorption to ... - ACS Publications

ACS Paragon Plus Environment. Environmental ..... phosphonate group rather than the deprotonated amine group (slope: 0.99 vs 0.96, SD: 30 vs 59,. 189 ...
1 downloads 18 Views 1MB Size
Subscriber access provided by MT ROYAL COLLEGE

Article

Molecular Insights into Glyphosate Adsorption to Goethite Gained from ATR-FTIR, Two-Dimensional Correlation Spectroscopy, and DFT Study Wei Yan, and Chuanyong Jing Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b05643 • Publication Date (Web): 22 Jan 2018 Downloaded from http://pubs.acs.org on January 22, 2018

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

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

Page 1 of 28

Environmental Science & Technology

Molecular Insights into Glyphosate Adsorption to Goethite Gained from ATR-FTIR, Two-Dimensional Correlation Spectroscopy, and DFT Study

Wei Yana and Chuanyong Jinga,b*

a

State Key Laboratory of Environmental Chemistry and Ecotoxicology, Research

Center for Eco-Environmental Sciences, Chinese Academy of Sciences, Beijing 100085, China

b

University of Chinese Academy of Sciences, Beijing 100049, China

*Corresponding author: Dr. Chuanyong Jing; Tel: +86 10 6284 9523; Fax: +86 10 6284 9523; E-mail: [email protected]

1

ACS Paragon Plus Environment

Environmental Science & Technology

1

Abstract

2

Glyphosate (PMG) complexation on iron (hydr)oxides impacts its fate and transport in the

3

environment. To decipher the molecular-level interfacial configuration and reaction mechanism

4

of PMG on iron (hydr)oxides, the PMG protonation process, which influences the chemical and

5

physical properties of PMG, was first determined using ATR-FTIR spectroscopy. The FTIR

6

results reveal that the deprotonation occurs at carboxylate oxygen when pKa1< pH < pKa2, at

7

phosphonate oxygen when pKa2< pH < pKa3, and at amino nitrogen when pH > pKa3. PMG

8

complexation on goethite was investigated using in situ flow-cell ATR-FTIR, two-dimensional

9

correlation spectroscopy (2D-COS), and density functional theory (DFT) calculations. The

10

results indicate that the phosphonate group on PMG interacts with goethite to form inner-sphere

11

complexes with multiple configurations depending on pH: binuclear bidentate (BB) and

12

mononuclear bidentate (MB) without proton under acidic conditions (pH 5), mononuclear

13

monodentate (MM) with proton and BB without proton at pH 6-8, and MM without proton under

14

alkaline conditions (pH 9). Phosphate competition significantly impacted the PMG adsorption

15

capacity and its interfacial configurations. As a result, the stability of the adsorbed PMG was

16

impaired, as evidenced by its elevated leachability. These results improve our understanding of

17

PMG-mineral interactions at the molecular level and have significant implications for risk

18

assessment for PMG and structural analog pollutants.

19 20 21

2

ACS Paragon Plus Environment

Page 2 of 28

Page 3 of 28

Environmental Science & Technology

22 23 24

Table of Contents (TOC)/Abstract Art

25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 3

ACS Paragon Plus Environment

Environmental Science & Technology

43

Introduction

44

Glyphosate (N-(phosphonomethyl)glycine, PMG) is a nonselective, post-emergence herbicide for

45

weed control.1 Due to the worldwide application of PMG, its degradation, bioavailability, and

46

transport in the environment are of great concern, and these environmental processes are mainly

47

regulated by its chemistry at mineral/aqueous interfaces.2-5

48

PMG interacts with mineral surfaces through its three polar functional groups, namely,

49

phosphonate, amine, and carboxylate (Figure S1, in the Supporting Information (SI)). These

50

functional groups are readily deprotonated and/or dissociated depending on pH. The dissociation

51

constants have been accurately determined, but the deprotonation sequence is still a subject of

52

controversy. Some earlier studies reported that the deprotonation of PMG occurred at the

53

carboxyl group when pKa1< pH < pKa2, at the phosphonate group when pKa2< pH < pKa3, and

54

at the amine group when pH > pKa3 (Figure S2, a).6-8 Recent DFT9 and NMR10 studies suggested

55

a different deprotonation sequence which is carboxyl, amine, and phosphonate group as pH

56

increase (Figure S2, b).

