Mechanistic Insights from Discrete Molecular Dynamics Simulations of

Others, however, have investigated direct benefits of intentionally. 45 applying ... The details provided by DMD such as binding. 77 configuration ...
0 downloads 0 Views 878KB Size
Subscriber access provided by UNIVERSITY OF CONNECTICUT

Article

Mechanistic Insights from Discrete Molecular Dynamics Simulations of Pesticide-Nanoparticle Interactions Nicholas K. Geitner, Weilu Zhao, Feng Ding, Wei Chen, and Mark R. Wiesner Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b01674 • Publication Date (Web): 07 Jul 2017 Downloaded from http://pubs.acs.org on July 7, 2017

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 27

Environmental Science & Technology

1

Mechanistic Insights from Discrete Molecular Dynamics Simulations of

2

Pesticide− −Nanoparticle Interactions

3

Nicholas K Geitner,a Weilu Zhao,b Feng Ding,c Wei Chen,b Mark R Wiesnera*

4

a

5

Environmental Engineering, Duke University, Durham, NC 27708, USA

6

b

7

China

8

c

9

Corresponding Author:

Center for the Environmental Implications of NanoTechnology, Department of Civil and

College of Environmental Science and Engineering, Nankai University, Tianjin 300350,

Department of Physics and Astronomy, Clemson University, Clemson, SC 29634, USA

10

Box 90287, Durham, NC 27708

11

Phone: 919-660-5292

12

Fax: 919-660-5219

13

E-mail: [email protected]

14 15

ABSTRACT

16

Nano-scale particles have the potential to modulate the transport, lifetimes, and ultimate

17

uptake of pesticides that may otherwise be bound to agricultural soils. Engineered

18

nanoparticles provide a unique platform for studying these interactions. In this study, we

19

utilized discrete molecular dynamics (DMD) as a screening tool for examining

20

nanoparticle−pesticide adsorptive interactions. As a proof-of-concept, we selected a library

21

of 15 pesticides common in the United States and 4 nanomaterials with likely natural or 1 ACS Paragon Plus Environment

Environmental Science & Technology

22

incidental sources, and simulated all possible nanoparticle− −pesticide pairs. The resulting

23

adsorption coefficients derived from DMD simulations ranged over several orders of

24

magnitude, and in many cases were significantly stronger than pesticide adsorption on clay

25

surfaces, highlighting the significance of specific nano-scale phases as a preferential media

26

with which pesticides may associate. Binding was found to be significantly enhanced by

27

the capacity to form hydrogen bonds with slightly hydroxylated fullerols, highlighting the

28

importance of considering the precise nature of weathered nanomaterials as opposed to

29

pristine precursors. Results were compared to experimental adsorption studies using

30

selected pesticides, with a Pearson correlation coefficient of 0.97.

31 32

INTRODUCTION

33

The presence and implications of anthropogenic nanomaterials in various natural

34

environments is a rapidly growing body of research.1 This attention is due to the rapidly

35

growing applications of a wide variety of nanomaterials in both consumer and industrial

36

applications, ranging from silver nanoparticles in textiles to titania particles in sunscreens

37

or nano-scale ceria as diesel fuel additives. However, the release of naturally occurring and

38

incidental nanomaterials is several orders of magnitude greater than those of intentional

39

engineered origin,2 and in many cases, the engineered and natural or incidental material

40

may be virtually indistinguishable.

41

From a human health and exposure perspective, agricultural environments are

42

clearly of critical importance. Work in this area frequently focuses on the acute toxicity of

43

pristine nanomaterials to agricultural plants. Previous work has highlighted carbon

2 ACS Paragon Plus Environment

Page 2 of 27

Page 3 of 27

Environmental Science & Technology

44

nanotubes in soil affecting seed germination3 as well as graphene toxicity towards several

45

agricultural plants.4 Others, however, have investigated direct benefits of intentionally

46

applying nanomaterials to agricultural soil. For example, carbon nanoparticles from

47

biochar (that includes significant quantities of C60 and other fullerene-like materials) were

48

found to boost wheat plant growth.5 Potential beneficial effects of fullerols on a wide

49

variety of organisms have been reported.6 Still others employ a variety of metal and metal

50

oxide nanoparticles for contaminant remediation or the prevention of plant disease.7-8

51

Whether intentional or incidental, there are a wide variety and growing number of

52

agricultural exposure routes to a diverse population of nanomaterials.1, 5, 7, 9-12

53

Many nanomaterials, however, do not appear to be acutely toxic at environmentally

54

relevant concentrations and exposures.13-15 Therefore, attention has begun to turn to

55

secondary effects such as the trophic transfer (propagation through food webs) of these

56

materials.1,

57

delivering chemicals to cells

58

agricultural soils may alter the transport kinetics and subsequent uptakes of pesticides and

59

other agricultural chemicals into plants.21 In order to understand the potential for

60

nanomaterials to exhibit such modulation of pesticide transport in an agricultural setting, a

61

detailed understanding of interactions between nano-scale particles and agricultural

62

chemicals is required. The large number of agricultural chemicals and the diversity of

63

nano-scale phases challenge the experimental characterization of all possible interactions.

64

A predictive computational approach with high efficiency and accuracy has the potential of

65

accelerating the description of relevant interactions, while reducing reliance on expensive

66

and labor-intensive experimental studies.

12, 16-18

There have also been suggestions of “Trojan Horse” mechanisms in 19-20

Recent studies have reported that nanomaterials in

3 ACS Paragon Plus Environment

Environmental Science & Technology

67

Some computational studies have been done using molecular dynamics (MD) to

68

examine small molecule interactions with carbon nanotubes22 as well as polynomial fitting

69

based on previous experimental results and a biological surface adsorption index (BSAI).23

70

Discrete Molecular Dynamics (DMD) has been previously employed to simulate

71

nanoparticle protein corona formation,24 polymeric nanoparticle oil dispersants,25 and the

72

formation of stripe-like binary molecules on nanoparticle surfaces.26 Here, we consider

73

scenarios of agricultural relevance; the adsorption of commonly used pesticides with

74

reference nanomaterials such as fullerene, fullerols and ceria nanoparticles. By simulating

75

the interactions between several nanoparticles and a library of relevant pesticides, we aim

76

to provide a heuristic for quickly assessing combinations of chemical and nanoparticles of

77

interest for possible experimental study. The details provided by DMD such as binding

78

configuration, mechanisms, kinetics, and environmental sensitivity may also inform a so-

79

called “safer-by-design” approach to risk management for engineered nanoparticles.

