Light-Induced Efficient Molecular Oxygen Activation on a Cu(II

Jan 3, 2015 - Lianwei Wang , Jianxin Liu , Yawen Wang , Xiaochao Zhang ... Zhong-Ting Hu , Yen Nan Liang , Jun Zhao , Yingdan Zhang , En-Hua Yang ...
0 downloads 0 Views 828KB Size
Subscriber access provided by SELCUK UNIV

Article

Light-induced Efficient Molecular Oxygen Activation on Cu(II) Grafted TiO2/graphene Photocatalyst for Phenol Degradation Hui Zhang, Liang-Hong Guo, Dabin Wang, Lixia Zhao, and Bin Wan ACS Appl. Mater. Interfaces, Just Accepted Manuscript • Publication Date (Web): 03 Jan 2015 Downloaded from http://pubs.acs.org on January 4, 2015

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.

ACS Applied Materials & Interfaces 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

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

1

Light-induced Efficient Molecular Oxygen

2

Activation on Cu(II) Grafted TiO2/graphene

3

Photocatalyst for Phenol Degradation

4

Hui Zhang, Liang-Hong Guo*, Dabin Wang, Lixia Zhao, Bin Wan

5 6

State Key Laboratory of Environmental Chemistry and Eco-toxicology, Research Centre for

7

Eco-environmental Sciences, Chinese Academy of Sciences, 18 Shuangqing Road, P.O. Box

8

2871, Beijing 100085, China

9 10

ACS Paragon Plus Environment

1

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 28

11

ABSTRACT: An efficient photocatalytic process involves two closely-related steps: charge

12

separation and the subsequent surface redox reaction. Herein, a ternary hybrid photocatalytic

13

system was designed and fabricated by anchoring Cu(II) cluster onto TiO2/reduced graphene

14

oxide (RGO) composite. Microscopic and spectroscopic characterization revealed that both TiO2

15

nanoparticles and Cu(II) clusters were highly dispersed on graphene sheet with intimate

16

interfacial contact. Compared with pristine TiO2, TiO2/RGO/Cu(II) composite yielded an almost

17

3-fold enhancement in the photodegradation rate toward phenol degradation under UV

18

irradiation. ESR spectra and electrochemical measurements demonstrated that the improved

19

photocatalytic activity of this ternary system benefited from the synergetic effect between RGO

20

and Cu(II), which facilitates the interfacial charge transfer and simultaneously achieves in-situ

21

generation of H2O2 via two-electron reduction of O2. These results highlight the importance to

22

harmonize the charge separation and surface reaction process in achieving high photocatalytic

23

efficiency for practical application.

24

KEYWORDS: TiO2, photocatalysis, oxygen reduction, phenol degradation

25

ACS Paragon Plus Environment

2

Page 3 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

26

ACS Applied Materials & Interfaces

INTRODUCTION

27

Photo-induced oxidation/reduction reactions at semiconductor surface present a promising

28

approach to address the emerging environmental issues.1-3 In a typical photocatalytic reaction,

29

molecular oxygen is activated by photogenerated electrons to produce ·O2- or H2O2, while holes

30

are trapped by surface absorbed hydroxyls (OH - ) to generate ·OH.4 These reactive oxygen

31

species (ROS) are identified as one of the most active initiators of photocatalytic degradation of

32

organic contaminants. Previous studies demonstrate that the photocatalytic efficiency is critically

33

dependent upon light adsorption, electron/hole separation and surface chemical reactions.5

34

Unfortunately, optimizing all these processes in a single photocatalyst has remained a great

35

challenge. Although UV light excited wide-band-gap photocatalysts (e.g., TiO2) usually exhibit

36

more efficiency in initiating or accelerating photodegradation reaction because of the high redox

37

potential of photogenerated electron/hole in the conduction/valance band,6 the quantum

38

efficiency of molecular oxygen activation is often hindered due to the charge recombination in

39

photoexcited semiconductor and the lack of active sites on the catalyst surface.

40

To overcome these problems, heterostructured photocatalysts have been constructed to

41

improve the quantum efficiency of TiO2.7,8 For example, TiO2/carbon nanomaterial (e.g., C60,

42

carbon nanotube) hybrids greatly promoted the charge separation across their interface.9,10 Metal

43

nanoparticles (e.g., Au, Ag and Pt) were used to quickly capture electrons from photoexcited

44

TiO2.11,12 In recent years, to avoid the high cost of noble-metal nanoparticles, surface

45

modification of TiO2 with transition metal ions (such as Cu(II) and Fe(III)) has been employed to

46

promote multielectron reduction reaction of oxygen and subsequently enhance the photocatalytic

47

performance.13-17 However, the increased quantum efficiency of Cu(II)- or Fe(III)-grafted TiO2 is

48

still limited due to the weak charge separation ability in the hybrid system. Indeed, both charge

ACS Paragon Plus Environment

3

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 28

49

separation and surface catalytic reaction are equally important, and the slower one will be the

50

rate determining step for the photocatalytic reaction. The challenge in achieving a high-

51

performance photocatalyst lies in ensuring a harmony of charge separation and the subsequent

52

surface catalytic reaction for molecular oxygen activation, which requires deliberate control of

53

the interactions among the photocatalyst components in a systematic and rational fashion.

54

With its monolayer structure and superior electron mobility, graphene is highly desirable as a

55

flexible two-dimensional (2D) catalyst support.18-20 Recently, graphene based photocatalysts

56

have been demonstrated to manipulate the charge transfer across semiconductor-graphene

57

interface.21,22 Upon excitation of TiO2 by light, electrons are transferred into RGO, which then

58

participate in red-/ox- reaction or lead to their storage within the C-C network in the absence of

59

additional electron acceptor.23,24 These elaborate works provide a hint to interface the light

60

harvesting semoconductor and surface reaction sites onto an individual RGO sheet to accelerate

61

charge transfer process and surface catalytic reaction simultaneously. Herein, we report for the

62

first time anchoring Cu(II) on TiO2/RGO mat to achieve an efficient photocatalytic process for

63

phenol degradation. The charge transfer and molecular oxygen activation process in the system

64

were investigated to explore the possible reaction mechanism. Our work demonstrates a

65

promising way to simultaneously tune charge separation and molecular oxygen activation,

66

consequently the overall photocatalytic performance.