57

The disagreement in PMG deprotonation seriously hinders the molecular-level understanding

58

of PMG adsorption on minerals including goethite,11-13 and divergent phosphonate group

59

involved mechanisms have been proposed on PMG surface complexation. For example, Sheals et

60

al6 speculated that PMG was adsorbed on goethite surfaces via a primary monodentate complex

61

and a secondary bidentate complex at neutral pH. Tribe et al13 found that PMG only formed

62

monodentate surface complexes in all the pH ranges. In contrast, Barja et al14 identified two

63

predominant surface species co-existing on goethite, namely, monodentate protonated and

64

bridging bidentate complexes. Recent studies suggest that adsorbed ions on goethite may exhibit 4

ACS Paragon Plus Environment

Page 4 of 28

Page 5 of 28

Environmental Science & Technology

65

a variety of surface configurations depending on the exposed crystal faces, adding complexity to

66

the structural identification.15-17

67

The ambiguities in deprotonation sequence and surface molecular structure may largely be

68

attributed to the lack of effective approach for the structural determination. As a widely-used

69

technique to probe the solid-liquid interface reactions, attenuated total reflectance (ATR) Fourier

70

transform Infrared (FTIR) spectroscopy often results in equivocal peak assignment due to its

71

intrinsic low resolution to resolve convoluted peaks,18 which may impair the accurate

72

identification of surface complexes. Thus, complementary techniques such as two-dimensional

73

correlation spectroscopy (2D-COS) analysis, and DFT calculations are central to re-examine the

74

PMG deprotonation process and the surface complexation.

75

The application of PMG is usually accompanied by phosphorus fertilization. The abundant

76

phosphate in the soil can compete with PMG for the available surface sites on minerals, leading

77

to the leaching and runoff of PMG to deeper layers of the soil.19 Nevertheless, the

78

molecular-level impacts of phosphate on the PMG interaction with surfaces and the

79

corresponding structures of PMG surface complexes are far from well understood.

80

The objective of this study was therefore to explore the molecular-scale interaction of PMG on

81

Fe(III) (hydr)oxide surfaces using multiple complementary techniques, including in situ

82

flow-cell ATR-FTIR spectroscopy (in H2O and D2O), 2D-COS analysis, and DFT calculations.

83

The insights gained from this study improve our knowledge in predicting the environmental fate

84

of PMG and structural analog pollutants.

85

Material and Methods

5

ACS Paragon Plus Environment

Environmental Science & Technology

Page 6 of 28

86

Materials. All chemicals were analytical reagent grade or higher and were used without further

87

purification. Milli-Q water (18.2 MΩ) for ATR-FTIR analysis was boiled for 60 min and cooled

88

with N2 gas purging to remove CO2.

89

Goethite was selected as a typical iron (hydr)oxide, considering its ubiquity in the

90

environment and highly reactive surfaces. Goethite was synthesized via iron nitrate hydrolysis as

91

detailed in our previous report (Figure S3).20,

92

determined with an ASAP2000 instrument (Micromeritics Instrument Corp., USA). The point of

93

zero charge (PZC) of goethite was determined to be 8.9 with a Zetasizer Nano ZS (Malvern

94

Instruments, U.K.).

95

ATR-FTIR

Measurements.

ATR-FTIR

21

The BET surface area (84.7 m2/g) was

measurements

were

performed

using

a

96

Thermo-Nicolet Nexus 6700 FTIR spectrometer equipped with a liquid-nitrogen cooled MCT

97

detector. Both H2O and D2O (99.9 atom % D, Sigma) were used as solvent in the FTIR

98

experiments. The IR spectra were collected using a horizontal attenuated total reflectance (HATR)

99

cell (PIKE Tech) with a 45°ZnSe (pH/pD 5-9) or Ge (beyond pH/pD range of 5-9) ATR crystal

100

in 0.01-0.2 M NaCl (Figure S4). A total of 1000 and 256 scans were recorded for static and in

101

situ flow cell ATR-FTIR spectrum, respectively, over 1800-800 cm−1 at a resolution of 4 cm−1.22

102

The details of static and in situ flow cell ATR-FTIR measurements are provided in the SI.

103

For better identification of peaks, the peak fittings of all IR spectra were carried out with

104

Peakfit v.4.12 software using the second derivative fitting algorithm. The goodness of fit R factor

105

in all of the fitted spectra was greater than 0.999. The details are provided in the Figure S5.