80

Additionally, the computational efficiency of DMD allows for the simulation of larger,

81

more complex systems over longer time scales than traditional MD. To evaluate the

82

predictive power of this DMD-based approach, we compare the results with experimentally

83

measured binding affinities.

84

The nanoscales selected for the experimental portion of this study were a C60

85

fullerene, fullerols with either 8 (fullerol-8) or 24 (fullerol-24) hydroxyl groups, and ceria

86

nanoparticles. This set of nanoparticles represents carbon-based nanomaterials with

87

varying degrees of hydrophobicity as well as a metal oxide nanoparticle. Ceria

88

nanoparticles, a member of the broad class of metal oxide nanoparticles, find applications

89

is catalysis, silica polishing, and fuel additives, resulting in significant potential for

4 ACS Paragon Plus Environment

Page 4 of 27

Page 5 of 27

Environmental Science & Technology

90

environmental expsure.27-33 Inclusion of hydroxylated fullerene derivatives is motivated by

91

the systematic range of derivatizations of a C60 cage that these materials provide as well as

92

environmental relevance of fullerols and other derivatized or environmentally aged carbon-

93

cage nanomaterials34-36 as natural and incidental materials that may be present in soils.

94

Such carbon-based nanomaterials are commonly found as a product of carbon-based

95

material combustion, such as coal ash or fossil fuel emissions.1, 9-10, 37 Thirteen pesticides

96

were selected from the US Environmental Protection Agency list of most commonly used

97

pesticides in the United States, discarding some of those with high similarity in chemical

98

structure.

99

dichlorodiphenyldichloroethylene38

100

We

then

added

two

legacy (DDE,

pesticides

of

breakdown

particular product

interest, of

dichlorodiphenyltrichloroethane, DDT) and chlordane.39

101 102

METHODS

103

DMD Simulations

104

DMD is a subclass of MD approach which employs several improvements in calculation

105

efficiency, including implicit solvent, discretized potential functions, and re-calculation of

106

atomic ballistic equations only when atoms have participated in a collision.40 These

107

features yield significant gains in calculation speed and efficiency, allowing for the

108

simulation of larger molecular systems on longer time scales. Interatomic potentials

109

include electrostatic, van der Waals (VDW), hydrogen bond, and solvation interactions.

110

Solvation energy in these implicit solvent simulations was calculated using the

111

Lazaridis−Karplus EEF1 (Effective Energy Function 1) model.41 The Debye−Hückel

112

approximation was used to model electrostatic screening, with a Debye length of 48 Å, 5 ACS Paragon Plus Environment

Environmental Science & Technology

113

corresponding to a solution ionic strength of approximately 4 mM×e2, as may be expected

114

in freshwater and many pore water systems.

115

Small molecule ligands were modeled in all-atom DMD simulations with the

116

MedusaScore force field,42 which was parameterized on a large set of small molecule

117

ligands and transferrable to different molecular systems. The predictive power of

118

MedusaScore has been validated in various benchmark studies, including recent CSAR

119

(community structure-activity resource) blind ligand-receptor docking prediction

120

exercises43-44 where the performance of MedusaScore was among the best approaches in

121

predicting the near-native ligand-binding poses and binding affinities.

122

All nanomaterials and molecules were constructed using the Avogadro molecular

123

builder (v 1.1.1),45 including all heavy atoms and hydrogens. Ceria nanoparticles were

124

assigned a diameter of 2.7 nm and included 500 total atoms. Atoms buried in the core of

125

the particle were omitted to minimize the computational load for ceria nanoparticle

126

simulations. For those species with potentially charged moieties, a pH of 7.4 was assumed

127

and sodium ions were incorporated into the simulation to balance the system net charge.

128

When incorporating cerium into the DMD simulation package, the VDW constants of

129

atomic Ce27,

130

angles)48 were obtained from the literature. Specifically, the VDW radii (σVDW) and energy

131

(εVDW) of Ce were assigned 2.4 Å and 0.38 kcal/mol, respectively. The corresponding

132

desolvation energy (εsolv) was chosen as -1.0 kcal/mol (i.e., the nanoparticle is weakly

133

hydrophilic) in the EEF1 calculation. Clay aluminosilicate was similarly parameterized. Si:

134

εVDW = 0.30 kcal/mol, σVDW = 1.92 Å, εsolv = -5.0 kcal/mol; Al: εVDW = 0.28 kcal/mol, σVDW =

135

1.84 Å, εsolv = -5.0 kcal/mol (i.e., clay is more hydrophilic than CeO2).

46-47

and the hydrophobicity of CeO2 surfaces (from experimental contact

6 ACS Paragon Plus Environment

Page 6 of 27

Page 7 of 27

Environmental Science & Technology

136

Simulations were carried out with all possible pairings of a single nanoparticle and

137

agricultural agent in a simulation box size of 80 Å3 with periodic boundary conditions.

138

Unless otherwise noted, the temperature for each simulation was 0.6 kcal/(mol·kB), which

139

corresponds approximately to 300 K. Constant temperature was maintained using the

140

Anderson thermostat.49 After initialization, energy minimization was carried out for 10,000

141

time steps of approximately 500 ps each. Production simulations were subsequently run for

142

500,000 time steps (approximately 25 ns). We then extracted time-dependent data over the

143

entire simulation for estimating the Gibbs free energy of binding (∆G) and the number of

144

atomic contacts between nanoparticles and pesticides. Two atoms were defined as being in

145

contact when within 6.5 Å of each other. All simulations were carried out on the Duke

146

Compute Cluster, housed at Duke University. Each simulation was carried out on a single

147

CPU core and completed in approximately 1.9 hours of computational time.

148

To estimate the binding free energies, we first obtained the potential energy (E) of

149

the unbound system by calculating the mean and standard deviation values for all time

150

points where the molecule was far from the nanoparticle. The molecule and nanoparticle

151

were considered bound when the value of interatomic contacts was non-zero. During all

152

such time points, we again calculated the mean and standard deviations of the potential

153

energy of the bound state. We then used these two mean values to calculate the Δ of

154

binding for each nanoparticle–molecule pair. Assuming that the conformational flexibility

155

of a molecule in the unbound states was due to independent rotation of its rotatable bonds

156

and that the molecule lost this conformational flexibility upon binding, changes in entropy

157

were then estimated as the number of rotatable bonds in each pesticide, NRB, such that

7 ACS Paragon Plus Environment

Environmental Science & Technology

158

Δ ≈ −   . Subsequently, changes in free energy were calculated as ∆ = ∆ − ∆.