67

EXPERIMENTAL SECTION

68

Chemicals and materials. Graphite powder was purchased from Bodi Chemical Co. Ltd

69

(Tianjin, China). Ti(OCH4)4, phenol and 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) were

70

purchased from Sigma-Aldrich (St.Louis, MO, USA). HPLC-grade methanol was obtained from

71

Fisher Scientific (Pittsburgh, PA, USA). All other chemicals were of analytical grade and used as

ACS Paragon Plus Environment

4

Page 5 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

72

received. All solutions were prepared with deionized (DI) water obtained using a Millipore Milli-

73

Q water purification system (Bedford, MA, USA).

74

Synthesis of Photocatalysts. Graphene oxide (GO) was synthesized by oxidation-exfoliation

75

of graphite powder according to a modified Hummers’ method.25 TiO2/RGO composites were

76

prepared by a hydrothermal method. Typically, GO (3.3 mg) was dispersed in a mixed solution

77

containing ethanol (15 ml) and water (30 ml) by ultrasonic treatment. Then, a solution of

78

Ti(OC4H9)4 (2.9 mL) and ethanol (15mL) was added dropwise to the GO suspension under

79

magnetic stirring. After stirring for 2 h, a homogeneous suspension with a light gray color was

80

obtained and then transferred into a 100 mL Teflon-lined autoclave and maintained at 180 0C for

81

10 h. The resultant TiO2/RGO composite was collected by centrifugation, washed repeatedly by

82

water and ethanol, and dried under vacuum at room temperature. Blank TiO2 sample was

83

prepared following a similar method without the addition of GO.

84

The grafting of Cu(II) ions onto TiO2/RGO composites was performed using an impregnation

85

method. In a typical preparation, TiO2/RGO sample (0.5 g) was dispersed into CuCl2 solution

86

(20 mL). The weight fraction of Cu2+ ions relative to TiO2/RGO was set to be 1%. Then, the

87

suspension was heated at 90 0C for 1 h using a water bath under stirring. The resultant

88

suspension was collected by centrifugation, washed by distilled water, and dried at 90 0C for 24

89

h. The final products were ground into a powder using an agate mortar. TiO2/Cu(II) sample was

90

prepared following the same method.

91

Characterizations. Morphology of the synthesized photocatalysts was investigated by

92

transmission electron microscopy (TEM) using JEOL JEM 2010 instrument (Tokyo, Japan)

93

operated at 120 kV. The elemental mapping images were acquired using a scanning TEM

94

attachment by equipping with an Oxford energy-dispersive X-ray detector (EDX) (Bucks, U.K.).

ACS Paragon Plus Environment

5

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 28

95

For TEM measurements, a droplet of the sample-ethanol solution was put onto a Mo grid.

96

Structural characteration of the prepared samples was performed by powder X-ray diffraction

97

(XRD) on a X′Pert Pro MPD (Eindhoven, Netherlands) with Cu Kα radiation (λ = 0.15418 nm)

98

and Raman spectra on a Renishaw InVia Raman spectrometer (Wotton-under-Edge, UK) with

99

exciting wavelength at 532 nm. The ionic characterstics were investigated by X-ray

100

photoelectron spectroscopy (XPS) on a Thermo VG ESCALAB 250 spectrometer (East Grinsted,

101

UK) with Al Kα radiation at 1486.6 eV. Elemental analyses of the samples were performed using

102

an inductively coupled plasma/optical emission spectrometer (ICP-OES) on a Perkin-Elmer

103

Optima 8300 model instrument (Norwalk, CT, USA). Brunauer−Emmett−Teller (BET) surface

104

area was obtained at 77 K on a QuadraSorb SI analyzer (Quantachrome, USA). Fluorescence

105

spectra were measured on a Horiba Fluoromax-4 spectrofluorimeter (Edison, NJ, USA). Electron

106

spin resonance (ESR) signal of the paramagnetic radicals trapped by 5,5-dimethyl-1-pyrroline-N-

107

oxide (DMPO) was recorded on a Bruker ER073 spectrometer (Karlsruhe, Germany) during

108

irradiation of the suspension (catalyst samples, 0.05mg/mL; DMPO, 100 mM) with a 500 W

109

mercury lamp. The settings were the following: center field 3503.95 G, microwave frequency

110

9.84 GHz and power 20 mW. Electrochemical experiments were performed in a three-electrode

111

cell on CHI 630B workstation (Shanghai, China) with a platinum plate as the counter electrode, a

112

saturated KCl Ag/AgCl electrode as the reference electrode. The working electrodes were

113

prepared by coating photocatalyst-ethanol slurry onto FTO glass and a small amount of Nafion

114

solution (0.5%) was added to paste the slurry. Current-potential curves and cyclic

115

voltammograms of the prepared electrode were measured in a 0.1 M NaClO4 electrolyte solution.

116

High-purity O2 gas was used to bubble the electrolyte during the electrochemical O2 reduction

117

experiments. The normal hydrogen electrode (NHE) potential was converted from the Ag/AgCl

ACS Paragon Plus Environment

6

Page 7 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

118

electrode using: ENHE = E(Ag/AgCl) + 0.197 V. Electrochemical impedance spectra (EIS) were

119

measured in 0.1 M KCl solution containing 2.5 mM K3[Fe(CN)6]/K4[Fe(CN)6] (1:1) by applying

120

5 mV alternative signal versus the reference electrode over the frequency range of 1MHz to 100

121

mHz.

122

Evaluation of Photocatalytic Properties. Photodegradation of a phenol solution in ambient

123

condition was performed to evaluate the photocatalytic activity of the prepared catalysts.

124

Typically, an aqueous solution containing phenol (50 mL, 10 mg/L) and a photocatalyst (50 mg)

125

was placed in a cylindrical Pyrex glass vessel and kept in dark for 30 min to ensure

126

adsorption/desorption equilibrium. Then the photoreaction vessel was irradiated by a 500 W

127

tubelike high-pressure mercury lamp with an irradiation intensity of 1.0 mW/cm2 and the main

128

emission wavelength at 365 nm. The suspension was continuously stirred by a Teflon-coated

129

magnetic stirrer bar during the irradiation. The temperature of the reaction solution was

130

maintained at 20 oC by recirculating a cooling water system during the reaction. At given time

131

intervals, aliquots of the irradiated solution were collected, centrifuged and analyzed by high-

132

performance liquid chromatograph (HPLC) on Aligent 1260 (Palo Alto, CA, USA) with a

133

Poroshell 120 EC-C18 column. The mobile phase was methanol and water (v:v = 0.55:0.45) at a

134

flow rate of 0.3 mL min-1, and the detection wavelength was set at 280 nm.

