JES-00089; No of Pages 9 J O U RN A L OF E N V I RO N ME N TA L S CI EN CE S X X (2 0 1 4 ) XX X–XXX

Available online at www.sciencedirect.com

ScienceDirect www.journals.elsevier.com/journal-of-environmental-sciences

3Q3

Yaxin Zhang1,2 , Zeyu Zhou2 , Tan Chen2 , Hongtao Wang2,⁎, Wenjing Lu2

5

O

1. College of Environmental Science and Engineering, Hunan University, Changsha 410082, China. E-mail: [email protected] 2. School of Environment, Tsinghua University, Beijing 100084, China

R O

4

F

2

Graphene TiO2 nanocomposites with high photocatalytic activity for the degradation of sodium pentachlorophenol

1

6 8

AR TIC LE I NFO

ABSTR ACT

10

Article history:

A series of graphene–TiO2 photocatalysts was synthesized by doping TiO2 with graphene oxide 15

11

Received 15 November 2013

via hydrothermal treatment. The photocatalytic capability of the catalysts under ultraviolet 16

12

Revised 26 March 2014

irradiation was evaluated in terms of sodium pentachlorophenol (PCP-Na) decomposition and 17

13

Accepted 28 March 2014

mineralization. The structural and physicochemical properties of these nanocomposites were 18

D

P

9

characterized by X-ray diffraction, N2 adsorption–desorption, transmission electron micros- 19

E

14

copy, scanning electron microscopy, Ultraviolet–visible diffuse reflectance spectroscopy, X-ray 20 Keywords:

photoelectron spectroscopy, electron paramagnetic resonance spectra, and Fourier-transform 21

32

Photocatalysis

infrared spectroscopy. The graphene–TiO2 nanocomposites exhibited higher photocatalytic 22

33

Graphene

efficiency than commercial P25 for the degradation of PCP-Na, and 63.4% to 82.9% of the total 23

34

TiO2

organic carbon was fully mineralized. The improved photocatalytic activity may be attributed 24

35

PCP-Na

to the accelerated interfacial electron-transfer process and the significantly prolonged lifetime 25

36

Degradation

of electron‒hole pairs imparted by graphene sheets in the nanocomposites. However, 26

E

C

T

31

37

excessive graphene and the inhomogeneous aggregation of TiO2 nanoparticles may decrease 27 28

R

photodegradation efficiency.

R

© 2014 The Research Center for Eco-Environmental Sciences, Chinese Academy of Sciences. 29

39 40 38 41

Introduction

44

Heterogeneous nanocrystalline titania (TiO2) is commonly used because it is readily available, inexpensive, nontoxic, and relatively chemically stable. However, two problems limit the use of pure TiO2 in photocatalysis: the fast recombination of electron–hole pairs and the mismatch between band gap energy and solar radiation spectrum (Zhang et al., 2010b). The wide band gap causes selective ultraviolet (UV, λ < 380 nm) absorption by the catalyst (Khalid et al., 2013), and the short lifetime of photogenerated electron–hole pairs (2 to 3 μS) (Homa et al., 2009) causes the loss of approximately 90% of the generated carriers (Venkatachalam et al., 2007). To overcome

46 47 48 49 50 51 52 53 54

U

45

N C O

43 42

Published by Elsevier B.V. 30

these disadvantages, studies have focused on doping or co-doping of metal and non-metal elements into TiO2 to enhance quantum efficiency and extend the light-response range (Liu et al., 2011a). The recently discovered carbonaceous material graphene has elicited extensive attention in the fields of condensedmatter physics, electronics, and material science (Wang et al., 2013). Graphene shows various superior properties, such as high charge carrier mobility (>200,000 cm2/(V·sec), specific surface area (2630 m2/g), optical transmittance (~97.7%) (Zhang et al., 2012a), and Fermi velocity (106 m/sec) (Song et al., 2011). Thus, graphene–TiO2 (GT) nanocomposites have become a promising material for photocatalytic applications. GT has high stability,

⁎ Corresponding author. E-mail: [email protected] (Hongtao Wang).

http://dx.doi.org/10.1016/j.jes.2014.08.011 1001-0742/© 2014 The Research Center for Eco-Environmental Sciences, Chinese Academy of Sciences. Published by Elsevier B.V.