159

Log adsorption constants (log Kd) were then calculated per,

160 161

∆ = −  ,

(1)

where kB (kcal mol-1 K-1) is the Boltzmann constant and T is the system temperature, 300 K.

162 163 164

Pesticide Adsorption Experiments 1. Materials

165

Sublimed fullerene powder (C60, >99.5 %) was purchased from SES Research

166

(Houston, TX, USA). Cerium oxide nanopowder (99%), and permethrin (99.5%) were purchased from J&K Chemical (Shanghai,

170

China). Stock solutions of the pesticides were prepared in methanol, and were stored at -

171

20°C. The inorganic salts (NaH2PO4·2H2O and Na2HPO4·12H2O) were obtained from

172

VICTOR Co. (Tianjin, China).

173

Glass optical fibers coated with polyacrylate (thickness 35 mm; volume 15.4 µL/m)

174

were purchased from Polymicro Technologies (Phoenix, AZ). The fibers were cut to the

175

desired length (generally 1–3 cm, depending on the expected partition coefficient), cleaned

176

3 times by shaking in 50:50 methanol/water (v:v), then washed with deionized (DI) water

177

to remove the solvent, and stored in DI water until further use.

178

2. Preparation of nC60 stock suspension

8 ACS Paragon Plus Environment

Page 8 of 27

Page 9 of 27

Environmental Science & Technology

179

An nC60 stock suspension was prepared by magnetically stirring 150 mg C60 in 300

180

ml DI water for 30 min, followed by sonicating the mixture (using a sonication probe, 100

181

W) in an ice bath in the dark for 1 h. This step was repeated 3 times. The concentration of

182

C60 in the stock suspension was determined using an oxidation−toluene extraction

183

procedure.50

184

3. Procedures of negligible depletion-solid-phase microextraction

185

To determine the sorption coefficients of the pesticides to the fiber, the sorption

186

isotherms of the six pesticides to the fiber were obtained using previously developed

187

procedures.51-54 To initiate a fiber sorption experiment, first 20 ml of 10 mM phosphate

188

buffer (NaH2PO4/Na2HPO4, pH 7.0) was transferred to each of a series of 20-ml amber

189

glass vials. Next, a stock solution of a pesticide (in methanol) was added to each vial, and

190

the volume percentage of methanol was kept below 0.1% (v/v) to minimize cosolvent

191

effects. Next, a piece of fiber was added to the vial. The vial was capped and tumbled at 8

192

rpm at room temperature in the dark for 28 d. The time required to reach sorption

193

equilibrium was pre-determined. Then, the fibers were taken out, wiped with wet tissues,53-

194

54

195

other compounds) to analyze the mass of the pesticides on the fibers. The aqueous

196

solutions were extracted with hexane (and if needed, solvent-exchanged to methanol) to

197

analyze the concentration of the pesticides in the dissolved phase. All experiments were

198

run in duplicate. The sorption data were fitted with the linear sorption isotherm: Cfiber =

199

Kfiber ⋅ CW, where Cfiber (mg/L) and CW (mg/L) are the equilibrium concentrations of a

200

pesticide on the fiber and in the solution, respectively; and Kfiber (L/L) is the fiber–water

and extracted with hexane (for tefluthrin, bifenthrin, permethrin) or methanol (for all

9 ACS Paragon Plus Environment

Environmental Science & Technology

201

distribution coefficient. The fiber sorption isotherms are shown in SI Fig. S1, and the fitted

202

Kfiber values are summarized in SI Table S1.

203

4. Adsorption isotherms of pesticides to nC60 or CeO2

204

Prior to initiating an nC60 adsorption experiment, the stock suspension of nC60 was

205

diluted in 10 mM phosphate buffer (NaH2PO4/Na2HPO4, pH 7.0) to obtain an nC60

206

suspension of ~30 mg/L. Next, 20 mL the suspension was transferred to a series of 20-ml

207

amber glass vials. Then, the suspensions were spiked with a test pesticide (in methanol)

208

and tumbled at 8 rpm at room temperature in the dark for 7 d. Afterward, fibers were added

209

to the vials and were allowed to equilibrate for 28 d. Finally, the fibers were treated as

210

described above to analyze the concentrations of the pesticides on the fibers. The

211

concentrations of freely dissolved compounds were calculated based on the concentrations

212

on the fibers and the fiber–water distribution coefficients. The concentrations of the

213

pesticides on nC60 were calculated based on a mass balance approach.

214

For the nCeO2 adsorption experiments, 0.002 g CeO2 was added to 20 mL electrolyte

215

solution (10 mM phosphate buffer, NaH2PO4/Na2HPO4, pH 7.0)55, and then spiked with a

216

test pesticide (in methanol) and tumbled at 8 rpm at room temperature in the dark for 7 d.

217

Afterward, the vials were centrifuged (15,000 g, 10 min) and the supernatants were

218

withdrawn and extracted as stated above to analyze the concentrations of pesticides in the

219

aqueous solutions.56

220

5. Analytical methods

221

Chlorpyrifos was analyzed using a Waters high performance liquid chromatography

222

system equipped with a symmetry reversed-phase C18 column (4.6 × 150 mm), and was

10 ACS Paragon Plus Environment

Page 10 of 27

Page 11 of 27

Environmental Science & Technology

223

detected with a Waters 2489 UV/visible detector at a wavelength of 300 nm; the mobile

224

phase was acetonitrile–DI water (90:10, v:v; 1.0 ml/min). Tefluthrin, bifenthrin,

225

permethrin, chlordane, and p,p’-DDE were analyzed with a gas chromatograph (GC6890N,

226

Agilent Corp., Santa Clara, CA) equipped with an electron capture detector.57-59 No peaks

227

were detected in the spectra for potential degraded/transformed products of the compounds

228

tested.

229

In order to compare experimental adsorption coefficients to DMD simulation results,

230

experimental adsorption isotherms were extrapolated to very low adsorbent concentration,

231

as done by Zou et al.60 Briefly, the concentration in the adsorbed phase, q (mg/kg), was

232

plotted against the aqueous concentration of pesticide, Cw (mg/L). The slope of this curve

233

resulted in the theoretical Kd coefficient for a single pesticide–nanomaterial interaction and

234

was compared directly to DMD calculations.