135

H2O2 generation during UV irradiation of the photocatalyst solution was determined by

136

dimerization of p-hydrophenylacetic acid (POPHA) in the presence of H2O2 and horseradish

137

peroxidase to yield a fluorescence product 5,5′-dicarboxymethyl-2,2-dihydroxybiphenyl (λex =

138

315 nm, λem = 406 nm).26 Aliquots (0.5 mL) of irradiated catalyst solution was taken out and

139

kept in dark for 30 min. Subsequently, 0.5 mL of the fluorometric reagent (8 mg of POPHA and

140

2 mg of horseradish peroxidase dissolved in 50 mL Tris buffer (0.1 M, pH 8.8) solution) was

ACS Paragon Plus Environment

7

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 28

141

added and allowed to react for 30 min. Then, the above solution was centrifugated and the

142

supernatant was taken for fluorescence measurement. The calibration curves were obtained by

143

standard additions of H2O2 to fluorometric reagents.

144

RESULS AND DISCUSSION

145

Characterization of TiO2/RGO/Cu(II) composite. Cu(II) grafted TiO2/RGO composite was

146

synthesized by a two-step process: (i) anchoring TiO2 nanoparticles (NPs) on RGO sheet as a 2-

147

D scaffold and (ii) adsorption of Cu(II) on TiO2/RGO. AFM image shows GO exists as

148

fragments with a thickness of ca. 1.1 nm (Figure S1 in Supporting Formation (SI)), which

149

confirms the structure of one or few layered GO sheet with oxygen-functional groups upon

150

chemical exfoliation. These oxygen groups on GO sheet would facilitate in situ growth of TiO2

151

nanocrystals.27 Meanwhile, the negative potential of GO given by these oxygen-functional

152

groups could induce Cu2+ adsorption through electrostatic interactions.28 Figure 1a shows typical

153

TEM images of the resulting TiO2/RGO/Cu(II) composites. It can be seen that spherically

154

structured TiO2 NPs are uniformly anchored on 2-D RGO sheet with a crystallite size of ca. 8 nm

155

(Figure S2). Both TiO2 NPs and the edge of RGO sheet are evident. Such decorated TiO2 NPs

156

can act as spacer to partially prevent exfoliated graphene sheet from stacking. High-resolution

157

TEM images (Figure 1b) exhibit a lattice fringes spacing of 0.35 nm, which can be assigned to

158

anatase (101) planes of TiO2, and a layered structure with interlayer spacing of 0.34 nm

159

corresponding to RGO.29 It needs to note that, compared with TiO2/RGO hybrids (Figure S3),

160

surface modification with Cu(II) does not change their morphology and size in the ternary

161

composite. Cu(II) on TiO2/RGO samples is hardly observed by TEM, probably due to the limited

162

amount of Cu(II) and an amorphous form at the low-grafting temperature.13 However, EDX

163

spectra confirm the presence of Cu(II). Figure 1c-e displays the representative element mapping

ACS Paragon Plus Environment

8

Page 9 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

164

images of Ti and Cu in the hybrids, demonstrating an homogeneous distribution of TiO2 NPs and

165

Cu(II) on the RGO sheet. Thus, an intimate interfacial contact among TiO2, RGO, and Cu(II)

166

components was achieved.

167

The structure character of the TiO2/RGO/Cu(II) composites was further determined by XRD

168

patterns (Figure S4). TiO2 in all the samples crystallize in an anatase phase structure (JCPDS No.

169

21-1272). Notably, a characteristic (002) peak of GO at 2θ = 7.76o (corresponding to an

170

interlayer distance of ~1.1 nm) disappeared completely after hydrothermal process, suggesting

171

the great reduction of GO and the successful intercalation of TiO2 NPs and Cu(II) to the GO

172

stacks in the final composites. No peaks associated with Cu-like compounds were detected,

173

implying Cu(II) exists as an amorphous form. Raman spectra provide addition insights into the

174

structure information of the composites (Figure 2). The Raman bands at 144 cm−1 (Eg), 399 cm−1

175

(B1g), 513 cm−1 (A1g) and 638 cm−1 (Eg) match well with TiO2 anatase structure.30 Two bands at

176

1326 cm-1 (D band) and 1597 cm-1 (G band) confirm the presence of RGO in the composites.31

177

Significantly, a characteristic 2D band (at ~ 2665 cm-1) corresponding to sp2 bounded carbon

178

atoms is also observed in the TiO2/RGO/Cu(II) hybrid. It is noted that the intensity ratio (I2D/IG)

179

usually reflects the recovery degree of sp2 hybridized graphitic structure.32 The increase of I2D/IG

180

ratio to 0.14 after the solvothermal process demonstrates an efficient restoration of sp2 carbon in

181

GO. This would favor the formation of 2-D catalyst assembly with high quality.

182

Further evidence for the interactions of components in the TiO2/RGO/Cu(II) composites come

183

from XPS spectra (Figure 3). A successful reduction of GO to RGO by solvothermal process is

184

verified by C 1s signal, according to the diminished C-O, C=O and O-C=O groups intensity. It is

185

worth noting that the remaining C-O group but with much lower intensity (blue dashed line) in

186

the composites provides the adsorption sites for Cu(II) anchoring. Moreover, a characteristic

ACS Paragon Plus Environment

9

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 28

187

peak at 288.7 eV can be assigned to Ti-O-C bond, suggesting TiO2 NPs interact with RGO sheet

188

through an esterification reaction.27,30 The successful loading of Cu(II) on the TiO2/RGO is

189

further demonstrated by Cu 2p3/2 core-level XPS signal at 923.6 eV, which is consistent with

190

previous results.13,14 By analogy with the reported Cu(II)-TiO2 system, Cu (II) is grafted on the

191

TiO2/RGO composites in a distorted, amorphous CuO structure with a five-coordinated square

192

pyramidal form. Although it is difficult to measure the location of Cu(II) in the composites, the

193

Cu(II) cluster may prefer to attach on RGO sheet due to the great complex ability of oxygen-

194

containing groups toward Cu2+ ions. ICP-OES measurements displayed that the loading amount

195

of Cu(II) was calculated to be ~0.92 wt% in the TiO2/RGO/Cu(II) composites, nearly equal to

196

the starting ratios used in the preparation. The cyclic voltammograms of the prepared

197

TiO2/RGO/Cu(II) electrode provide additional evidence for the loading of Cu (II) (Figure S5). A

198

couple of redox peaks were obviously observed, which are similar to the redox peaks of a Cu2+

199

solution on a bare TiO2 electrode. The reduction peak located at about 0.10 V is assigned to the

200

reduction of Cu(II) to Cu(I).33 In the reverse scan, a major oxidation peak at 0.24 V is detected

201

and attributed to Cu(I) oxidation. These results indicate Cu(II) clusters are successfully grafted

202

on TiO2/RGO composites.