Please cite this article as: Zhang, Y., et al., Graphene TiO2 nanocomposites with high photocatalytic activity for the degradation of sodium pentachlorophenol, J. Environ. Sci. (2014), http://dx.doi.org/10.1016/j.jes.2014.08.011

55 56 57 58 59 60 61 62 63 64 65 66 67

2

77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98

1. Material and methods

101

1.1. Chemicals and reagents

102 Q4

109

PCP was purchased from Dima (USA). Commercial P25 TiO2 (80% anatase, 20% rutile; Brunauer–Emmett–Teller (BET) area, 54.77 m2/g) was supplied by Degussa Corporation (Germany). Graphene oxide (>99%) was purchased from Nanjing Xianfeng Nanomaterial Technology Co., Ltd. (China). All chemicals used in this study were of analytical grade and used without further purification. Double-distilled water was used in catalyst preparation and subsequent catalytic tests.

110

1.2. Catalysts preparation

111

Graphene oxide (100 mg) was sonicated in 250 mL of double-distilled water at room temperature for at least 1 hr to ensure thorough dispersion of graphene oxide in the solution. Then, a certain amount of P25 TiO2 was added to the above solution with magnetic stirring (120 r/min). The weight addition ratios of graphene oxide to P25 were 1:10, 1:20, 1:50, and 1:100. After stirring for 2 hr, the suspension was transferred into a 100 mL Teflon-lined autoclave and subjected to hydrothermal treatment at 180°C, for 6 hr. The hydrothermal process ensured sufficient reduction of graphene oxide to graphene (Yang et al., 2013). The obtained precipitates were centrifuged at 10,000 r/min for 10 min and then washed with double-distilled water until the

107 108

112 113 114 115 116 117 118 119 120 121 122

R

O

C

106 Q6

N

105

U

104 Q5

R

100 99

103

Crystal structures of samples were analyzed using X-ray diffraction (XRD) with a Bruker D8 Advance diffractometer with Cu Ka radiation (λ = 1.5418 Å) operating at 40 kV and 40 mA. BET specific surface area, Barrett–Joyner–Halenda (BJH) pore volume, average pore diameter, and fractal dimension of photocatalyst calculated from the N2 adsorption–desorption isotherms were obtained with a Quantachrome gas adsorption instrument (Autosorb AS-1N2). The samples were vacuum-degassed for 4 hr at 300°C before measurements. Ultraviolet–visible (UV–Vis) diffuse reflectance spectra (DRS) were obtained using a Hitachi U-3010 spectrophotometer equipped with an integrating sphere accessory for diffuse reflectance. Transmission electron micrographs were obtained with a Hitachi H-800 transmission electron microscope operating at an accelerating voltage of 200 kV. Scanning electron micrographs were acquired by a Hitachi S-5500 scanning electron microscope at an accelerating voltage of 20 kV. Surface composition of the nanocomposite was analyzed by X-ray photoelectron spectroscopy (XPS) on a Kratos Axis Ultra System with monochromatic Al Ka X-rays (1486.6 eV). Electron paramagnetic resonance spectra (EPR) were recorded at 70 K on a JEOL JES-FA200 spectrometer operating in the X-band with 30 mg of the sample introduced into an EPR quartz probe cell under N2 atmosphere. The FT-IR spectra of the catalysts were obtained using a spectrometer (Spectrum GX, PerkinElmer, USA) within the 4000 to 400 cm−1 region at a 4 cm−1 resolution. Approximately 2 mg of the milled sample was ground with 200 mg of KBr (FT-IR grade) and compressed into a pellet under vacuum at a pressure of 75 kN/cm for 3 min. A total of 32 scans were performed, averaged for each spectrum, and then corrected against air as background in each acquisition. The spectra were subjected to a Savitsky–Golay second derivative filter (using a third-order polynomial and seven-point smoothing) to discriminate the emerging peaks (Savitzky and Golay, 1964).