235 236

RESULTS AND DISCUSSION

237

Figure 1 contains renderings of the 4 simulated nanomaterials and Table 1 summarizes the

238

basic physicochemical properties and chemical structure of each pesticide. Counts of

239

hydrogen bond acceptors include trifluoromethanes.61 An example plot of system energy

240

as a function of time from a simulation of chlorpyrifos and a ceria nanoparticle is provided

241

in Figure 2. These plots were then used to calculate the free energies of adsorption as

242

described in Methods.

11 ACS Paragon Plus Environment

Environmental Science & Technology

C O

Page 12 of 27

H Ce

A

B

C

D

243 244

Figure 1. Rendering of the nanoparticles simulated. A: C60 fullerene; B: fullerol-8; C: fullerol-24; D: ceria

245

nanoparticle

246 247

Table 1. Names and basic properties of the agricultural chemicals for DMD simulations. MW

log

Net

Name

NRBb

NHBAc

NHBD Structure d

(g/mol)

KOW

Charge

Aldicarb

190.26

1.13

0

5

4

1

Atrazine

215.69

2.61

0

4

5

2

a

12 ACS Paragon Plus Environment

Page 13 of 27

Environmental Science & Technology

Bifenthrin

422.87

6.00

0

7

5

0

Carbofuran

221.26

2.32

0

3

4

1

Chlordane

409.76

6.16

0

0

0

0

Chlorpyrifos

350.59

4.96

0

6

4

0

DDE

318.02

6.51

0

2

0

0

Dicamba

221.03

2.21

-1

2

3

1

13 ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 27

Glyphosate

169.07

-3.4

-3

4

6

4

Imidacloprid

255.66

0.57

0

4

7

1

Methyldithiocarbamate

107.19

N/A

-1

1

1

1

Metolachlor

283.80

3.13

0

7

3

0

Permethrin

391.29

6.5

0

7

3

0

Tefluthrin

418.74

6.4

0

6

5

0

Terbufos

288.42

4.48

0

8

2

0

248

a

249

b

250

c

Net charges were calculated at pH 7 NRB: number of rotatable bonds

NHBA: number of hydrogen bond acceptors

14 ACS Paragon Plus Environment

Page 15 of 27

251

Environmental Science & Technology

d

NHBD: number of hydrogen bond donors

252

253 254

Figure 2. A sample of simulation data output for 500,000 time steps. Green: change in system potential

255

energy relative to initial conditions. Blue: number of inter-atomic contacts between the pesticide and

256

nanoparticle, in this case between chlorpyrifos and nano ceria.

257 258

The resulting log Kd (adsorption coefficients) from DMD simulations are shown in

259

Figure 3 for all nanoparticle–pesticide pairs. Across all nanomaterials, these adsorption

260

coefficients cover a wide range, from near 0 (methyldithiocarbamate on highly

261

hydroxylated fullerols) to 8.85 (tefluthrin on slightly hydroxylated fullerols). DDE

262

exhibited the strongest adsorption to fullerenes of all simulated pesticides. This may be

263

surprising, as it is not the most hydrophobic pesticide in the library, and it contains the

264

same number of aromatic rings as several other species. However, the unique geometry of

265

DDE allows its two chlorobenzene rings to rotate and “cradle” the nanoparticle, resulting 15 ACS Paragon Plus Environment

Environmental Science & Technology

266

in its extremely high binding affinity (Figure 4A). There are some noteworthy trends

267

observed between nanomaterials, as well. First, adsorption to fullerol-8 is nearly

268

universally stronger than that to C60 fullerenes. However, further hydroxylation to fullerol-

269

24 causes a decrease in adsorption strength for all tested chemicals. This drop is highly

270

dependent on the chemical agent; in some cases, it returns to approximately the same value

271

as fullerene adsorption (i.e. aldicarb), in some cases remains stronger than on fullerenes

272

(i.e. atrazine), and in many others becomes significantly weaker. We hypothesize that these

273

differences are due to the balance between interaction mechanisms. While the fullerenes

274

rely solely on van der Waals, π–π, and hydrophobic interactions for adsorption, fullerol-8

275

nanoparticles also have the capacity for hydrogen bonding while still allowing access to

276

the hydrophobic carbon surface. A typical example of this phenomenon can be seen in

277

Figure 4B, in which a bifenthrin molecule was observed to simultaneously exhibit both

278

hydrogen bonding (dotted yellow line) and strong hydrophobic interactions with a single

279

fullerol-8 nanoparticle. This mechanism is further evidenced in noting that chlordane, with

280

no capacity for forming hydrogen bonds, does not exhibit enhanced binding to fullerol-8

281

particles. An abundance of hydroxyl groups in fullerol-24 blocks the access to hydrophobic

282

surfaces, thus reducing the potential for hydrophobicity-driven adsorption (Figure 4C).

283

The weakest binding was observed for charged and more water-soluble pesticides, as

284

may be expected. Methyldithiocarbamate exhibited particularly weak binding on highly

285

hydroxylated Fullerols. This is likely due to its high water solubility and limited capacity

286

for hydrogen bonding. Aldicarb and atrazine, while not charged, bound weakly to each

287

nanoparticle due to slight water solubility and a lack of π–π interaction capacity. The

288

strongest binding, in contrast, was generally observed for pesticides with at least 1

16 ACS Paragon Plus Environment

Page 16 of 27

Page 17 of 27

Environmental Science & Technology

289

aromatic ring and significant capacity for hydrogen bonding on slightly hydroxylated

290

fullerols, as described above.

291

In most cases, adsorption on ceria nanoparticles is approximately equal to or weaker

292

than that on fullerenes. This is generally due to the lower hydrophobicity of the ceria

293

nanoparticle. Terbufos most significantly exhibited stronger adsorption to ceria

294

nanoparticles than to fullerenes or fullerols. We hypothesize that this is due to its highly

295

flexible “arms” containing oxygen and methyl groups, available to interact with the surface

296

through multiple mechanisms simultaneously (Figure 4D). These mechanisms are more

297

diverse than those available for adsorption to the purely carbon-based fullerenes.