203

Photocatalytic performance of TiO2/RGO/Cu(II) catalyst. Phenol, a typical toxic aromatic

204

compound with low degradability by conventional decomposition method, was selected as the

205

target molecule to evaluate the photocatalytic activity of the prepared composites. Figure 4a

206

shows the phenol degradation curves over various photocatalysts under UV light irradiation.

207

Both the direct photolysis of phenol and the adsorption of phenol on catalysts in the dark are

208

negligible, suggesting the observed phenol degradation is initiated by semiconductor

209

photocatalysis. Under UV light irradiation, the unmodified TiO2 sample shows a low

ACS Paragon Plus Environment

10

Page 11 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

210

photocatalytic activity. It is evident that the introduction of RGO and Cu(II) resulted in a

211

significant improvement in the photocatalytic phenol degradation. The temporal evolutions of

212

phenol concentration are well fitted with a pseudo-first-order rate equation. The rate constants (k)

213

for TiO2/Cu(II), TiO2/RGO and TiO2/RGO/Cu(II) composites are calculated to be 0.017 min-1,

214

0.014 min-1 and 0.023 min-1, respectively, which is up to almost three-fold enhancement than the

215

bare TiO2 (0.006 min-1) (Figure S6). Both TiO2/Cu(II) and TiO2/RGO exhibit a decent

216

photocatalytic activity, which might arise from efficient molecular oxygen activation by Cu(II)

217

clusters17 and enhanced charge separation by RGO,21,22 respectively. Importantly, the highest

218

photocatalytic activity of TiO2/RGO/Cu(II) clearly demonstrates Cu(II) and RGO act a

219

synergistic effect in boosting the photocatalytic performance of TiO2.

220

A separate experiment of TiO2/RGO/Cu(II) in the N2 purged suspension under UV light shows

221

a negligible phenol degradation, indicating that O2 participates in the photocatalytic degradation

222

reaction of organic pollutants. Once O2 is consumed by photogenerated electrons for phenol

223

degradation, it would be immediately supplied from air. Dissolved oxygen analysis displayed

224

that its concentration was maintained at 8.50 mg/L during the photocatalysis process. Indeed, as

225

the only molecular that can be reduced in ambient atmosphere, molecular O2 activation during

226

photocatalysis can not only inhibit the recombination of electron-hole pairs by capturing

227

electrons but also yield ·O2- or H2O2, which would participate in the subsequent photocatalytic

228

reaction. To further investigate the improved photocatalytic performance of TiO2/RGO/Cu(II)

229

composite, the generation of reactive oxygen species (ROS) during UV light irradiation was

230

probed. A DMPO spin-trapping ESR technique was employed to detect ·O2- and ·OH generation

231

(Figure S7). No signal was observed under the dark. Upon UV light irradiation, a characteristic

232

quartet peak of the DMPO-·OH adduct with intensity ratio of 1:2:2:1 was clearly observed.·O2-

ACS Paragon Plus Environment

11

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 28

233

was also captured to form DMPO-·O2- adduct with a quartet peaks in methanol suspension,

234

which usually accompany with DMPO-·CH2OH adduct signal due to the attack of ·OH radicals

235

onto the CH3OH solvent molecules.34 These results reveal that both ·O2- and ·OH were

236

generated from the irradiated photocatalyst samples.

237

To further assess ROS generation, temporal evolutions of ESR signal were displayed in Figure

238

5a and 5b, which were performed under the same conditions for comparison. It displays that the

239

rate of ·OH generation increases in the order of TiO2/RGO/Cu(II) > TiO2/RGO ≈ TiO2/Cu(II) >

240

TiO2, the trend of which is in good agreement with the observed phenol degradation. In general,

241

·OH radicals can generate from the photogenerated holes oxidation (h+ + OH- = ·OH), O2

242

reduction by electrons, and/or both. Figure 5b displays that, after Cu(II) modification, the rate of

243

·O2- generation over TiO2/RGO/Cu(II) is inhibited compared with that over TiO2/RGO. Indeed,

244

photogenerated electrons in the catalyst system could reduce O2 to ·O2- or H2O2 via one- or two-

245

electron reduction process, respectively (O2 + e− =·O2-, −0.046 V vs NHE; O2 + 2H+ + 2e− =

246

H2O2, 0.682 V vs NHE).35 Thus, a POPHA fluorescence method was employed to determine

247

H2O2 generation amount from different photocatalysts and further check the pathway of

248

molecular oxygen activation (Figure 5c). It was found that significantly more H2O2 were

249

generated over TiO2/RGO/Cu(II) than TiO2/RGO, which is contrary to the trend of ·O2-

250

generation. The rapid formation of H2O2 and slow growth of ·O2- indicate that one two-electron

251

reduction pathway (O2 → H2O2) is governed over TiO2/RGO/Cu(II), while an one-electron

252

reduction process for ·O2- generation is the main pathway over TiO2/RGO. Although the

253

generated ·O2- can further convert into H2O2 by the subsequent reaction, the amount of generated

254

H2O2 became steady after 40 min irradiation, indicating H2O2 was produced independent of ·O2-.

255

This result further suggests that H2O2 is generated via a two-electron reduction of O2 under the

ACS Paragon Plus Environment

12

Page 13 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

256

catalytic action of the grafted Cu(II) cluster. Indeed, considering the work function of RGO (-

257

4.42 eV) and the Fermi level of RGO at about 0 V (vs NHE), which is lower than that of

258

Cu2+/Cu+ (0.16 V vs NHE),28,36 Cu2+ clusters could notably capture the accumulated electrons on

259

RGO sheet. In addition, the potential of Cu2+/Cu+ is more positive than the one-electron

260

reduction of O2 →·O2-. Thus, a multielectron reduction pathway for H2O2 is more favored over

261

TiO2/RGO/Cu(II). This result is consistent with the multielectron reduction of O2 over the

262

reported Cu(II)/TiO2, Fe(III)/TiO2, and Rh(III)/TiO2 system.15,16,37 BET surface area analysis

263

displayed that the introduction of RGO and Cu(II) didn’t alter the surface area of TiO2 (Table

264

S1), indicating that the enhanced photoreactivity of the TiO2/RGO/Cu(II) was not ascribed to the

265

difference of surface area. Thus, it can be concluded that the grafted Cu(II) clusters on

266

TiO2/RGO work as efficient reactive sites to transfer the photogenerated electrons to O2

267

molecule, resulting in a highly efficient multielectron reduction of O2 to H2O2. The greatly

268

enhanced ROS generation demonstrate the synergic effect of RGO and Cu(II) in promoting

269

charge carrier separation.