128

1.4. Photocatalytic test

162

The photocatalytic degradation experiments were carried out in a vigorously stirred batch reactor containing 500 mL of PCP-Na (50 mg/L) solution and 100 mg of catalyst (GT or P25). The reactor was equipped with a high-pressure mercury lamp (300 W), and the temperature was maintained at 25°C. The photocatalytic degradation experiments were performed in a simple targeted PCP-Na aqueous system. Before irradiation, the suspension was stirred in darkness for 30 min to obtain an adsorption–desorption equilibrium, which proved to be sufficient (Zhang et al., 2012b). At designated time intervals, 1 mL aliquots of the suspension were collected and transferred into centrifuge tubes (CNW, 8 mL) with Teflon-lined screw caps. To quantitatively recover the analytes adsorbed on the catalyst, an equal volume of methanol was added to the suspension sample, which was then thoroughly shaken. The processed suspension was filtered through a 0.45 μm pore size membrane

163

F

76

127

O

75

1.3. Characterization

R O

74

P

73

D

72

123

E

71

pH of the supernatant was neutral. The sediment samples were dried at 333 K for 24 hr in an electric oven and then ground in an agate mortar. The as-synthesized photocatalysts of different ratios were denoted as GT 1:10, 1:20, 1:50, and 1:100.

T

70

high photocatalytic activity, and enhanced photovoltaic properties (Long, 2013) because of its enhanced quantum efficiency and decreased band gap energy (Geng et al., 2013). Several studies have been focused on the photocatalytic activity of GT. However, most of these studies utilized dyes or simple substances as target pollutants, such as methylene blue (Zhang et al., 2010a), benzene (Zhang et al., 2010c), Rhodamine B (Li et al., 2013), propanol (Kamegawa et al., 2010), and Acid Orange (An and Yu, 2011), which cannot fully represent the toxic substances existing in the environment. In aqueous systems, the harmful effect of persistent organic pollutants (POPs) on human health is becoming a global concern. Pentachlorophenol (PCP), one of the common POPs, is widely used as a wood preservative, anticorrosive, fungicide, and pesticide, leaving serious contamination in the environment (Xie et al., 2013). PCP is highly toxic, mutagenic, and carcinogenic, and was listed as one of the priority pollutants in the US, European Union, and China (Liu et al., 2011a). PCP is resistant to biodegradation and photolytic degradation in aqueous systems because of its stable chemical structure; thus, investigating the photocatalytic degradation of PCP using GT as a catalyst is of great interest. In this study, GT nanocomposites of different carbon material ratios were prepared via a facile hydrothermal treatment. The function of graphene in the enhancement of light absorption, quantum efficiency, and photoactivity of the photocatalysts was systematically investigated. The photocatalytic activity of the as-prepared nanoparticles was examined by the photodegradation of sodium pentachlorophenol (PCP-Na) under UV irradiation and the mineralization of the target pollutant. This study provides a new approach for the decomposition of persistent hazardous compounds.

C

69

E

68

J O U RN A L OF E N V I RO N ME N TA L S CIE N CE S X X (2 0 1 4 ) XXX –XXX

Please cite this article as: Zhang, Y., et al., Graphene TiO2 nanocomposites with high photocatalytic activity for the degradation of sodium pentachlorophenol, J. Environ. Sci. (2014), http://dx.doi.org/10.1016/j.jes.2014.08.011

124 125 126

129 Q7 130 131 132 133 134 135 Q8 136 137 138 139 140 Q9 141 142 143 Q10 144 145 146 Q11 147 148 Q12 149 150 Q13 151 Q14 152 153 154 155 156 157 158 159 160 161

164 165 166 167 168 169 170 171 172 173 174 Q15 175 176 177 178

3

J O U RN A L OF E N V I RO N ME N TA L S CI EN CE S X X (2 0 1 4 ) XX X–XXX

179

filter. The concentrations of PCP-Na and surfactants were determined using high-performance liquid chromatography 181 Q16 (Agilent 1260) fitted with a UV detector and an Agilent TC-C18 182 column (4.5 mm × 250 mm; 5 μm). The mobile phase of high183 performance liquid chromatography was methanol–water 184 (80:20) at a flow rate of 1.00 mL/min. Chromatography was 185 performed at 35°C. UV wavelengths were set at 249 nm for 186 PCP-Na. During photodegradation, 10 mL of the sample was 187 collected at certain time intervals and filtered through a 0.45 μm 188 pore size membrane filter for total organic carbon testing 189 Q17 (Shimadzu TOC-VCPH). All data were expressed as the means 190 of two replicates with ±5% deviation.