6

ACS Paragon Plus Environment

Page 7 of 28

Environmental Science & Technology

106

Frequency Calculations. The IR frequency of the PMG molecule and PMG surface

107

complexes with different configurations were calculated using the Gaussian 09 program with the

108

B3LYP hybrid DFT method.23 The cluster models of goethite, PMG, and PMG surface

109

complexes were constructed and the details are provided in the SI. The cluster model results

110

could only be used as a qualitative guide in IR peak analysis rather than as an absolute energy

111

value. Because PMG surface complexes are not dispersion-dominated and no interaction

112

energies are involved, the dispersion corrections to standard DFT functions are not applied.

113

2D-COS Analysis. In order to determine the origin of the IR bands in the obtained IR spectra,

114

2D-COS analysis was performed using 2Dshige software (Shigeaki Morita, Japan). Compared to

115

conventional FTIR analysis, 2D-COS has the advantage of being able to distinguish one band

116

from another by its origins, which has been successfully applied to probe complicated surface

117

interaction processes in different systems.20-22 A detailed description of 2D-COS analysis is

118

provided in the SI.

119

Batch Competitive Adsorption Isotherms. To explore the impact of phosphate on PMG

120

adsorption and the corresponding structures of PMG complexes, competitive adsorption

121

experiments of PMG and phosphate on goethite surfaces were conducted at pH 7. The details of

122

batch adsorption experiments and the sample analysis, including HPLC-ESI-MS/MS and

123

colorimetric assay of PMG and phosphate, are provided in the SI.

124

Results and Discussion

125

PMG deprotonation process. Because of its central importance, the PMG deprotonation

126

process was determined first by analyzing the IR spectra for various dissolved PMG species

127

(Figure 1a). According to the dissociation constants (pKa1=2.3, pKa2=6.0, pKa3=11.0),7 the 7

ACS Paragon Plus Environment

Environmental Science & Technology

128

predominant species of PMG at pH 1.5, 4.2, 9.0, and 12.5 are neutral (PMG), monoanion

129

(PMG−), dianion (PMG2−), and trianion (PMG3−), respectively (Figure S1). It worth note that the

130

phosphonate group can protonate to form a positively charged species at pH 0.99) between the results of

187

DFT calculations and deconvoluted ATR-FTIR spectra (Figure S6, S8, S9, Table S3-S6).

188

Specifically, the DFT results for the PMG2− species are aligned with the deprotonated

189

phosphonate group rather than the deprotonated amine group (slope: 0.99 vs 0.96, SD: 30 vs 59,

10

ACS Paragon Plus Environment

Page 10 of 28

Page 11 of 28

Environmental Science & Technology

190

Figure S10), supporting our FTIR results that the second deprotonation of PMG occurs at the

191

phosphonate oxygen rather than the amino nitrogen.

192

Surface Complexation of PMG on Goethite. The protonation process of PMG has a

193

substantial impact on its solid-water interfacial interactions in the environment. Since PMG and

194

PMG3− exist, respectively, under extremely acidic and alkaline conditions, only environmentally

195

relevant PMG species, i.e. PMG− and PMG2− at pH 5-9, were investigated regarding their surface

196

complexation.

197

Though PMG is a tridentate chelating ligand, mounting evidence suggests that the

198

carboxylate and amine groups do not directly interact with goethite.6, 13, 14 As shown in Figures 2

199

and S6, adsorption did not shift the peak of υas(COO−) at 1400 cm−1, indicating either no or weak

200

outer-sphere carboxylate surface interactions.6 In addition, the XPS analysis by Sheals et al.

201

showed no evidence of direct interaction between the amine group and goethite.6 Conversely, the

202

IR vibrations of the phosphonate group (1200-900 cm−1) exhibited dramatic changes upon

203

adsorption at different pH values (Figure 2), indicating the formation of inner-sphere surface

204

complexes on goethite via the phosphonate group.

205

For an in-depth understanding of the PMG-goethite interaction mechanism, DFT

206

calculations were performed to provide the theoretical IR bands of various PMG surface

207

complexes. The possible PMG surface configurations included mononuclear monodentate (MM),

208

mononuclear bidentate (MB), and binuclear bidentate (BB) structures. In addition, each structure

209

considered two protonation states of phosphonate moiety: protonated (Type I) and deprotonated

210

(Type II) species, leading to an overall of six PMG complexes (Figure 3).

11

ACS Paragon Plus Environment

Environmental Science & Technology

211

The comparison of DFT-calculated and experimental IR spectra showed that none of the single

212

configuration match the experimental IR data. The disagreement can be attributed to two facts: 1)

213

more peaks appear in the calculated spectra than in the experimental spectra due to the

214

asymmetry of cluster models, and 2) the DFT-calculated spectrum only corresponds to one

215

specific surface configuration, whereas more than one PMG surface configurations may coexist

216

at certain pH.