270

Based on the above results, a schematic illustration of the photocatalytic reaction mechanism

271

over TiO2/RGO/Cu(II) is shown in Figure 6. RGO was demonstrated to store and shuttle

272

electrons to the surface adsorbed species.23 Upon UV light irradiation, a stepwise transfer of

273

electrons from photoexcited TiO2 to Cu(II) cluster via RGO (which acts as a “highway” for

274

electron shuttling) occurs and subsequently achieves O2 multielectron reduction. This process

275

makes phtogenerated holes and electrons to carry out selective catalytic reaction at separate sites.

276

Owing to the efficient consumption of electrons, more remaining holes participate in the

277

subsequent photocatalytic degradation reaction. Since Cu+/Cu2+ redox system provides only one

278

electron, twice amount of Cu species are required to achieve two-electron reduction of oxygen

ACS Paragon Plus Environment

13

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 28

279

molecules (O2 + 2Cu+ + 2H+ = H2O2 + 2Cu2+). This structure has been confirmed by the

280

aforementioned XPS results. The generated H2O2 could be further converted into ·OH radicals

281

through a Fenton-like reaction (Cu+ +H2O2 = Cu2+ + ·OH + OH-)28 or photocatalytic reactions

282

(H2O2 + UV light = 2·OH and H2O2 + e- =·OH + OH-).38 This indicates the synergetic effect

283

between RGO and Cu(II) not only suppresses the charge recombination by improving the

284

interfacial charge transfer, but also facilitates the surface catalytic reaction by providing reactive

285

centers. In addition, some photogenerated electrons can also be transferred directly to Cu(II) on

286

TiO2 surface or to RGO, and subsequently participate in photocatalytic reaction, which is

287

consistent with the decent photocatalytic activity over TiO2/Cu(II), TiO2/RGO composites.

288

The photocatalytic performance of the prepared composites under visible light irradiation (λ >

289

400 nm) was further investigated (Figure S8). The bare TiO2 showed rather poor photocatalytic

290

activity under visible light, while TiO2/Cu(II), TiO2/RGO and TiO2/RGO/Cu(II) composites

291

exhibited much improvements in the photocatalytic degradation of phenol. These results indicate

292

the narrowing of the band gap of TiO2 after Cu(II) and RGO incorporation, which can be

293

attributed the photoinduced interfacial charge transfer between TiO2 and the adsorbed Cu(II)

294

species13,14,36 and the formation of Ti-O-C band in the TiO2/RGO composites,39,40 respectively.

295

However, the photocatalytic efficiency of these composites under visible light is still low.

296

To evaluate the stability of the TiO2/RGO/Cu(II) photocatalyst, a recycling photocatalytic

297

experiment was carried out. As shown in Figure 4b, the high performance of TiO2/RGO/Cu(II)

298

was well maintained during the repeated light irradiation cycles. Moreover, the amount of Cu(II)

299

species in the five-cycle repeated TiO2/RGO/Cu(II) was about 0.87 wt%, which is comparable to

300

the as-prepared TiO2/RGO/Cu(II) composites. Few Cu species in the photocatalytic reaction

301

solution also indicated the good stability of Cu(II) in the composites (Figure S9). Thus, the high

ACS Paragon Plus Environment

14

Page 15 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

302

photocatalytic activity and stability of TiO2/RGO/Cu(II) composites suggest its great potential in

303

practical applications.

304

Electrochemical measurements of TiO2/RGO/Cu(II) catalyst. To further examine the

305

oxygen reduction process, electrochemical O2 reduction was investigated on the prepared

306

catalyst film electrode at bias potential to avoid the complicated redox process under light

307

irradiation conditions. Figure 7a shows the current-potential curves of the prepared composite

308

electrodes in Ar- or O2-saturated electrolyte. Little current was observed in Ar-bubbled solution

309

regardless of the electrodes, while in O2-saturated solution, a reduction current was observed,

310

indicating that O2 reduction occurred on the composite electrodes. Compared with the bare TiO2

311

electrode, a larger O2 reduction current is observed over TiO2/RGO and TiO2/Cu(II) electrodes

312

due to the important roles of RGO and Cu(II) in promoting charge transfer and facilitating O2

313

reduction, respectively. Under the synergetic effect of RGO and Cu(II), TiO2/RGO/Cu(II)

314

ternary system exhibited the highest O2 reduction current. Another interesting observation is that

315

the onset potential for oxygen reduction gradually shifted from below 0 V on TiO2 and

316

TiO2/RGO electrode, which corresponds to one-electron reduction of O2, to the positive region

317

(about 0.15 V) in the case of Cu(II) grafted electrode, in good agreement with the redox potential

318

of Cu2+/Cu+. At this potential, multielectron reduction of O2 probably occurred for H2O2

319

generation. Thus, the electrochemical oxygen reduction behavior of the electrodes further

320

demonstrate that the grafted Cu2+ accept electrons from TiO2 via RGO, and then transfer

321

electrons to O2 via multielectron reduction reaction, which is consistent with the observed ROS

322

generation during photocatalysis process.

323

The charge-carrier migration behavior in the TiO2/RGO/Cu(II) ternary system was further

324

characterized by EIS spectra (Figure 7b). Compared with the pure TiO2 sample, both TiO2/Cu(II)

ACS Paragon Plus Environment

15

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 28

325

and TiO2/RGO show depressed semicircles at high frequencies, suggesting a decrease in the

326

contact resistance at solid state interface layer and the charge transfer resistance at the electrode-

327

electrolyte solution contact interface.41 Notably, the TiO2/RGO/Cu(II) ternary system exhibited

328

the shortest circle among all three hybrid samples, indicating a faster interfacial charge transfer

329

and more efficient charge separation due to the excellent electron shuttling feature of RGO and

330

effective Cu(II) cluster reactive sites for electron capture. The efficient charge transport behavior

331

was further confirmed by the cyclic voltammograms (Figure S4). The anodic and cathodic

332

current density of Cu(II) on the TiO2/RGO/Cu(II) composite electrode is larger than that of TiO2

333

/Cu(II), implying enhanced electron transfer due to the introduction of RGO sheet as a 2-D

334

conducting substrate. These results demonstrate the significant role of RGO and Cu(II) in

335

improving charge separation and subsequent O2 activation for photocatalytic application.