298

Additionally, due to the lack of hydrogen bonding capacity, pristine ceria is still more

299

hydrophobic than either fullerol (see Methods),48 and previous work has illuminated

300

several mechanisms of oxygen-mediated adsorption onto CeO2 surfaces which appears to

301

enhance turbufos binding to the ceria nanoparticle.62 10 9

Fullerene

Fullerol-8

Fullerol-24

Ceria

8 7

log Kd

6 5 4 3 2 1 0

302 303

Figure 3. Predicted log Kd values for each nanoparticle-pesticide pair from DMD simulations

17 ACS Paragon Plus Environment

Environmental Science & Technology

304 305

A

C B O H Ce Cl F S P

C

D

E

306 307

Figure 4. A collection of simulation snapshots highlighting interaction mechanisms. A: DDE interaction with

308

a fullerene; B: bifenthrin interaction with fullerol-8; C: bifenthrin with fullerol-24; D: terbufos with a ceria

309

nanoparticle. Yellow dotted lines indicate the formation of hydrogen bonds; E: the two observed binding

310

modes of permethrin on fullerenes

311 312

The increase in the number of hydroxyl groups from fullerene to fullerol-8 and

313

fullerol-24 results into increased capability to form hydrogen bonds with water molecules,

314

and thus, increasing water solubility. The fact that fullerol-8 interacted most strongly with

315

most chemicals also suggests that the adsorption of those chemicals is not driven solely by

316

their low water solubility. We confirmed this by plotting the predicted adsorption

317

coefficients against each chemical’s log octanol–water partition coefficient, KOW, and 18 ACS Paragon Plus Environment

Page 18 of 27

Page 19 of 27

Environmental Science & Technology

318

performed correlation analysis (SI, Figures S4-7). Pearson correlation coefficients between

319

calculated adsorption coefficients and log KOW for adsorption to fullerenes, fullerol-8,

320

fullerol-24, and ceria nanoparticles are 0.78, 0.75, 0.58, and 0.82, respectively. Thus, while

321

there was a trend seen between adsorption and chemical hydrophobicity in these cases, this

322

property alone is insufficient to precisely predict actual adsorption strength and rank

323

between chemicals. This is because KOW values fail to account for other factors important

324

to molecular adsorption to nanomaterials including molecular flexibility, entropic effects,

325

and specific modes of interaction such as hydrogen bonding and π–π interactions. The

326

many physical factors that must be considered for all species involved make prediction

327

difficult from a single or small subset of simple physical parameters. The use of DMD,

328

however, allows for adsorption coefficient prediction with little prior knowledge of the

329

chemical or nanomaterial in question other than chemical structure, and can also provide

330

additional insight such as the binding mechanisms at work as well as kinetic and

331

configurational information. System parameters such as ionic strength and temperature can

332

also be freely adjusted in DMD, which allows for the facile investigation of a diverse set of

333

scenarios.

334

Importantly, we also experimentally verified the predictive power of our DMD

335

simulations. We selected 2 nanomaterials (fullerenes and ceria nanoparticles) and 6

336

pesticides with which to obtain experimental adsorption isotherms. From these isotherms,

337

we extracted values for Kd and compared them to the simulation results, shown in Figure 5.

338

The dotted line represents the ideal case in which the experimental results exactly equal the

339

simulated predictions. The average absolute difference in experimental and predicted log

340

Kd values is 0.16. A correlation analysis yielded a Pearson correlation coefficient of 0.97.

19 ACS Paragon Plus Environment

Environmental Science & Technology

341

There is one outlier adsorption, permethrin on fullerenes. We hypothesize that this may be

342

due to the long, fairly inflexible structure of permethrin. In examining simulation

343

snapshots, we observed that permethrin exhibited two distinct binding modes on fullerenes

344

(Figure 4E). In one mode, one benzene ring and the hydrophobic triangular tail are bound

345

to the nanoparticles; in the other, both benzene rings are bound to the surface, but the tail is

346

not due to insufficient flexibility of the molecule. This limits the potential binding capacity

347

on a single fullerene particle. However, this limitation is not present on the much larger

348

ceria nanoparticles, and is not expected to be present for a small cluster of fullerenes, as

349

was likely the case during adsorption isotherm experiments. This hypothesis will be tested

350

in future studies. Overall, this experimental verification confirms the validity and high

351

precision of our simulations, which can be performed far more quickly and economically

352

than experimental measures while also providing a rich variety of mechanistic information. 9.5

Fullerenes Ceria

9.0

log Kd - Exprimental

8.5 8.0 7.5 7.0 6.5 6.0 5.5 5.0 4.5 4.5

353

5.5

6.5

7.5

8.5

9.5

log Kd - DMD Simulations

354

Figure 5. Correlation plot comparing log Kd values from DMD simulations and experimental adsorption

355

isotherms for 6 pesticides adsorbing to fullerenes (green squares) and nano ceria (orange triangles).

20 ACS Paragon Plus Environment

Page 20 of 27

Page 21 of 27

Environmental Science & Technology

356

In using this method to determine which nanoparticle–pesticide pairs may

357

potentially affect pesticide transport and impact through binding, one must also consider

358

competitive adsorption surfaces already present in the environment. To this end, we

359

selected 5 of the strongest interacting pesticides and compared their adsorption coefficients

360

on nanoparticles to those on a clay surface, which is one of several possible such

361

competitive surfaces. For a model clay surface, we constructed a fragment of kaolinite 1

362

aluminosilicate layer thick and approximately 2 × 3.5 nm2 in size, shown in Figure 6a with

363

a tefluthrin molecule for illustration. As shown in Figure 6b, we found that each of these 5

364

pesticides adsorbed much more weakly to the clay surface than to Fullerol-8 nanoparticles,

365

and much more weakly than to fullerenes in all but one case. Because the adsorption onto

366

the nanomaterials is so much stronger than onto clay, these adsorption processes would be

367

expected to be environmentally relevant even at low nanoparticle concentrations.

368

Additionally, the mechanistic insights described above will be particularly important in

369

considering weathered nanomaterials, or those which have previously interacted with

370

organic matter, which as a result are expected to exhibit significantly different binding

371

capacities for organic co-contaminants, further highlighting these findings’ environmental

372

relevance. Mechanistic and quantitative results from this study will form the basis of

373

experimental and targeted computational studies using those nanoparticle-pesticide pairs

374

with the highest binding affinities. These studies should capture the complexity of

375

agricultural soil conditions more completely. Specifically, additional studies on the impact

376

of various forms of organic matter on the transport of the pesticides which most strongly

377

bind with nanoparticles are left for future studies. However, effects are expected to range

21 ACS Paragon Plus Environment

Environmental Science & Technology

378

from an additional sink of pesticide through adsorption and competitive binding with clay

379

and nanoparticle surfaces, depending on the form of organic matter.63

A

10

B

Fullerene

Fullerol-8

Clay

9

8

log Kd

7

6

5

4

3

2

380

Chlordane

Chlorpyrifos

DDE

Imidaclorpid

Tefluthrin

381

Figure 6. A: Rendering of tefluthrin adsorbed to the surface of kaolinite fragment. B: Comparing select

382

pesticide adsorption to fullerene (Green), fullerol-8 (Blue) and kaolinite clay (Brown hashed) surfaces.

