New Insight into the Aggregation of Graphene Oxide Using Molecular

Aug 3, 2017 - A comparative experimental and molecular dynamics (MD) simulation study was carried out to investigate the aggregation of graphene oxide...
0 downloads 0 Views 1MB Size
Subscriber access provided by Warwick University Library

Article

New Insight into the Aggregation of Graphene Oxide Using Molecular Dynamics Simulations and Extended Derjaguin-Landau-Verwey-Overbeek Theory Huan Tang, Ying Zhao, Xiaonan Yang, Dongmei Liu, Penghui Shao, Zhigao Zhu, Sujie Shan, Fuyi Cui, and Baoshan Xing Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b01668 • Publication Date (Web): 03 Aug 2017 Downloaded from http://pubs.acs.org on August 3, 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

New Insight into the Aggregation of Graphene Oxide Using Molecular Dynamics

2

Simulations and Extended Derjaguin-Landau-Verwey-Overbeek Theory

3

Huan Tang

4

Zhu a,b, Sujie Shan a,b, Fuyi Cui a,b,*, and Baoshan Xing c,*

5

a

6

b

7

Harbin 150090, China

8

c

a,b,c

, Ying Zhao

a,b

, Xiaonan Yang

a,b

, Dongmei Liu

a,b

, Penghui Shao

a,b

, Zhigao

State Key Laboratory of Urban Water Resource and Environment, Harbin 150090, China School of Municipal and Environmental Engineering, Harbin Institute of Technology,

Stockbridge School of Agriculture, University of Massachusetts, Amherst, MA, 01003, USA

9

1

ACS Paragon Plus Environment

Environmental Science & Technology

Table of Contents (TOC) Image Amount of H-bond

10

400 Lower pH

Higher pH

320

0

Interaction Energy (mJ/m2)

11

12

Page 2 of 27

5000 Time (ps)

10000

XDLVO

0 pH=3 pH=6 pH=9

-20 0

2 4 Separation distance (nm)

13

2

ACS Paragon Plus Environment

Page 3 of 27

Environmental Science & Technology

14

ABSTRACT

15

A comparative experimental and molecular dynamics (MD) simulation study was carried out

16

to investigate the aggregation of graphene oxide (GO). Mechanisms behind the effects of

17

solution chemistries (pH, metal ions, and tannic acid (TA)) and GO topology (carboxyl

18

content, GO size, and GO thickness) were uncovered. For example, MD results showed that

19

more hydrogen bonds formed between GO and water at higher pH, according well with the

20

increased hydrophilicity of GO calculated based on contact angle measurements. Radial

21

distribution functions analysis suggested Ca2+ interacted more strongly with GO than Na+,

22

which explained the experimental observations that Ca2+ was more effective in accelerating

23

the aggregation process than Na+. The adsorption-bridging and steric effects of TA were

24

simulated, and TA was found to be unfolded upon wrapping on GOs, leading to an increased

25

capacity for ion and solvent binding. The evaluations of contributions to GO hydrophilicity,

26

electrostatic energy, and intensities of interactions with metal ions indicated carboxyl group is

27

the essential functional group in mediating the stability of GO. Overall, by combining MD

28

simulations with experimental measurements, we provided molecular-level understandings

29

towards the aggregation of GO, indicating MD, if used properly, can be applied as a useful

30

tool to obtain insights into the aggregation of nanomaterials.

31

INTRODUCTION

32

Graphene oxide, abbreviated as GO, has been receiving increasing attention due to its

33

exceptional properties.1, 2 GOs are widely used in the field of electronics, composite materials,

34

and biomedical applications.3 With many promising applications, there are growing concerns

35

that GOs will enter natural water bodies throughout their life cycle. Recent studies have 3

ACS Paragon Plus Environment

Environmental Science & Technology

5

Page 4 of 27

36

shown GOs are cytotoxic toward human cells,4,

37

understand their fate in water and evaluate their environmental risks. Since the aggregation of

38

nanoparticles (including GO) is one of the most important factors that ultimately control their

39

fate,6 it is necessary to explore the aggregation of GOs in water.

demonstrating the research need to

40

Aggregation of GOs in aquatic environments has been studied extensively in recent