217

To solve these problems, we first categorized the observed IR peaks into different groups by

218

2D-COS analysis due to its merits in distinguishing one band from another by its origins upon

219

external perturbations.20 The 2D-COS analysis results in Figure 4 contoured the synchronous and

220

asynchronous correlation maps for PMG-goethite complexes, which were obtained from the

221

dynamic IR spectra of PMG on goethite at pH 5, 7, and 9 (Figure S11). According to the detailed

222

discussion in the SI, the IR peaks can be classified into the following five groups at pH 5-9: a)

223

1142, 1018, 987 cm−1 and b) 1122, 1052, 956 cm−1 at pH 5; c) 1025, 987, 938 cm−1 and d) 1128,

224

1094, 1060 cm−1 at pH 7; while only e) 1102, 1004, and 980 cm−1 at pH 9. Compared to the

225

DFT-calculated spectra, group a) and c) belong to BB II; group b), d), and e) belong to MB II,

226

MM I, and MM II, respectively (Figure 3, Figure S12). This 2D-COS result highlights the

227

diversity of PMG configurations on goethite surfaces.

228

To confirm the existence of multiple PMG surface complexes, ATR-FTIR spectra in D2O were

229

collected and compared with those in H2O. As shown in Figure 5, most IR peaks at pD 5 and 9 in

230

D2O had the same positions to those in H2O. This observation implied that the phosphonate

231

group of the PMG surface complexes at pH 5 and 9 were without proton.28 At pH/pD=7,

232

however, the group d as classified by the above 2D-COS analysis, namely, 1128, 1094, and 1060 12

ACS Paragon Plus Environment

Page 12 of 28

Page 13 of 28

Environmental Science & Technology

233

cm−1, had a significant shift to lower frequencies, whereas the peaks in the group c (1025, 987,

234

and 938 cm−1) remained approximately the same positions upon the deuterium exchange. This

235

result strongly suggested that two PMG surface complexes coexisted at pH 7. Integrating IR

236

spectra, 2D-COS analysis, and DFT results concluded that BB II and MB II complexes were

237

predominant under a slightly acidic condition (pH 5), MM I and BB II under neutral condition

238

(pH 6–8), while only MM II under alkaline condition (pH 9). The comparison of the calculated

239

and experimental IR vibrational frequencies and the assignment of possible PMG complexes was

240

summarized in Table 1.

241

The pH dependence of PMG surface complexes can be attributed to the ligand-exchange

242

mechanism. Under acidic conditions, the ligand, phosphonate group on PMG, is concentrated to

243

the goethite surface (pHpzc=8.9) by the electrostatic attraction, which increases the energetic

244

favorability to replace the surface hydroxyl group to form bidentate structures, including BB and

245

MB configurations. As the pH increases to neutral, the competition of OH− hinders the ligand

246

exchange reaction of PMG with surfaces, resulting in a successive disappearance of the MB

247

configuration. MB is a four-membered ring structure which has a relatively unfavorable energy

248

compared with the BB configuration.14 Under alkaline conditions, the MM complex without

249

proton becomes the predominant configuration due to the strong competition with OH−.

250

Competition of phosphate with PMG for Surface Complexation. Because PMG forms

251

surface complexes through the phosphonate group, the competitive adsorption of PMG and

252

phosphate should be of great environmental significance considering the co-application of

253

phosphate fertilizers and the herbicide PMG.19

13

ACS Paragon Plus Environment

Environmental Science & Technology

254

Both PMG and phosphate adsorption conformed to the Langmuir isotherm, and the

255

adsorption capacity of PMG (42.3 mg/g or 0.25 mM/g, Figure 6a) was much lower than that of

256

phosphate (159.3 mg/g or 1.021 mM/g, Figure 6b). In the competitive system, phosphate

257

dramatically reduced PMG adsorption by up to about 50% when the molar ratio of phosphate to

258

PMG increased to 5:1 (Figure 6a). In contrast, PMG had negligible effect on phosphate

259

adsorption, even when the molar ratio was as high as 10:1 (Figure 6b).

260

The difference in the macroscopic competitive adsorption behaviors of PMG and phosphate

261

can be explained by the FTIR results (Figure 6c). The spectrum of the PMG and phosphate

262

coexisting system (molar ratio = 1:1, bottom of Figure 6c) was almost identical to that of

263

phosphate adsorption alone (middle of Figure 6c), both of which were in stark contrast to that of

264

PMG adsorption alone (top of Figure 6c). This observation indicates that phosphate has a

265

stronger affinity for goethite surfaces. The unchanged profile and peak positions at 1590, 1400,

266

and 1324 cm−1 in the spectrum of PMG and phosphate coadsorption (Figure S13) indicate that

267

the carboxyl group of PMG may not directly interact with goethite.