2.2. BET surface areas and pore distributions

212

The specific surface areas of P25 and GT with different addition ratios were determined by the BET method using adsorption data, whereas the pore volume and pore size distribution were derived using the BJH model based on desorption isotherms. Fig. 2 (inset) shows the adsorption–desorption isotherms and the corresponding pore size distribution curves of the catalysts. All catalysts exhibited similar type IV isotherms (Sing et al., 1985) with an H3 hysteresis loop at a relative pressure close to unity, which are typical characteristics of mesoporous materials (Liu et al., 2011b). The pore size distributions (inset in Fig. 2) of the GT samples were broad (from 2 to over 100 nm), indicating the presence of mesopores and macropores (Xiang et al., 2011b). As shown in Table 1, the specific surface area and pore volume of the as-prepared catalysts were primarily dominated by graphene compared with those of P25. However, the pore diameter was uniform and lower than 5 nm. The BET surface area of the GT samples significantly increased from 53.74 to 72.73 m2/g with increasing graphene content. This result can be attributed to the fraction of bare graphene in the composites. However, the BET surface area of the GT samples was still relatively low compared with that of pure graphene (2600 m2/g). This result can be ascribed to the minimal doping level of graphene. Thus, surface area and porosity are mainly dominated by the TiO2 component. Another parameter that can provide additional information on the catalyst structure in terms of “roughness exponent” is the surface fractal dimension (D) (Zhu et al., 2009). Fractal dimension describes the topography of the real surface based on fractal geometry. This factor is usually between 2 (smooth surface) and 3 (rough surface). The complex pore and surface structure can be adequately described by D values (Zhang et al., 2012c). D was derived using the Frenkel– Halsey–Hill and Neimark–Kiselev methods (Table 1). The D values of the GT samples increased with increasing graphene content. This result suggests that the doping with graphene,

213

O

The XRD patterns of the as-prepared GT composites and P25 TiO2 are shown in Fig. 1. GT nanocomposites with different weight addition ratios of graphene and P25 clearly had similar XRD patterns. Peaks located at 25.3°, 37.8°, 48.0°, 55.1°, and 62.7° can be indexed to the (101), (004), (200), (211), and (204) crystal planes of anatase TiO2 (JCPDS No. 21-1272), respectively. No typical diffraction peaks of graphene were observed in the nanocomposites. This result can be attributed to the main characteristic peak of graphene at 25.0° overlapping with the (101) peak of anatase TiO2 (Zhang et al., 2011) and to the relatively low weight fraction of graphene in the nanocomposites. However, the GT composites slightly increased the amount of rutile phase TiO2 (JCPDS No. 21-1276). The XRD results indicate that doping graphene with TiO2 promotes the phase transformation from anatase to rutile. In other words, graphene has a catalytic function in the phase transformation from anatase to rutile during the hydrothermal treatment because of its special structural properties (Li et al., 2013).

C

E

R

250

0.010

Anatase Rutile

GT (1:100) GT (1:50) GT (1:20)

GT 1:100

GT 1:20 GT 1:50

P25 GT 1:10

0.008

200

150

dV/dW (cm3/(g.min))

Adsorbed volume (cm3/g, STP)

211

(224)

210

R

209

(215)

208

(204)

207

(116)

206

(211) (220)

205

N C O

204

(200)

203

(210)

202

(101) (004)

201

(111)

200

U

199

(101)

198

(110)

197

Intensity (a. u.)

196

R O

194

P

2.1. XRD patterns of the photocatalysts

D

193

E

2. Results and discussion

T

192 191

195

F

180

0.006

0.004

0.002

100 0.000

10

100 Pore diameter (nm)

50

GT (1:10) P25 20

30

40

50

60

70

80

90

100

2θ (°) Fig. 1 – XRD patterns of the graphene–TiO2 (GT) photocatalyst and P25.

0 0.0

0.2

0.4

0.6

0.8

1.0

Relative pressure (P/P0) Fig. 2 – Nitrogen adsorption–desorption isotherm and pore size distribution curve (inset) of graphene–TiO2 (GT) and P25.