336

Conclusion

337

In summary, Cu(II) grafted TiO2/reduced graphene oxide (RGO) composite was prepared for

338

developing high-performance photocatalyst. Under the synergetic effect between RGO and

339

Cu(II), this composition engineering not only optimizes charge transfer pathways for improved

340

charge separation, but also provide abundant photocatalytic reactive sites to reduce oxygen

341

effectively, resulting in enhanced photocatalytic performance for phenol degradation. In

342

addition, Cu(II) cluster modified TiO2/RGO preferentially reduce O2 to H2O2 via two-electron

343

transfer. The modulation of charge separation and molecule oxygen activation pathway described

344

here can be potentially applied to other photocatalysis system for environmental cleanup.

345 346

ASSOCIATED CONTENT

ACS Paragon Plus Environment

16

Page 17 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

347

Supporting Information. AFM image of GO, size distribution of TiO2 NPs in TiO2/RGO/Cu(II)

348

composites, TEM images of TiO2/RGO, XRD patterns, cyclic voltammograms, ESR spectra of

349

OH and·O2− species, BET surface area and visible light photoactivity of the prepared

350

photocatalysts, Cu-species concentration in the TiO2/RGO/Cu(II) composites suspension. This

351

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

352

AUTHOR INFORMATION

353

Corresponding Author

354

*Dr. Liang-Hong Guo, phone: 86 10-62849685, email: [email protected].

355

Notes

356

The authors declare no competing financial support.

357

ACKNOWLEDGMENT

358

This work was supported by the National Basic Research Program of China (2011CB936001)

359

and National Nature Science Foundation of China (21207146, 21177138, and 21477146).

360 361

ACS Paragon Plus Environment

17

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

362 363 364 365 366

Page 18 of 28

REFERENCES (1) Hoffmann, M. R.; Martin, S. T.; Choi, W.; Bahnemann, D. W. Environmental Applications of Semiconductor Photocatalysis. Chem. Rev. 1995, 95, 69–96. (2) Tachikawa, T.; Fujitsuka, M.; Majima, T. Mechanistic Insight into the TiO2 Photocatalytic Reactions: Design of New Photocatalysts. J. Phys. Chem. C 2007, 111, 5259–5275.

367

(3) Teoh, W. Y.; Scott, J. A.; Amal, R. Progress in Heterogeneous Photocatalysis: from

368

Classical Radical Chemistry to Engineering Nanomaterials and Solar Reactors. J. Phys. Chem.

369

Lett. 2012, 3, 629–639.

370

(4) Xu, W.; Jain, P. K.; Beberwyck, B. J.; Alivisatos, A. P. Probing Redox Photocatalysis of

371

Trapped Electrons and Holes on Single Sb-doped Titania Nanorod Surfaces. J. Am. Chem. Soc.

372

2012, 134, 3946−3949.

373 374

(5) Tong, H.; Ouyang, S.; Bi, Y.; Umezawa, N.; Oshikiri, M.; Ye, J. Nano-photocatalytic Materials: Possibilities and Challenges. Adv. Mater. 2012, 24, 229−251.

375

(6) Liu, M.; Qiu, X. Q.; Miyauchi, M.; Hashimoto, K. Energy-Level Matching of Fe(III) Ions

376

Grafted at Surface and Doped in Bulk for Efficient Visible-Light Photocatalysts. J. Am. Chem.

377

Soc. 2013, 135, 10064−10072.

378 379 380 381

(7) Kubacka, A.; Fernández-García, M.; Colón, G. Advanced Nanoarchitectures for Solar Photocatalytic Applications. Chem. Rev. 2012, 112, 1555−1614. (8) Yang, J.; Wang, D.; Han, H.; Li, C. Roles of Cocatalysts in Photocatalysis and Photoelectrocatalysis. Acc. Chem. Res. 2013, 46, 1900–1909.

ACS Paragon Plus Environment

18

Page 19 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

382 383 384 385

ACS Applied Materials & Interfaces

(9) Lin, J.; Zong, R.; Zhou, M.; Zhu, Y. Photoelectric Catalytic Degradation of Methylene Blue by C60-modified TiO2 Nanotube Array. Appl. Catal., B 2009, 89, 425–431. (10) Kongkanand, A.; Kamat, P. V. Electron Storage in Single Wall Carbon Nanotubes. Fermi Level Equilibration in Semiconductor−SWCNT Suspensions. ACS Nano 2007, 1, 13−21.

386

(11) Wang, F.; Jiang, Y.; Gautam, A.; Li, Y.; Amal, R. Exploring the Origin of Enhanced

387

Activity and Reaction Pathway for Photocatalytic H2 Production on Au/B-TiO2 Catalysts. ACS

388

Catal. 2014, 4, 1451–1457.

389

(12) Su, R.; Tiruvalam, R.; He, Q.; Dimitratos, N.; Kesavan, L.; Hammond, C.; Lopez-

390

Sanchez, J. A.; Bechstein, R.; Kiely, C. J.; Hutchings, G. J.; Besenbacher, F. Promotion of

391

Phenol Photodecomposition over TiO2 Using Au, Pd, and Au-Pd Nanoparticles. ACS Nano 2012,

392

6, 6284–6292.

393

(13) Irie, H.; Miura, S.; Kamiya, K.; Hashimoto, K. Efficient Visible Light-sensitive

394

Photocatalysts: Grafting Cu(II) Ions onto TiO2 and WO3 Photocatalysts. Chem. Phys. Lett. 2008,

395

457, 202–205.

396

(14) Irie, H.; Kamiya, K.; Shibanuma, T.; Miura, S.; Tryk, D. A.; Yokoyama, T.; Hashimoto,

397

K. Visible Light-Sensitive Cu(II)-Grafted TiO2 Photocatalysts: Activities and X-ray Aabsorption

398

Fine Structure Analyses. J. Phys. Chem. C 2009, 113, 10761−10766.

399

(15) Nishikawa, M.; Mitani, Y.; Nosaka, Y. Photocatalytic Reaction Mechanism of Fe(III)-

400

Grafted TiO2 Studied by Means of ESR Sspectroscopy and Cchemiluminescence Photometry. J.

401

Phys. Chem. C 2012, 116, 14900–14907.