41

years.7-10 Findings from these studies illustrated that the aggregation of GO followed

42

Derjaguin - Landau - Verwey - Overbeek (DLVO) theory; however, hydrophobic force was

43

not considered in DLVO theory, and extended DLVO (XDLVO) theory should be employed

44

to fully understand the aggregation of GOs. The response of hydrodynamic diameters (Dh) to

45

different solution chemistries were measured to investigate the hydrodynamic properties of

46

GO, and electrophoretic mobility (EPM) comparisons were employed to understand the

47

effects of pH and ionic strength (IS). The increased stability at high pH was attributed to the

48

increased electrostatic repulsion between GOs, however, Wu et al. claimed it was not

49

sufficient to explain the GO stabilities under different pH regimes by using EPM

50

measurement alone.10 In addition, previous investigations mainly focused on macroscopic

51

experiments, elucidating the roles of pH, IS, and natural organic matter (NOM). Limited

52

efforts have been devoted to illustrate how these physiochemical conditions affect the

53

behavior of GO from a microscopic perspective. Furthermore, carboxyl group was found to

54

play an important role in determining the properties of GO in the field of materials synthesis,

55

and limited published research has been conducted to evaluate the contributions of carboxyl

56

groups on GO aqueous stability.11 GO sheet is essentially a network of hydrophobic benzene

57

rings, bearing hydroxyl, epoxy and carboxyl groups on the basal plane and sheet edge.12 4

ACS Paragon Plus Environment

Page 5 of 27

Environmental Science & Technology

58

Therefore, evaluating the effects of different groups is important in investigating the behavior

59

of GO.

60

Atomic-scale investigations with molecular dynamics (MD) simulations could contribute

61

significantly to understanding microscopic processes and furnish many details that are not

62

accessible experimentally.13-16 By utilizing MD, the microscopic aggregation of GO was

63

found to be a point-line-plane process, and dominant forces including van der Waals (vdWs)

64

attraction, electrostatic interaction, hydrogen bonds (H-bonds) interaction, π-π stacking, and

65

hydrophobic interaction were found to be different in respective stages.17 MD simulations

66

further suggested that the aggregates exhibit a GO-water-GO structure,18 and intra-layer and

67

inter-layer configurations of H-bonds forming between GO and water were put forward.19

68

However, these simulations were all carried out in pure water without containing ions or

69

NOM, which are ubiquitous in real natural water.

70

In addition to enhance the fundamental understanding of microscopic process, MD is also

71

essential to make the link between experiment and theory; for example, in exploring the

72

destructive extraction of phospholipids from Escherichia coli membranes by graphene

73

nanosheets, transmission electron microscopy was employed to show the rough stages of

74

extraction, and MD simulations to reveal the atomic details of the process.20 Gray et al.

75

studied the membrane fouling using Liquid Chromatography UV254, and the results were

76

supported by MD results.21 Liu et al. investigated uranyl and uranyl carbonate adsorption on

77

aluminosilicate surfaces by performing MD simulations and the structures of the adsorbed

78

complexes compared favorably with X-ray absorption spectroscopy results.13 Therefore, it is

79

reasonable and important to explore the aggregation of GO combining experiment 5

ACS Paragon Plus Environment

Environmental Science & Technology

80

Page 6 of 27

measurements and MD simulations.

81

With the aforementioned discussion, no studies have been conducted to uncover the

82

mechanisms behind the effects of solution chemistries and GO topologies on the aggregation

83

of GOs, and no xDLVO calculations of the aggregation of GO were available. Therefore, we

84

investigated the aggregation processes of GO using MD simulations and xDLVO calculations,

85

and the effect of GO topology was also evaluated.

86

MATERIALS AND METHODS

87

Preparation of GO and Carboxylated GO Suspensions

88

GO was synthesized through the reaction of graphite powder (Nanjing XFNANO Materials

89

Tech Co., Ltd) with KMnO4 in a concentrated H2SO4 solution (the Hummers method).22

90

NaOH and ClCH2COONa were used to convert GO to carboxylated GO (GO-COOH).23

91

Detailed preparing methods were shown in Supporting Information Part One (SI1).