383

Acknowledgement

384

This material is based upon work supported by the National Science Foundation (NSF) and

385

the Environmental Protection Agency (EPA) under NSF Cooperative Agreement EF-

386

0830093, Center for the Environmental Implications of NanoTechnology (CEINT). Any

387

opinions, findings, conclusions or recommendations expressed in this material are those of

388

the author(s) and do not necessarily reflect the views of the NSF or the EPA. This work

389

has not been subjected to EPA review and no official endorsement should be inferred. 22 ACS Paragon Plus Environment

Page 22 of 27

Page 23 of 27

390

Environmental Science & Technology

This work was further supported by NSF CAREER CBET-1553945 (Ding) and the

391

Ministry of Science and Technology of China 2014CB932001 (Chen).

392

Supplementary Information

393

Additional data pertaining to experimental adsorption isotherms, including statistical

394

analysis, controls, and method calibration. Also includes correlation plots for KOW vs

395

calculated Kd values.

396 397

CITATIONS

398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424

1. Biswas, P.; Wu, C.-Y., Nanoparticles and the Environment. Journal of the Air & Waste Management Association 2005, 55 (6), 708-746. 2. Wiesner, M. R.; Lowry, G. V.; Casman, E.; Bertsch, P. M.; Matson, C. W.; Di Giulio, R. T.; Liu, J.; Hochella, M. F., Jr., Meditations on the ubiquity and mutability of nano-sized materials in the environment. ACS Nano 2011, 5 (11), 8466-70. 3. Khodakovskaya, M.; Dervishi, E.; Mahmood, M.; Xu, Y.; Li, Z.; Watanabe, F.; Biris, A. S., Carbon nanotubes are able to penetrate plant seed coat and dramatically affect seed germination and plant growth. ACS Nano 2009, 3 (10), 3221-7. 4. Begum, P.; Ikhtiari, R.; Fugetsu, B., Graphene phytotoxicity in the seedling stage of cabbage, tomato, red spinach, and lettuce. Carbon 2011, 49 (12), 3907-3919. 5. Saxena, M.; Maity, S.; Sarkar, S., Carbon nanoparticles in ‘biochar’ boost wheat (Triticum aestivum) plant growth. RSC Adv. 2014, 4 (75), 39948. 6. Gao, J.; Wang, Y.; Folta, K. M.; Krishna, V.; Bai, W.; Indeglia, P.; Georgieva, A.; Nakamura, H.; Koopman, B.; Moudgil, B., Polyhydroxy fullerenes (fullerols or fullerenols): beneficial effects on growth and lifespan in diverse biological models. PLoS One 2011, 6 (5), e19976. 7. Ram, P.; Vivek, K.; Kumar, S. P., Nanotechnology in sustainable agriculture: Present concerns and future aspects. African Journal of Biotechnology 2014, 13 (6), 705-713. 8. Dimkpa, C. O.; McLean, J. E.; Britt, D. W.; Anderson, A. J., Bioactivity and Biomodification of Ag, ZnO, and CuO Nanoparticles with Relevance to Plant Performance in Agriculture. Industrial Biotechnology 2012, 8 (6), 344-357. 9. Utsunomiya, S.; Jensen, K. A.; Keeler, G. J.; Ewing, R. C., Uraninite and Fullerene in Atmospheric Particulates. Environmental Science & Technology 2002, 36 (23), 4943-4947. 10. Silva, L.; DaBoit, K.; Serra, C.; Mardon, S.; Hower, J., Fullerenes and Metallofullerenes in Coal-Fired Stoker Fly Ash. Coal Combustion and Gasification Products 2010, 2, 66-79. 11. Navarro, D. A.; Kookana, R. S.; McLaughlin, M. J.; Kirby, J. K., Fullerol as a Potential Pathway for Mineralization of Fullerene Nanoparticles in Biosolid-Amended Soils. Environmental Science & Technology Letters 2016, 3 (1), 7-12. 23 ACS Paragon Plus Environment