Please cite this article as: Zhang, Y., et al., Graphene TiO2 nanocomposites with high photocatalytic activity for the degradation of sodium pentachlorophenol, J. Environ. Sci. (2014), http://dx.doi.org/10.1016/j.jes.2014.08.011

214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247

4

t1:3

t1:4

Table 1 – BET specific surface area, pore volume, average pore diameter, and fractal dimension of graphene–TiO2 nanocomposites and P25. Sample

BET surface area (m2/g)

Total pore volume (mL/g)

t1:6 t1:7 t1:8 t1:9 t1:10

P25 GT 1:10 GT 1:20 GT 1:50 GT 1:100

t1:11 t1:13 t1:12 t1:14 t1:15

GT: graphene–TiO2. D values calculated at P/P0 > 0.35. b NK: D was derived using Neimark–Kiselev method. c FHH: D was derived using Frenkel–Halsey–Hill method.

53.74 72.73 64.39 58.39 55.63

Fractal dimension (D) a

Pore diameter (nm)

0.264 0.361 0.372 0.338 0.332

3.092 3.109 3.117 3.113 3.115

NK b

R2

FHH c

R2

2.459 2.568 2.527 2.493 2.488

0.998 0.979 0.980 0.990 0.989

2.555 2.517 2.512 2.504 2.491

0.993 0.995 0.995 0.997 0.977

t1:5

a

250 251 252 253 254

2.3. Scanning electron microscopy and transmission electron microscopy The scanning electron microscopy (SEM) and transmission electron microscopy (TEM) images of GT 1:10 and 1:100 are shown in Fig. 3. As shown in the figure, the TiO2 nanoparticles did not easily complex with graphene and were entrapped

inside the graphene sheets. However, the TEM images of GT reveal inhomogeneous dispersion of TiO2 particles in the graphene matrix. The particles accumulated along the wrinkles and edges of the graphene sheets. This accumulation limited the interfacial contact between TiO2 and graphene, which resulted in insufficient utilization of the excellent electron conductivity of graphene (Zhang et al., 2011). Hence, further studies should focus on achieving homogenous dispersion of solid TiO2 particles on a graphene sheet to effectively utilize the “structure-directing” ability of graphene

a

T

E

255

with its wrinkles and stacking structure, forms composites with a “rougher” surface.

P

249

D

248

R O

O

F

t1:1 t1:2

J O U RN A L OF E N V I RO N ME N TA L S CIE N CE S X X (2 0 1 4 ) XXX –XXX

b

TEM-GT 1:100

U

c

N

C

O

R

R

E

C

TEM-GT 1:10

Q1

SEM GT 1:10

d

SEM GT 1:100

Fig. 3 – TEM and SEM images of graphene–TiO2 (GT). (a) TEM of GT 1:10, (b) TEM of GT 1:100; (c) SEM of GT 1:10; (d) SEM of GT 1:100. Please cite this article as: Zhang, Y., et al., Graphene TiO2 nanocomposites with high photocatalytic activity for the degradation of sodium pentachlorophenol, J. Environ. Sci. (2014), http://dx.doi.org/10.1016/j.jes.2014.08.011

256 257 258 259 260 261 262 263 264 265

5

J O U RN A L OF E N V I RO N ME N TA L S CI EN CE S X X (2 0 1 4 ) XX X–XXX

oxide, which possesses abundant oxygen-containing functional groups on the basal plane and edge that can provide reactive sites.

2400

269

2.4. UV–Vis DRS

1200

270

286

Fig. 4 shows the UV–Vis DRS of the as-prepared GT composites. For comparison, P25 TiO2 was also tested under the same condition. P25 showed a mild response within 200 to 400 nm, which can be attributed to the electron transition from the valence band to the conduction band. The addition of graphene increased the light absorption intensity in both UV and visible light regions. Furthermore, a qualitative red shift to higher wavelength was observed in the absorption edge of GT nanocomposites (50 nm for GT 1:100). This shift can be attributed to the hybridization of C2p and O2p atomic orbits to form a new valence band; hence, the band gap energy was reduced in the GT system (Yang et al., 2013). Within the 400 to 800 nm range, light absorption intensity increased with increasing graphene content. This phenomenon can be attributed to the surface plasmon resonance effect of the graphene unit (Li et al., 2013). These findings suggest that the introduction of graphene can promote the visible light response of the nanocomposites.