ACS Paragon Plus Environment

19

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 28

402

(16) Nosaka, Y.; Takahashi, S.; Sakamoto, H.; Nosaka, A. Reaction Mechanism of Cu(II)-

403

Grafted Visible-Light Responsive TiO2 and WO3 Photocatalysts Studied by Means of ESR

404

Spectroscopy and Chemiluminescence Photometry. J. Phys. Chem. C 2011, 115, 21283–21290.

405

(17) Lin, M.; Inde, R.; Nishikawa, M.; Qiu, X.; Atarashi, D.; Sakai, E.; Nosaka, Y.;

406

Hashimoto, K.; Miyauchi, M. Enhanced Photoactivity with Nanocluster-Grafted Titanium

407

Dioxide Photocatalysts. ACS Nano 2014, 8, 7229–7238.

408

(18) Geim, A. K.; Novoselov, K. S. The Rise of Graphene. Nat. Mater. 2007, 6, 183–191.

409

(19) Bonaccorso, F.; Sun, Z.; Hasan, T.; Ferrari, A. C. Graphene Photonics and

410

Optoelectronics. Nat. Photonics 2010, 4, 611–622.

411

(20) Wei, Z.; Wang, D.; Kim, S.; Kim, S.-Y.; Hu, Y.; Yakes, M. K.; Laracuente, A. R.; Dai, Z.;

412

Marder, S. R.; Berger, C.; King, W. P.; de Heer, W. A.; Sheehan, P. E.; Riedo, E. Nanoscale

413

Tunable Reduction of Graphene Oxide for Graphene Electronics. Science 2010, 328, 1373–1376.

414

(21) Xiang, Q.; Yu, J.; Jaroniec, M. Graphene-Based Semiconductor Photocatalysts. Chem.

415 416 417

Soc. Rev. 2012, 41, 782–796. (22) Kamat, P. V. Graphene-Based Nanoassemblies for Energy Conversion. J. Phys. Chem. Lett. 2011, 2, 242–251.

418

(23) Krishnamurthy, S.; Kamat, P. V. Galvanic Exchange on Reduced Graphene Oxide:

419

Designing a Multifunctional Two-dimensional Catalyst Assembly. J. Phys. Chem. C 2013, 117,

420

571−577.

ACS Paragon Plus Environment

20

Page 21 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

421

(24) Xiang, Q.; Yu, J.; Jaroniec, M. Synergetic Effect of MoS2 and Graphene as Cocatalysts

422

for Enhanced Photocatalytic H2 Production Activity of TiO2 Nanoparticles. J. Am. Chem. Soc.

423

2012, 134, 6575–6578

424 425

(25) Hummers, W. S.; Offeman, R. E. Preparation of Graphitic Oxide. J. Am. Chem. Soc. 1958, 80, 1339.

426

(26) Zhao, K.; Zhang, L.; Wang, J.; Li, Q.; He, W.; Yin, J. J. Surface Structure-Dependent

427

Molecular Oxygen Activation of BiOCl Single-Crystalline Nanosheets. J. Am. Chem. Soc. 2013,

428

135, 15750−15753.

429

(27) Sun, J.; Zhang, H.; Guo, L.-H.; Zhao, L. Two-Dimensional Interface Engineering of a

430

Titania−Graphene Nanosheet Composite for Improved Photocatalytic Activity. ACS Appl. Mater.

431

Interfaces 2013, 5, 13035−13041.

432 433 434 435

(28) Xiong, Z.; Zhang, L. L.; Zhao, X. S. Visible-Light-Induced Dye Degradation over Copper-Modified Reduced Graphene Oxide. Chem. Eur. J. 2011, 17, 2428–2434. (29) Yeh, T. F.; Syu, J. M.; Cheng, C.; Chang, T. H.; Teng, H. S. Graphite Oxide as a Photocatalyst for Hydrogen Production from Water. Adv. Funct. Mater. 2010, 20, 2255−2262.

436

(30) Perera, S. D.; Mariano, R. G.; Vu, K.; Nour, N.; Seitz, O.; Chabal, Y.; Balkus, K. J.

437

Hydrothermal Synthesis of Graphene-TiO2 Nanotube Composites with Enhanced Photocatalytic

438

Activity. ACS Catal. 2012, 2, 949– 956.

439

(31) Rao, R.; Podila, R.; Tsuchikawa, R.; Katoch, J.; Tishler, D.; Rao, A.; Ishigami, M. Effects

440

of Layer Stacking on the Combination Raman Modes in Graphene. ACS Nano 2011, 5, 1594–

441

1599.

ACS Paragon Plus Environment

21

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 28

442

(32) Su, C. Y.; Xu, Y. P.; Zhang, W. J.; Zhao, J. W.; Liu, A. P.; Tang, X. H.; Tsai, C. H.;

443

Huang, Y. Z.; Li, L. J. Highly Efficient Restoration of Graphitic Structure in Graphene Oxide

444

Using Alcohol Vapors. ACS Nano 2010, 4, 5285–5292.

445

(33) Reyna-González, J. M.; Torriero, A. A. J.; Siriwardana, A. I.; Burgar, I. M.; Bond, A. M.

446

Extraction of Copper(II) Ions from Aqueous Solutions with a Methimazole-Based Ionic Liquid.

447

Anal. Chem. 2010, 82, 7691–7698.

448

(34) Wu, T.; Lin, T.; Zhao, J.; Hidaka, H.; Serpone, N. TiO2-Assisted Photodegradation of

449

Dyes. 9. Photooxidation of a Squarylium Cyanine Dye in Aqueous Dispersions under Visible

450

Light Irradiation. Environ. Sci. Technol. 1999, 33, 1379–1387.

451

(35) Wang, P.; Xia, Y.; Wu, P.; Wang, X.; Yu, H.; Yu, J. Cu(II) as a General Cocatalyst for

452

Improved Visible-Light Photocatalytic Performance of Photosensitive Ag-Based Compounds. J.

453

Phys. Chem. C 2014, 118, 8891−8898.

454

(36) Lightcap, I. V.; Kosel, T. H.; Kamat, P. V. Anchoring Semiconductor and Metal

455

Nanoparticles on a Two-Dimensional Catalyst Mat. Storing and Shuttling Electrons with

456

Reduced Graphene Oxide. Nano Lett. 2010, 10, 577–583.