Environmental Science & Technology

425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468

Page 24 of 27

12. Nair, R.; Varghese, S. H.; Nair, B. G.; Maekawa, T.; Yoshida, Y.; Kumar, D. S., Nanoparticulate material delivery to plants. Plant Science 2010, 179 (3), 154-163. 13. Colman, B. P.; Espinasse, B.; Richardson, C. J.; Matson, C. W.; Lowry, G. V.; Hunt, D. E.; Wiesner, M. R.; Bernhardt, E. S., Emerging contaminant or an old toxin in disguise? Silver nanoparticle impacts on ecosystems. Environ Sci Technol 2014, 48 (9), 5229-36. 14. Levard, C.; Hotze, E. M.; Colman, B. P.; Dale, A. L.; Truong, L.; Yang, X. Y.; Bone, A. J.; Brown, G. E., Jr.; Tanguay, R. L.; Di Giulio, R. T.; Bernhardt, E. S.; Meyer, J. N.; Wiesner, M. R.; Lowry, G. V., Sulfidation of silver nanoparticles: natural antidote to their toxicity. Environ Sci Technol 2013, 47 (23), 13440-8. 15. Baun, A.; Hartmann, N. B.; Grieger, K.; Kusk, K. O., Ecotoxicity of engineered nanoparticles to aquatic invertebrates: a brief review and recommendations for future toxicity testing. Ecotoxicology 2008, 17 (5), 387-95. 16. Ferry, J. L.; Craig, P.; Hexel, C.; Sisco, P.; Frey, R.; Pennington, P. L.; Fulton, M. H.; Scott, I. G.; Decho, A. W.; Kashiwada, S.; Murphy, C. J.; Shaw, T. J., Transfer of gold nanoparticles from the water column to the estuarine food web. Nat Nanotechnol 2009, 4 (7), 441-4. 17. Rico, C. M.; Majumdar, S.; Duarte-Gardea, M.; Peralta-Videa, J. R.; Gardea-Torresdey, J. L., Interaction of nanoparticles with edible plants and their possible implications in the food chain. J Agric Food Chem 2011, 59 (8), 3485-98. 18. Geitner, N. K.; Marinakos, S. M.; Guo, C.; O'Brien, N.; Wiesner, M. R., Nanoparticle Surface Affinity as a Predictor of Trophic Transfer. Environ Sci Technol 2016, 50 (13), 6663-9. 19. Hsiao, I. L.; Hsieh, Y. K.; Wang, C. F.; Chen, I. C.; Huang, Y. J., Trojan-horse mechanism in the cellular uptake of silver nanoparticles verified by direct intra- and extracellular silver speciation analysis. Environ Sci Technol 2015, 49 (6), 3813-21. 20. Kovochich, M.; Espinasse, B.; Auffan, M.; Hotze, E. M.; Wessel, L.; Xia, T.; Nel, A.; Wiesner, M. R., Comparative Toxicity of C60 Aggregates towards mammalian cells: Role of Tetrahydrofuran (THF) decomposition. Environmental Science & Technology 2009, 43 (16), 6378-6384. 21. De La Torre-Roche, R.; Hawthorne, J.; Deng, Y.; Xing, B.; Cai, W.; Newman, L. A.; Wang, Q.; Ma, X.; Hamdi, H.; White, J. C., Multiwalled carbon nanotubes and c60 fullerenes differentially impact the accumulation of weathered pesticides in four agricultural plants. Environ Sci Technol 2013, 47 (21), 12539-47. 22. Comer, J.; Chen, R.; Poblete, H.; Vergara-Jaque, A.; Riviere, J. E., Predicting Adsorption Affinities of Small Molecules on Carbon Nanotubes Using Molecular Dynamics Simulation. ACS Nano 2015, 9 (12), 11761-74. 23. Chen, R.; Zhang, Y.; Monteiro-Riviere, N. A.; Riviere, J. E., Quantification of nanoparticle pesticide adsorption: computational approaches based on experimental data. Nanotoxicology 2016, 10 (8), 1118-28. 24. Ding, F.; Radic, S.; Chen, R.; Chen, P.; Geitner, N. K.; Brown, J. M.; Ke, P. C., Direct observation of a single nanoparticle-ubiquitin corona formation. Nanoscale 2013, 5 (19), 9162-9. 25. Geitner, N. K.; Wang, B.; Andorfer, R. E.; Ladner, D. A.; Ke, P. C.; Ding, F., Structurefunction relationship of PAMAM dendrimers as robust oil dispersants. Environ Sci Technol 2014, 48 (21), 12868-75. 26. Ge, X.; Ke, P. C.; Davis, T. P.; Ding, F., A Thermodynamics Model for the Emergence of a Stripe-like Binary SAM on a Nanoparticle Surface. Small 2015, 11 (37), 4894-9. 24 ACS Paragon Plus Environment

Page 25 of 27

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

Environmental Science & Technology

27. Hung, V. V.; Thanh, L. T. M.; Masuda-Jindo, K., Study of thermodynamic properties of cerium dioxide under high pressures. Computational Materials Science 2010, 49 (4), S355-S358. 28. Si, R.; Flytzani-Stephanopoulos, M., Shape and crystal-plane effects of nanoscale ceria on the activity of Au-CeO2 catalysts for the water-gas shift reaction. Angew Chem Int Ed Engl 2008, 47 (15), 2884-7. 29. Reed, K.; Cormack, A.; Kulkarni, A.; Mayton, M.; Sayle, D.; Klaessig, F.; Stadler, B., Exploring the properties and applications of nanoceria: is there still plenty of room at the bottom? Environ. Sci.: Nano 2014, 1 (5), 390-405. 30. Parveen, S.; Misra, R.; Sahoo, S. K., Nanoparticles: a boon to drug delivery, therapeutics, diagnostics and imaging. Nanomedicine 2012, 8 (2), 147-66. 31. Gantt, B.; Hoque, S.; Willis, R. D.; Fahey, K. M.; Delgado-Saborit, J. M.; Harrison, R. M.; Erdakos, G. B.; Bhave, P. V.; Zhang, K. M.; Kovalcik, K.; Pye, H. O., Near-road modeling and measurement of cerium-containing particles generated by nanoparticle diesel fuel additive use. Environ Sci Technol 2014, 48 (18), 10607-13. 32. Erdakos, G. B.; Bhave, P. V.; Pouliot, G. A.; Simon, H.; Mathur, R., Predicting the effects of nanoscale cerium additives in diesel fuel on regional-scale air quality. Environ Sci Technol 2014, 48 (21), 12775-82. 33. Andreescu, D.; Bulbul, G.; Özel, R. E.; Hayat, A.; Sardesai, N.; Andreescu, S., Applications and implications of nanoceria reactivity: measurement tools and environmental impact. Environ. Sci.: Nano 2014, 1 (5), 445-458. 34. Pycke, B. F.; Chao, T. C.; Herckes, P.; Westerhoff, P.; Halden, R. U., Beyond nC60: strategies for identification of transformation products of fullerene oxidation in aquatic and biological samples. Anal Bioanal Chem 2012, 404 (9), 2583-95. 35. Brant, J. A.; Labille, J.; Robichaud, C. O.; Wiesner, M., Fullerol cluster formation in aqueous solutions: implications for environmental release. J Colloid Interface Sci 2007, 314 (1), 281-8. 36. Schreiner, K. M.; Filley, T. R.; Blanchette, R. A.; Bowen, B. B.; Bolskar, R. D.; Hockaday, W. C.; Masiello, C. A.; Raebiger, J. W., White-Rot Basidiomycete-Mediated Decomposition of C60Fullerol. Environmental Science & Technology 2009, 43 (9), 3162-3168. 37. Bakshi, S.; He, Z. L.; Harris, W. G., Natural Nanoparticles: Implications for Environment and Human Health. Critical Reviews in Environmental Science and Technology 2014, 45 (8), 861904. 38. Beyer, W. N.; Gale, R. W., Persistence and changes in bioavailability of dieldrin, DDE, and heptachlor epoxide in earthworms over 45 years. Ambio 2013, 42 (1), 83-9. 39. Lee, W. Y.; Lannucci-Berger, W.; Eitzer, B. D.; White, J. C.; Mattina, M. I., Persistent organic pollutants in the environment: chlordane residues in compost. J Environ Qual 2003, 32 (1), 224-31. 40. Proctor, E. A.; Ding, F.; Dokholyan, N. V., Discrete molecular dynamics. Wiley Interdisciplinary Reviews: Computational Molecular Science 2011, 1 (1), 80-92. 41. Lazaridis, T.; Karplus, M., Effective energy functions for protein structure prediction. Curr Opin Struct Biol 2000, 10 (2), 139-45. 42. Yin, S.; Biedermannova, L.; Vondrasek, J.; Dokholyan, N. V., MedusaScore: an accurate force field-based scoring function for virtual drug screening. J Chem Inf Model 2008, 48 (8), 1656-62. 25 ACS Paragon Plus Environment