287

2.5. XPS

288

The XPS surface probe technique was adopted to study the chemical states of Ti, O, and C species and the interaction of graphene and TiO2 on the surface of GT composites. Fig. 5 shows the XPS of P25 and GT 1:100 composites in the Ti 2p, O 1s, and C 1s binding energy regions. For GT, the measured binding energies of Ti 2p3/2 and Ti 2p1/2 were 458.3 and 464.0 eV, respectively, consistent with P25 TiO2. The deconvoluted peaks centered at the binding energies of 529.5, 530.4, and 531.7 eV were assigned to Ti\O, O\H, and C\O functional groups, respectively. Fig. 5b shows the shift in the O 1s binding energies, illustrating that electron transformation at the interface between graphene and TiO2 perturbs the O electronic environment (Li et al., 2013).

282 283 284 285

289 290 291 292 293 294 295 296 297 298 299 300

Intensity (a. u.)

1.5

1.2

0.9

F 466

462

460

458

456

454

452

Binding energy (eV)

2200 1650

2750

O1s

C-O 531.7

O-H 530.4

1100

b

Ti-O 529.5

GT 1:100

2750

P25

2200 1650 1100 536

534

532

530

528

526

524

Binding energy (eV) 3500

Celement, C-H, C-C GT 1:100

2800

c

284.2

C1s

2100 1400

GT 1:10

464

O

0 468

R O

281

P25

2250

P

280

3000

D

279

Intensity (a. u.)

278

600

750

E

277

Intensity (a. u.)

276

Ti2p

1500

T

275

C

274

a

Ti 2p3/2 458.3 eV Ti 2p1/2 464.0 eV

E

273

R

272

GT 1:100

1800

R

271

N C O

268

700

C=O 288.3

C-OH 285.1

30000 P25

2250

GT 1:20

1500

U

267

Absorbence (a. u.)

266

GT 1:50 GT 1:100

0.6

750 0 292

290

288

286

284

282

280

278

276

Binding energy (eV) 0.3 P25

0.0 200

300

400

500

600

700

800

Fig. 5 – High-resolution XPS spectra of the synthesized graphene–TiO2 (GT 1:100) and P25. (a) Ti 2p region; (b) O 1s region; (c) C 1s region.

Wavelength (nm) Fig. 4 – UV–Vis diffuse reflectance spectra of the synthesized graphene–TiO2 (GT) and P25.

Fig. 5c shows the high-resolution C 1s XPS spectra of the 301 GT 1:100 sample in comparison with that of P25. The binding 302 energy of 284.2 eV can be attributed to the carbon, C\C, and 303

Please cite this article as: Zhang, Y., et al., Graphene TiO2 nanocomposites with high photocatalytic activity for the degradation of sodium pentachlorophenol, J. Environ. Sci. (2014), http://dx.doi.org/10.1016/j.jes.2014.08.011

6

2.6. EPR analysis

318

An in situ EPR study was performed for GT composites and P25 at liquid N2 temperature (77 K) under UV irradiation. Photo-induced EPR signals of trapped electrons and holes on GT nanocomposites and P25 are shown in Fig. 6. Two sets of features were observed in the EPR spectrum: signals from electrons (A) and holes (B). EPR signals due to GT were characterized by g values. The sharp signal at g┴ = 1.995 and g// = 1.970 corresponded to surface electron trapping sites and are represented as signal A (Gopal et al., 2008; Kumar et al., 2006). EPR signals with g factors g1 = 2.030, g2 = 2.016, and g3 = 2.008 corresponded to surface hole trapping sites and are represented as signal B (Ribbens et al., 2011; Shkrob et al., 2011). Fig. 6 demonstrates apparent variation in EPR intensities between trapped electrons and holes of GT and P25. For P25, a very weak signal was observed presumably due to hole traps. This result indicates that the lifetime of electron and hole radicals is too short to be observed under such testing conditions. By contrast, the EPR intensities for the GT nanocomposites were higher and more intense, especially for GT 1:20, followed by GT 1:100. These results suggest that graphene doping can effectively prolong the lifetime of photoelectronhole pairs.