92

Characterization of GO

93

Physical dimensions of GO and GO-COOH were measured by AFM. XPS and FTIR were

94

utilized to determine surface functionalities of GO and GO-COOH. The area under peaks was

95

used for quantifying the relative concentration of functional groups. The contents of different

96

functional groups were calculated from the XPS spectra.

97

Aggregation Kinetics of GO

98

Electrophoretic mobilities (EPM) and zeta potential were measured with a Zeta Sizer Nano

99

ZS (Malvern Instrument, Worcestershire, U.K.). Changes of GO and GO-COOH

100

hydrodynamic diameters (Dh) as a function of IS, pH, and the presence of NOM were

101

measured by DLS. Because of the special geometrical structure of GO (a one-atom-thick 6

ACS Paragon Plus Environment

Page 7 of 27

Environmental Science & Technology

102

sheet with a lateral size of micrometers), it appears that the DLS technique is not

103

quantitatively reliable. Lotya and co-authors showed the mean nanosheet length of GO is

104

proportional to Dh of GO measured by DLS.24 Therefore, the DLS analysis presented here

105

should be viewed only as a qualitative indicator to shed light on the aggregation of GO, and

106

DLS has been widely used in exploring the hydrodynamic properties of non-spherical

107

particles (including GO).7, 9, 18, 25-30

108

The GO concentration of 25 mg/L provided a strong DLS signal and was therefore used in

109

all aggregation studies. The initial aggregation period was defined as the time period from

110

experiment initiation (t=0) to the time when measured Dh values exceeded 1.50 Dh,initial. The

111

particle attachment efficiency (α) was used to quantify particle aggregation kinetics:α =

112

 () ( )   → ()  ( ) ,   →, 

113

concentrations (CCC) of GO and GO-COOH were determined from the intersection of

114

extrapolated lines through the diffusion and reaction limited regimes.

115

Setup of the Simulation System

116

Although multiple models for the structure of GO have been proposed, the most widely

117

accepted is Lerf−Klinowski model, which suggests epoxy and hydroxyl groups are located on

118

the basal plane (both sides), and carboxyl groups are attached to the carbon atoms on the

119

edge.31, 32 Oxidized groups distribute randomly on either side of the basal plane.31 In order to

120

quantify the oxygen content and proportion of these oxygen-containing functional groups,

121

XPS characterization was employed. Based on our XPS characterization results (Table S3),

122

the model of GO (Figure 1(a)) was set to be C20O2(OH)2(COOH)1, and similar models were

123

employed by other researchers.33-35 The structure and chemical composition of tannic acid

, where N0 is the initial particle concentration.26 Critical coagulation

7

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 27

124

(TA) is known and has been used as a NOM surrogate in the literature,36 therefore, TA was

125

employed to represent NOM in our studies. The model of TA is shown in SI2. Initially, two

126

GO sheets were well separated (Figure 1 (b) and (c)), and the initial distance between the

127

geometric center of each adjacent molecule was 2 nm. The GOs were then solvated in a cubic

128

periodic box with the distance between the solutes and box boundary at least 10 Å. To study

129

the effects of solution chemistries and different functional groups on the aggregation of GO,

130

different simulation systems were employed and the details were provided in SI3. To make

131

the simulation results more convincing, a statistical approach was employed: four sets of MD

132

simulations were performed with different GO sizes, system sizes, water models, GO models,

133

and GO layers. Details of each set of MD simulations are shown in SI3, and the detailed

134

simulation methods are provided in SI4.

135 136

Figure 1. Setup of the simulation system. (a) The model of GO. (b) System setup for the aggregation of

137

GO with the presence of metal ions. (c) System setup for the aggregation of GO with the presence of

138

metal ions and tannic acids. H in white, O in red, C in black, the blue balls represent metal ions, the

139

tiny red lines represent water molecules, the two sheers represent GOs, and the three smaller molecules

140

represent tannic acids. The initial distance between each adjacent molecule was 2 nm.