457

(37) Kitano, S.; Murakami, N.; Ohno, T.; Mitani, Y.; Nosaka, Y.; Asakura, H.; Teramura, K.;

458

Tanaka, T.; Tada, H.; Hashimoto, K.; Kominami, H. Bifunctionality of Rh3+ Modifier on TiO2

459

and Working Mechanism of Rh3+/TiO2 Photocatalyst under Irradiation of Visible Light. J. Phys.

460

Chem. C 2013, 117, 11008–11016.

ACS Paragon Plus Environment

22

Page 23 of 28

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

461

(38) Li, X.; Zhen, X.; Meng, S.; Xian, J.; Shao, Y.; Fu, X.; Li, D. Structuring β-Ga2O3

462

Photonic Crystal Photocatalyst for Efficient Degradation of Organic Pollutants. Environ. Sci.

463

Technol. 2013, 47, 9911–9917.

464 465

(39) Zhang, H.; Lv, X.; Li, Y.; Wang, Y.; Li, J. P25-Graphene Composite as a High Performance Photocatalyst. ACS Nano 2010, 4, 380–386.

466

(40) Liu, B.; Huang, Y.; Wen, Y.; Du, L.; Zeng, W.; Shi, Y.; Zhang, F.; Zhu, G.; Xu, X.;

467

Wang, Y. Highly Dispersive {001} Facets-Exposed Nanocrystalline TiO2 on High Quality

468

Graphene as a High Performance Photocatalyst. J. Mater. Chem. 2012, 22, 7484–7491.

469

(41) Pan, X.; Zhao, Y.; Liu, S.; Korzeniewski, C. L.; Wang, S.; Fan, Z. Comparing Graphene-

470

TiO2 Nanowire and Graphene-TiO2 Nanoparticle Composite Photocatalysts. ACS Appl. Mater.

471

Interfaces 2012, 4, 3944–3950.

472

ACS Paragon Plus Environment

23

ACS Applied Materials & Interfaces

473 474

Figure 1. (a) TEM and (b) High-resolution TEM image of the prepared TiO2/RGO/Cu(II)

475

composite. (c-e) EDX element mapping images of Ti and Cu in the TiO2/RGO/Cu(II)

476

composites.

477

Intensity (a.u.)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 28

B1g Eg Eg A 1g

D

G

2D D+G GO TiO2/RGO/Cu(II) TiO2/RGO TiO2/Cu(II) TiO2

500

478

1000 1500 2000 2500 3000 -1 Raman shift (cm )

479

Figure 2. Raman spectra of GO, TiO2, TiO2/Cu(II), TiO2/RGO and TiO2/RGO/Cu(II)

480

composites.

ACS Paragon Plus Environment

24

Page 25 of 28

(a)

(b)

Intensity (a.u.)

C-O C=O Ti-O-C

TiO2/RGO/Cu(II)

Cu 2p3/2

Intensity (a.u.)

C-C

TiO2/RGO/Cu(II)

O=C-O

TiO2

GO

280

284 288 292 Bingding Energy (eV)

481

930

940 950 Bingding Energy (eV)

960

482

Figure 3. (a) Deconvoluted peaks of C 1s spectra of GO and TiO2/RGO/Cu(II) samples. (b) Cu

483

2p core-level spectra of TiO2 and TiO2/RGO/Cu(II) samples.

484 485 (b)

(a) 1.2 Photolysis

Dark Light On

1.0

5th

4th

0.8

TiO2/RGO/Cu(II)

C/C0

0.6 0.4

0.6 0.4 0.2

0.2

486

3rd

2nd

TiO2/Cu(II) TiO2/RGO

0.0 -30

1.0 1st

TiO2

0.8 C/C0

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

0

30

60

90 120 150 180

0.0

0

3

Time (min)

6

9

12

15

Time (h)

487

Figure 4. (a) Phenol degradation curves under UV light irradiation over TiO2, TiO2/Cu(II),

488

TiO2/RGO and TiO2/RGO/Cu(II) composites. (b) Cycling runs of TiO2/RGO/Cu(II) for

489

photocatalytic degradation of phenol.

490

ACS Paragon Plus Environment

25

TiO2/Cu(II)

0.12

TiO2/RGO TiO2/RGO/Cu(II)

0.09

0.5

Page 26 of 28

TiO2 TiO2/Cu(II) TiO2/RGO TiO2/RGO/Cu(II)

0.4

(b)

0.3 0.2

-

0.06 0.03 0.00 0

491

(a)

DMPO-·O2 Signal Intensity (a.u.)

TiO2

0.15

4 8 12 Irradiation Time (min)

H2O2 concentration (mM)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

DMPO-·OH Signal Intensity (a.u.)

ACS Applied Materials & Interfaces

16

0.0 0

4

8

12

16

Irradiation Time (min)

TiO2 TiO2/Cu(II) TiO2/RGO TiO2/RGO/Cu(II)

(c)

1.6

0.1

1.2 0.8 0.4 0.0 0

492

20

40

60 80 Time (min)

100

120

493

Figure 5. (a, b) The temporal evolution curves of DMPO-·OH adduct and DMPO-·O2- adduct

494

obtained from the first peak intensity. (c) The generation curves of H2O2 over different samples

495

under UV light irradiation.

496 497 498 499 500

ACS Paragon Plus Environment

26

Page 27 of 28

501 502

Figure 6. Schematic illustration of the structure of TiO2/RGO/Cu(II) composites (a) and the

503

possible photocatalytic process (b).

504

0.0 (a)

-2

-0.2

TiO2/Cu(II) (N2) TiO2/RGO (N2) TiO2/RGO/Cu(II) (N2)

-0.4

TiO2 (O2) TiO2/Cu(II) (O2) TiO2/RGO/Cu(II) (O2)

-0.2 0.0 0.2 Potential (V vs. NHE )

TiO2/Cu(II)

5000

TiO2/RGO

4000

TiO2/RGO/Cu(II)

3000 2000

TiO2/RGO (O2)

-0.6 -0.4

505

TiO2 (N2)

TiO2

(b)

6000 -Z′′ (ohm)

Current density (mA cm )

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

0.4

1000 0 0

1000

2000 3000 Z′ (ohm)

4000

5000

506

Figure 7. (a) Current-potential curves of the prepared electrodes in Ar or O2-saturated NaClO4

507

solutions. (b) EIS Nyquist plots of the the prepared electrodes in 0.1 M KCl solution containing

508

2.5 mM K3[Fe(CN)6]/K4[Fe(CN)6] (1:1).

509

ACS Paragon Plus Environment

27

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

510

Page 28 of 28

The Table of Contents

511

ACS Paragon Plus Environment

28