268

Further quantitative analysis on the area of deconvoluted interfacial peaks can provide insights

269

into the structural information. Compared to the phosphate-only system (middle of Figure 6c),

270

the relative peak areas of several peaks increased in the competitive system, especially the peak

271

at 1124 cm−1 (28% increase relative to 1050 cm−1, bottom of Figure 6c), the position of which is

272

close to the strongest peak (1128 cm−1) in the PMG-only system (top of Figure 6c). This result

273

indicated that a small amount of PMG complexes, though less than phosphate, contributed to the

274

inner-sphere complexation in the competitive system.

14

ACS Paragon Plus Environment

Page 14 of 28

Page 15 of 28

Environmental Science & Technology

275

Interestingly, phosphate preferentially occupied the goethite sites that originally were prone to

276

form the BB II configuration in the PMG-only system. For example, for the peak at 980 and 942

277

cm−1 in the competitive system, corresponding to 987 and 938 cm−1 of BB II in the PMG-only

278

system (top of Figure 6c), the relative peak areas only increased 8% and 1%, respectively

279

(bottom of Figure 6c). By comparison, for the peaks at 1124 and 1088 cm−1 in the competitive

280

system, which are close to the peaks at 1128 and 1094 cm−1 of MM I in the PMG-only system,

281

the relative peak areas appreciably increased by 28% and 21%, respectively. The uneven increase

282

in peak areas showed that phosphate not only displayed a stronger affinity than PMG for goethite

283

surfaces, but also preferentially occupied the limited surface sites which originally formed the

284

stable PMG surface complex, namely, BB II. As a result of phosphate competition, a great

285

proportion of PMG molecules were compelled to form relatively unfavorable MM I.

286

Notably, phosphate surface complexes may also consist of multiple species.17,

29

Because

287

phosphate dramatically reduced PMG adsorption whereas PMG had a negligible effect on

288

phosphate adsorption (Figure 6a, b), it is unlikely that PMG has a great impact on phosphate

289

surface structure. Therefore, the configurations of phosphate complexes were relatively stable in

290

the presence of PMG and the uneven increase in peak areas should be mainly attributed to the

291

change in the relative proportion of PMG surface species.

292

The impact of surface structures on the mobility of adsorbed PMG and phosphate motivated

293

our further study using extraction experiments. The release of PMG and phosphate followed the

294

pseudo-second-order kinetics model (Figure S14, Table S8-S9). The leaching results show that

295

19.1-29.5% of PMG was released in the ternary PMG-phosphate-goethite system, appreciably

296

higher than that in the binary PMG-goethite system (9.6-18.6%, Table S8). This result clearly 15

ACS Paragon Plus Environment

Environmental Science & Technology

297

demonstrates that phosphate competition altered the PMG surface complex to a less stable

298

configuration. Conversely, PMG resulted in an insignificant impact on phosphate desorption

299

(3.8-4.0% in ternary system vs 6.1-6.3% in binary system, Table S9),in agreement with their

300

adsorption behaviors (Figure 6).

301

Environmental Significance. PMG has become an agricultural panacea due to its high

302

efficiency and broad-spectrum properties, while its environmental behaviors is still unclear. Our

303

results here provide new insights into the molecular-level surface complexation of glyphosate on

304

goethite. The phosphonate group plays a key role in the formation of inner-sphere complexes on

305

goethite surfaces, while the carboxyl group may only take part in outer-sphere surface

306

interactions via electrostatic attraction. Notably, multiple PMG surface configurations may

307

coexist and transform depending on pH. Considering that the BB structure is more stable than

308

MB and MM, we should pay particular attention to the pH factor in evaluating the behavior,

309

bioavailability, and fate of PMG in the environment.

310

Our work also demonstrates that the presence of phosphate inhibits PMG adsorption by

311

competing for surface sites. Moreover, phosphate competition impairs the affinity of PMG

312

toward goethite surfaces by restricting PMG to forming an energetically unfavorable MM rather

313

than a more stable BB structure, resulting in higher mobility for adsorbed PMG. Given the

314

widespread phosphorus fertilization and its strong affinity to soil components, phosphate, the

315

structural analog to the active moiety of PMG, should be carefully considered in future studies

316

on PMG risk assessment.