141

XDLVO Theory

142

According to the XDLVO theory,37 the aggregation of GO (plate-plate system) is determined

143

by electrostatic double layer (EDL) interactions, vdWs interactions, and Lewis acid-base (AB) 8

ACS Paragon Plus Environment

Page 9 of 27

Environmental Science & Technology

144

interactions, which can be calculated using the following equations: 38, 39

145

ΦEDL = ε ε ψ 2 [ () + 1 - coth (kh)]

146

ΦvdW = - 

147

ΦAB = ∆G!" exp ( 

148

The total interaction energy (Φ) between GOs: Φ = ΦEDL + ΦvdW + ΦAB

149

Permittivity of free space (ε0) and the dielectric constant of water (ε) were 8.854×10-12 C/V•

150

m and 78.5;7 zeta potential value was used in the place of surface potential ψ; h is the

151

separation distance between GOs. λ is the characteristic decay length of AB interactions in

152

water, whose value is between 0.2 and 1.0 nm, and a commonly used value of λ for aqueous

153

systems is 0.6 nm. 38, 39 h0 represents the minimum equilibrium cut-off distance and is usually

154

assigned a value of 0.157nm.40, 41 k (nm-1) is the inverse of Debye length and was calculated

155

using the following equation: 39

156

  # = &*

157

where kB is Boltzmann’s constant (1.38064852×10−23 J/K) 42, T is absolute temperature (300K

158

was employed in our study), NA is the Avogadro number (6.022140857×1023 mol−1)

159

ionic strength (mol/L) and e is the electron charge (1.6021766208×10−19 C) 44. Based on these

160

values, (5-1) can be transformed to:

161

 ≈ 3.28√I

162

A represents the Hamaker constant and was calculated using:45

163

 &1 45 45 2 A=24 π ℎ ( 23 − &15 )

164

∆G!" is the acid-base free energy per unit area and was calculated using: 45 

165

9 9 # ∆G!" = 2 [7189 (27123 − 718# ) +718#(27123 − 7189 ) −2 7γ# ;< γ;< ] 

166

To obtain the value of A andΔGAB, the contact angles (θ) of three probing liquids (water,

167

glycerol, and formamide) (3 µL) were acquired with a contact angle goniometer (Kino

168

SL200B) using thin GO films produced by drying concentrated GO suspensions on clean

169

glass slides. As the surface interfacial tension parameters of the selected probing liquids were





(1) (2)

 # ) $

(3)

' '( )  +, -

(4)

(5-1)

43

, I is

(5)

(6)

(7)

9

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 27

170

9 45 # known, the vdWs (123 ), electron-accepting (123 ), and electron-donating (123 ) interfacial

171

tension values for GO could be determined through the following equations:46

172

9 45 # 1=4 (1+cosθ) = 2&1=45 123 + 271=9 123 + 271=# 123

173

where i represent water, glycerol, and formamide, and the values of 1=∗ are shown in TableS4.

174

RESULTS AND DISCUSSION

175

Characterization of GO and GO-COOH Sheets

176

Physical dimensions of GO and GO-COOH measured by AFM are shown in Figure S4. The

177

GO and GO-COOH sheets are single-layered in the absence of ions. The average thickness of

178

GO and GO-COOH are determined to be 0.70 ± 0.14 nm, consistent with typical GO

179

samples.47 FT-IR and XPS results are summarized in SI5.

180

Electrokinetic and Hydrodynamic Properties under Different pH Regimes

181

As shown in Figure S6 (a), the EPMs of GO remained negative over the pH range from 0.3 to

182

12. As pH increased, the EPMs decreased sharply from about -(0.52±0.14)×10−8m2V−1s−1 to

183

-(3.5±0.01)×10−8m2V−1s−1. The changes of EPM at different pH conditions were caused by the

184

dissociable functional groups on GO.27 GO Dh measured after 15 minutes at different pH

185

levels are presented in Figure S6(b). The response of Dh to varying pH was similar to that of

186

EPM, with GO Dh being constant (∼150 nm) from pH 12 to 4, and then increasing sharply as

187

pH decreased. The increasing stability at higher pH was also observed for other

188

oxygen-containing functional group rich nanomaterials, and was attributed to a larger

189

electrostatic repulsive force as inidcated by EPM measuremenrs.48 However, EPM

190

measurements alone may have limitations in describing particle stability7,

191

simulations and XDLVO calculations were performed to further interpret the effect of pH on

(8)

49

, and MD

10

ACS Paragon Plus Environment

Page 11 of 27

Environmental Science & Technology

192

the behavior of GO.