Environmental Science & Technology

513 514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555

Page 26 of 27

43. Nedumpully-Govindan, P.; Jemec, D. B.; Ding, F., CSAR Benchmark of Flexible MedusaDock in Affinity Prediction and Nativelike Binding Pose Selection. J Chem Inf Model 2016, 56 (6), 1042-52. 44. Ding, F.; Dokholyan, N. V., Incorporating backbone flexibility in MedusaDock improves ligand-binding pose prediction in the CSAR2011 docking benchmark. J Chem Inf Model 2013, 53 (8), 1871-9. 45. Hanwell, M. D.; Curtis, D. E.; Lonie, D. C.; Vandermeersch, T.; Zurek, E.; Hutchison, G. R., Avogadro: an advanced semantic chemical editor, visualization, and analysis platform. J Cheminform 2012, 4 (1), 17. 46. Conesa, J. C., Computer Modeling of Surfaces and Defects on Cerium Dioxide. Surface Science 1995, 339 (3), 337-352. 47. Sayle, T. X. T.; Parker, S. C.; Catlow, C. R. A., The Role of Oxygen Vacancies on Ceria Surfaces in the Oxidation of Carbon-Monoxide. Surface Science 1994, 316 (3), 329-336. 48. Azimi, G.; Dhiman, R.; Kwon, H. M.; Paxson, A. T.; Varanasi, K. K., Hydrophobicity of rareearth oxide ceramics. Nat Mater 2013, 12 (4), 315-20. 49. Andersen, H. C., Molecular dynamics simulations at constant pressure and/or temperature. The Journal of Chemical Physics 1980, 72 (4), 2384. 50. Wang, L.; Hou, L.; Wang, X.; Chen, W., Effects of the preparation method and humic-acid modification on the mobility and contaminant-mobilizing capability of fullerene nanoparticles (nC60). Environmental science. Processes & impacts 2014, 16 (6), 1282-9. 51. Wang, F.; Haftka, J. J.; Sinnige, T. L.; Hermens, J. L.; Chen, W., Adsorption of polar, nonpolar, and substituted aromatics to colloidal graphene oxide nanoparticles. Environmental pollution 2014, 186, 226-33. 52. Wang, F.; Wang, F.; Zhu, D.; Chen, W., Effects of sulfide reduction on adsorption affinities of colloidal graphene oxide nanoparticles for phenanthrene and 1-naphthol. Environmental pollution 2015, 196, 371-8. 53. Heringa, M. B.; Hermens, J. L. M., Measurement of free concentrations using negligible depletion-solid phase microextraction (nd-SPME). TrAC Trends in Analytical Chemistry 2003, 22 (9), 575-587. 54. ter Laak, T. L.; Durjava, M.; Struijs, J.; Hermens, J. L. M., Solid Phase Dosing and Sampling Technique To Determine Partition Coefficients of Hydrophobic Chemicals in Complex Matrixes. Environmental Science & Technology 2005, 39 (10), 3736-3742. 55. Ji, P.; Zhang, J.; Chen, F.; Anpo, M., Study of adsorption and degradation of acid orange 7 on the surface of CeO2 under visible light irradiation. Applied Catalysis B: Environmental 2009, 85 (3-4), 148-154. 56. Vlasova, N. N.; Golovkova, L. P.; Stukalina, N. G., Adsorption of organic acids on a cerium dioxide surface. Colloid Journal 2015, 77 (4), 418-424. 57. Casas, V.; Llompart, M.; Garcia-Jares, C.; Cela, R.; Dagnac, T., Multivariate optimization of the factors influencing the solid-phase microextraction of pyrethroid pesticides in water. Journal of chromatography. A 2006, 1124 (1-2), 148-56. 58. Khan, I. A.; Riazuddin; Parveen, Z.; Ahmed, M., Multi-residue determination of synthetic pyrethroids and organophosphorus pesticides in whole wheat flour using gas chromatography. Bulletin of environmental contamination and toxicology 2007, 79 (4), 454-8.

26 ACS Paragon Plus Environment

Page 27 of 27

556 557 558 559 560 561 562 563 564 565 566 567 568 569 570

Environmental Science & Technology

59. Sayed, M., Analysis of Pesticides in Water Samples and Removal of Monocrotophos by γIrradiation. Journal of Analytical & Bioanalytical Techniques 2014, 05 (01). 60. Zou, M.; Zhang, J.; Chen, J.; Li, X., Simulating adsorption of organic pollutants on finite (8,0) single-walled carbon nanotubes in water. Environ Sci Technol 2012, 46 (16), 8887-94. 61. Hao, G. F.; Wang, F.; Li, H.; Zhu, X. L.; Yang, W. C.; Huang, L. S.; Wu, J. W.; Berry, E. A.; Yang, G. F., Computational discovery of picomolar Q(o) site inhibitors of cytochrome bc1 complex. J Am Chem Soc 2012, 134 (27), 11168-76. 62. Li, C.; Sakata, Y.; Arai, T.; Domen, K.; Maruya, K.-i.; Onishi, T., Carbon monoxide and carbon dioxide adsorption on cerium oxide studied by Fourier-transform infrared spectroscopy. Part 1.—Formation of carbonate species on dehydroxylated CeO2, at room temperature. Journal of the Chemical Society, Faraday Transactions 1: Physical Chemistry in Condensed Phases 1989, 85 (4), 929. 63. Radic, S.; Geitner, N. K.; Podila, R.; Kakinen, A.; Chen, P.; Ke, P. C.; Ding, F., Competitive binding of natural amphiphiles with graphene derivatives. Sci Rep 2013, 3, 2273.

571 572

573 574

TOC image

27 ACS Paragon Plus Environment