317

ASSOCIATED CONTENT

16

ACS Paragon Plus Environment

Page 16 of 28

Page 17 of 28

Environmental Science & Technology

318

Supporting Information. The Supporting Information is available free of charge on the ACS

319

Publications website at http://pubs.acs.org.

320

Details of in situ flow cell ATR-FTIR measurements; Peak fitting procedure; 2D-COS analysis;

321

PMG and phosphate Determination; analytical procedure; dynamic spectra of PMG on goethite

322

at pH 5, 7, and 9; and additional figures.

323 324

Notes The authors declare no competing financial interest. ACKNOWLEDGEMENTS

325

We acknowledge the financial support of the National Key Basic Research Program of China

326

(2015CB932003, 2014CB441102), the Strategic Priority Research Program of the Chinese

327

Academy of Sciences (XDB14020302), the National Natural Science Foundation of China

328

(21477144, 41425016, and 21321004).

329 330

References

331 332 333 334 335 336 337 338 339 340 341 342 343 344 345

(1) Sammons, R. D.; Gaines, T. A. Glyphosate resistance: State of knowledge. Pest Manag. Sci. 2014, 70, 1367-1377. (2) Doublet, J.; Mamy, L.; Barriuso, E. Delayed degradation in soil of foliar herbicides glyphosate and sulcotrione previously absorbed by plants: Consequences on herbicide fate and risk assessment. Chemosphere 2009, 77, 582-589. (3) Zablotowicz, R. M.; Accinelli, C.; Krutz, L. J.; Reddy, K. N. Soil depth and tillage effects on glyphosate degradation. J. Agric. Food Chem. 2009, 57, 4867-4871. (4) Dousset, S.; Jacobson, A. R.; Dessogne, J.-B.; Guichard, N.; Baveye, P. C.; Andreux, F. Facilitated transport of diuron and glyphosate in high copper vineyard soils. Environ. Sci. Technol. 2007, 41, 8056-8061. (5) Jonsson, C. M.; Persson, P.; Sjöberg, S.; Loring, J. S. Adsorption of glyphosate on goethite (α-FeOOH): Surface complexation modeling combining spectroscopic and adsorption data. Environ. Sci. Technol. 2008, 42, 2464-2469. (6) Sheals, J.; Sjoberg, S.; Persson, P. Adsorption of glyphosate on goethite: Molecular characterization of surface complexes. Environ. Sci. Technol. 2002, 36, 3090-3095.

17

ACS Paragon Plus Environment

Environmental Science & Technology

346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385

(7) Li, W.; Wang, Y.-J.; Zhu, M.; Fan, T.-T.; Zhou, D.-M.; Phillips, B. L.; Sparks, D. L. Inhibition mechanisms of Zn precipitation on aluminum oxide by glyphosate: A 31P NMR and Zn EXAFS study. Environ. Sci. Technol. 2013, 47, 4211-4219. (8) Harris, W. R.; Sammons, R. D.; Grabiak, R. C.; Mehrsheikh, A.; Bleeke, M. S. Computer simulation of the interactions of glyphosate with metal ions in phloem. J. Agric. Food Chem. 2012, 60, 6077-6087. (9) Peixoto, M. M.; Bauerfeldt, G. F.; Herbst, M. H.; Pereira, M. S.; da Silva, C. O. Study of the stepwise deprotonation reactions of glyphosate and the corresponding pKa values in aqueous solution. J. Phys. Chem. A 2015, 119, 5241-5249. (10) Liu, B.; Dong, L.; Yu, Q.; Li, X.; Wu, F.; Tan, Z.; Luo, S. Thermodynamic study on the protonation reactions of glyphosate in aqueous solution: potentiometry, calorimetry and NMR spectroscopy. J. Phys. Chem. B 2016, 120, 2132-2137. (11) Gimsing, A. L.; Borggaard, O. K.; Sestoft, P. Modeling the kinetics of the competitive adsorption and desorption of glyphosate and phosphate on goethite and gibbsite and in soils. Environ. Sci. Technol. 2004, 38, 1718-1722. (12) Mazzei, P.; Piccolo, A. Quantitative evaluation of noncovalent interactions between glyphosate and dissolved humic substances by NMR spectroscopy. Environ. Sci. Technol. 2012, 46, 5939-5946. (13) Tribe, L.; Kwon, K. D.; Trout, C. C.; Kubicki, J. D. Molecular orbital theory study on surface complex structures of glyphosate on goethite:  Calculation of vibrational frequencies. Environ. Sci. Technol. 2006, 40, 3836-3841. (14) Barja, B. C.; dos Santos Afonso, M. Aminomethylphosphonic acid and glyphosate adsorption onto goethite:  A comparative study. Environ. Sci. Technol. 2005, 39, 585-592. (15) Villalobos, M.; Perez-Gallegos, A. Goethite surface reactivity: A macroscopic investigation unifying proton, chromate, carbonate, and lead(II) adsorption. J. Colloid Interface Sci. 2008, 326, 307-323. (16) Villalobos, M.; Cheney, M.A.; Alcaraz-Cienfuegos, J. Goethite surface reactivity: II. A microscopic site-density model that describes its surface area-normalized variability. J. Colloid Interface Sci. 2009, 336, 412-422. (17) Kubicki, J.D.; Paul, K.W.; Kabalan, L.; Zhu, Q.; Mrozik, M.K.; Aryanpour M.; Pierre-Louis, A.M.; Strongin, D.R. ATR-FTIR and density functinal theory study of the structures, energetics, and vibrational spectra of phosphate adsorbed onto goethite. Langmuir 2012, 28, 14573-14587. (18) Barja, B. C.; dos Santos Afonso, M. An ATR−FTIR study of glyphosate and its Fe(III) complex in aqueous solution. Environ. Sci. Technol. 1998, 32, 3331-3335. (19) Sasal, M. C.; Demonte, L.; Cislaghi, A.; Gabioud, E. A.; Oszust, J. D.; Wilson, M. G.; Michlig, N.; Beldoménico, H. R.; Repetti, M. R. Glyphosate loss by runoff and its relationship with phosphorus fertilization. J. Agric. Food Chem. 2015, 63, 4444-4448. (20) Yang, Y.; Yan, W.; Jing, C. Dynamic adsorption of catechol at the goethite/aqueous solution interface: A molecular-scale study. Langmuir 2012, 28, 14588-14597. 18