193

The Aggregation of GO under Different pH Regimes

195

of carboxylic group, and 9.8 corresponding to the ionization of the hydroxyl group.50 The

196

range of pH usually observed in aquatic environment is from 5 to 9 7, indicating only

197

carboxylic groups in GO are deprotonated, while the hydroxyl groups are protonated.

198

Therefore, two GO models were employed: a) 90% of the carboxylic groups were protonated

199

(C20O2(OH)2(COOH)0.9(COO-)0.1) to simulate the GO at lower pH values, b) 90% of the

200

carboxylic groups were deprotonated (C20O2(OH)2(COOH)0.1(COO-)0.9) to simulate the GO at

201

higher pH values. (a)

6

0 3

distance 0

202

1

L-J Potential

10000 Time (ps)

-1 20000

4

(b) 0

3 -80 2 0

distance L-J potential 10000 20000 Time (ps)

(c) Amount of H-bond

Distance (nm)

9

Potential Energy (kJ/mol)

The pKa values of GO are 4.3, 6.6, and 9.8, with 4.3 and 6.6 corresponding to the ionization

Potential Energy (kJ/mol) Distance (nm)

194

400

Higher pH

320

0

Lower pH

10000 Time (ps)

20000

203

Figure 2. Molecular dynamics simulation results. (a) Distance and L-J potential energy between GOs

204

at higher pH values. (b) Distance and L-J potential energy between GOs at lower pH values. (c)

205

Amount of H-bonds forming between GOs and water at lower or higher pH values. The initial distance

206

between GOs was 2.0 nm. "Distance" here means the distance between the geometric centers of GOs.

207

Negative potential value implies attraction between GOs, and positive value implies repulsion. Blue

208

and purple dotted lines are used to distinguish the early aggregation stage and the aggregation stage.

209

At higher pH (Figure 2 (a) and video S1), the distance between GOs kept increasing and no

210

aggregation was observed during the simulation of 20 ns. As the distance was beyond the

211

scope (1.0nm) of vdWs interaction throughout the aggregation process, L-J potential energy

212

kept constant at 0 kJ / mol. 11

ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 27

213

At lower pH (video S2), the equilibrium state was obtained within about 5000 ps (Figure 2

214

(b)). Two distinctive representative stages for the aggregation are identified (although the

215

detailed time scales may different in different simulations, the overall trend is consistent in all

216

of them). (1) early aggregation stage (~0-2450 ps, blue dotted line): L-J potential energy and

217

electrostatic potential energy kept constant at 0 kJ/mol, with GO diffused freely in the solvent

218

and adjusted their motion direction. (2) aggregation stage (~2450 ps-4700ps, purple dotted

219

line): L-J potential energy kept negative and decreased to the minimum value, with GOs

220

aggregated quickly. It can be speculated that whether GOs can aggregate was governed by the

221

first stage. In this stage, GOs may separate from each other or approach to each other. If GOs

222

kept separating from each other, they would disperse in water and had no opportunity to

223

aggregate. If GOs approached one another at the end of this stage, then the subsequent

224

aggregation stage started. The initial distance between geometric centers of GOs was 2 nm,

225

which was beyond the scope of short-range forces (cut off for vdWs:1 nm, H-bond interaction:

226

0.35 nm

227

approaching was hydrophobic interaction53. Therefore, hydrophobic interaction played

228

dominant roles in the aggregation, which controls both the kinetics and thermodynamics.

51

, and π-π stacking: 0.35 nm

52

), indicating the driving force that accelerated the

229

GO is amphiphilic54, and the hydrophilicity of GO can be measured by the ability to form

230

H-bonds with water. As shown in Figure 2(c), at higher pH, more H-bonds formed between

231

GO and water, causing GO to favor the aqueous medium and decreasing the aggregation

232

efficiency. The acid-base interfacial free energy △ G!" is also a measure of 

233

hydrophilicity/hydrophobicity; when △ G!" >0, the material immersed in water is 

234

considered hydrophilic, and when △ G!"