ACS Paragon Plus Environment

Page 18 of 28

Page 19 of 28

386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418

Environmental Science & Technology

(21) Yan, W.; Wang, H.; Jing, C. Adhesion of Shewanella oneidensis MR-1 to goethite: A two-dimensional correlation spectroscopic study. Environ. Sci. Technol. 2016, 50, 4343-4349. (22) Yan, W.; Zhang, J.; Jing, C. Adsorption of enrofloxacin on montmorillonite: Two-dimensional correlation ATR/FTIR spectroscopy study. J. Colloid Interface Sci. 2013, 390, 196-203. (23) Frisch, M. J. T., G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; and Fox, D. J., Gaussian 09, revision A.01; Gaussian, Inc.: Wallingford, CT. 2009. (24) Sheals, J.; Persson, P.; Hedman, B. IR and EXAFS spectroscopic studies of glyphosate protonation and copper(II) complexes of glyphosate in aqueous solution. Inorg. Chem. 2001, 40, 4302-4309. (25) Barja, B. C.; Tejedor-Tejedor, M. I.; Anderson, M. A. Complexation of methylphosphonic acid with the surface of goethite particles in aqueous solution. Langmuir 1999, 15, 2316-2321. (26) Nakamoto, K. Infraed and raman spectra of inorganic and coordiantion compounds. Part A: Theory and Applications in Inorganic Chemistry. 5th Edition. John Wiley & Sons, Inc. 1997, 189-209. (27) Chabot, M.; Hoang, T.; Al-Abadleh, H.A. ATR-FTIR studies on the nature of surface complexes and desorption efficiency of p-arsanilic acid on iron (oxyhydr)oxides. Environ. Sci. Technol. 2009, 43, 3142–3147. (28) Johnston, Chad P.; Chrysochoou, M. Mechanisms of chromate adsorption on hematite Geochim. Cosmochim. Ac. 2014, 138, 146–157. (29) Elzinga, E.J.; Kretzschmar, R. In situ ATR-FTIR spectroscopic analysis of the co-adsorption of orthophosphate and Cd(II) onto hematite. Geochim. Cosmochim. Ac. 2013, 117, 53-64.

19

ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 28

Table 1. The Calculated IR Vibrational Frequencies (cm−1) and Assignment of Six Possible Phosphonate Group involved PMG Complexes with Goethite Surfaces. Solvent = H2O Assignment MM Ia MM IIa MB I MB II BB I BB II 1208 – 1181 – 1194 – δ (PO–H) of phosphonate group b 1129(1128) – 1126 (1122) – 1148(1142) υ (P=O) of phosphonate group – 1108(1102) – – – – υas (PO2) of uncomplexed P–O – 1037 – – – – υs (PO2) of uncomplexed P–O – – 1099 1048 (1052) 1105 1014(1018), υas (P–(OFe)2) of bidentate complexes – – 1002 957 (956) 1048 984(987), υs (P–(OFe)2) of bidentate complexes υ (P–OFe) of monodentate complexes

δ (PO–H) of phosphonate group υ (P=O) of phosphonate group υas (PO2) of uncomplexed P–O υs (PO2) of uncomplexed P–O υas (P–(OFe)2) of bidentate complexes υs (P–(OFe)2) of bidentate complexes υ (P–OFe) of monodentate complexes

1092(1094), 1054(1060)

1008(1004), 972(980)

MM I

MM II

1186



1118(1105)







935(936) –

Solvent = D2O MB I MB II

BB I

BB II

1160



1178





1130 (1118)



1146(1137)



1118(1102)











1030













1074

1044 (1052)

1096

1008(1014),





997

957 (956)

1040

1078(1073), 1044(1041)

1011(1005), 981(980)







981(984), 938(931) –

a

“I” and “II” represent protonated and deprotonated type, respectively. Values in the parentheses are experimental IR vibrational frequencies. MM: mononuclear monodentate; MB: mononuclear bidentate; BB: binuclear bidentate. All proposed interfacial configurations are shown in Figure 3.

b

20

ACS Paragon Plus Environment

Page 21 of 28

419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450

Environmental Science & Technology

Figure Captions

Figure 1. ATR-FTIR spectra of 1000 mg L−1 dissolved PMG (a) and sarcosine (b) in 0.1 M NaCl for pH values 1.5, 4.2, 9.0 and 12.5. Figure 2. Deconvoluted ATR-FTIR spectra of adsorbed PMG on goethite surfaces at pH 5, 6, 7, 8, and 9 at equilibrium. Figure 3. Phosphonate group involved PMG complexes with goethite surfaces. The possible interfacial configurations, including mononuclear monodentate (MM), mononuclear bidentate (MB), and binuclear bidentate (MB), were optimized. Each configuration has two types, namely, the phosphonate group with (I) and without (II) proton. The explicit H2O molecules were not shown for clarity. Fe: cyan; C: grey; O: red; N: blue; P: orange; H: white.

Figure 4. Synchronous (a, c, e) and asynchronous (b, d, f) correlation contour maps of dynamic IR spectra for the adsorption of PMG on goethite surface at pH 5, 7, and 9. The blue (red) regions were defined as negative (positive) correlation intensities. Figure 5. IR spectra of adsorbed PMG in H2O and D2O for pH/pD 5 (left), pH/pD 7 (middle), and pH/pD 9 (right). Figure 6. Adsorption isotherms of PMG (a) and phosphate (b) in the competitive systems at pH 7. Symbols represent the experimental results, and the solid lines are Langmuir adsorption model simulations. (c) Deconvoluted ATR-FTIR interfacial spectra of PMG (top), phosphate (middle), and the competition of phosphate with PMG (bottom) in 0.1 M NaCl solution at pH 7. The red values in the parentheses are relative peak areas, which are normalized to the most intense peak located at 1050 cm−1.

451 452 453 21

ACS Paragon Plus Environment

Environmental Science & Technology

454 455 456 457 458 459 460 461 462 463 464 465 466

Figure 1.

467 468 469 470 471 472 473 474 22

ACS Paragon Plus Environment

Page 22 of 28

Page 23 of 28

Environmental Science & Technology

475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490

Figure 2.

491 492 493 494 495 23

ACS Paragon Plus Environment

Environmental Science & Technology

496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511

The “I” and “II” after the abbreviation of configurations represent the phosphonate group with and without proton, respectively. Salient bonds are highlighted by red circles.

512 513

Figure 3.

514 515 516 517 24

ACS Paragon Plus Environment

Page 24 of 28

Page 25 of 28

Environmental Science & Technology

518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536

Figure 4.

537 538 25

ACS Paragon Plus Environment

Environmental Science & Technology

539 540 541 542 543 544 545 546 547 548 549

Wavenumber (cm−1)

550 551 552 553 554

Figure 5.

555 556 557 558

26

ACS Paragon Plus Environment

Page 26 of 28

Page 27 of 28

Environmental Science & Technology

559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578

Figure 6.

579 27

ACS Paragon Plus Environment

Environmental Science & Technology

580

28

ACS Paragon Plus Environment

Page 28 of 28