323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338

2.8. Photocatalytic degradation of PCP-Na

357

The photocatalytic activity of GT with different carbon ratios and P25 catalysts was evaluated through PCP-Na (initial concentration of 50 mg/L) degradation and TOC variation in aqueous solution under UV irradiation for 2 hr (Fig. 8a, b, respectively). After 120 min of UV irradiation, 31.1% of PCP-Na was degraded using P25 as the catalyst. Remarkable improvements in photodegradation were observed in all GT nanocomposites, wherein more than 97% of PCP-Na was decomposed. However, GT with different graphene ratios showed similar degradation efficiencies. The degradation curves were fitted to the pseudo-firstorder kinetic law described by the following equation:

358

R O

322

P

321

D

320

T

319

C

315

E

314

R

313

  c − ln ¼ kt c0

ð1Þ

R

312

340

O

311

14000

P25 GT 1:100 GT 1:50 GT 1:20 GT 1:10

12000

C

310

10000

N

309

The FT-IR spectra of GT and P25 TiO2 are described in Fig. 7. The broad peak at 3400 cm−1 in all samples indicates the presence of surface O\H stretching vibrations (Nguyen-Phan et al., 2011). Peaks at around 1393 and 1250 cm−1 can be assigned to tertiary C\OH and epoxy C\O groups, respectively (Nguyen-Phan et al., 2011; Shen et al., 2011). The absorption peak at around 1600 cm−1 is contributed by the skeletal vibration of graphene (Zhang and Pan, 2011). For the GT composites, the C\O stretching vibration peak at 1052 cm−1 and the C_O stretching vibration of \COOH group peak at 1720 cm−1 of graphene oxide completely disappeared. Thus, the oxygen-containing functional groups were completely removed through the thermal reaction. Meanwhile, strong absorption bands observed in the 400 to 1000 cm−1 region suggest the presence of Ti\O\Ti and Ti\O\C bonds, indicating the chemical interaction between the surface hydroxyl groups of TiO2 and the functional groups of graphene oxide (Nguyen-Phan et al., 2011).

8000 6000 4000 2000

0 -2000

P25 GT 1:10 GT 1:20 GT 1:50 GT 1:100

A

g=1.995

g =1.970

1600

U

308

Intensity (a.u.)

307

339

F

317

306

2.7. FT-IR spectra

O

316

C\H bonds in the composite. The peak at 285.1 eV can be ascribed to carbon and defect-containing sp2 hybridized carbon atoms (Xiang et al., 2011a). The relatively weak peak at 288.3 eV was contributed by carboxyl carbon (O_C\O), indicating the esterification reaction of \OH on TiO2 with \COOH on the graphene oxide surface and the formation of O_C\O\Ti bonds (Xiang et al., 2011b). The peak located at 287.0 eV, which corresponded to the oxygen-containing carbonaceous bands, disappeared in the C 1s XPS spectra of GT. Therefore, graphene oxide was effectively reduced to graphene during the hydrothermal treatment. Moreover, the sp2 carbon network was retained in graphene, as indicated by the C\C peak at 284.2 eV (Murugan et al., 2009).

305

E

304

J O U RN A L OF E N V I RO N ME N TA L S CIE N CE S X X (2 0 1 4 ) XXX –XXX

1393 1250

3393 1720

g2 = 2.016g3 =2.008 g1 = 2.030

-4000

B

-6000 -8000 316

318

320

322

324

326

328

330

332

334

Magnetic field (G) Fig. 6 – EPR spectra of the synthesized graphene–TiO2 (GT) and P25 at 70 K.

4000

3500

3000

2500

2000

1500

1000

500

Wavenumber (cm-1) Fig. 7 – FT-IR spectra of the synthesized graphene–TiO2 (GT) and P25.

Please cite this article as: Zhang, Y., et al., Graphene TiO2 nanocomposites with high photocatalytic activity for the degradation of sodium pentachlorophenol, J. Environ. Sci. (2014), http://dx.doi.org/10.1016/j.jes.2014.08.011

341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356

359 360 361 362 363 364 365 366 367 368 369

7

J O U RN A L OF E N V I RO N ME N TA L S CI EN CE S X X (2 0 1 4 ) XX X–XXX

1.0

1.0

a

b

0.8

0.8

TOC (%)

4 3

0.6

0.4

2 1 0

0

20

40 60 80 100 120 Time (min)

0.0 0

20

40

GT 1:100 GT 1:50 GT 1:20 GT 1:10

0.2

60

80

100

0.0

120

0

20

40

Time (min)

F

0.2

O

Ct /C0

0.4

5

-ln(Ct /C0)

P25 GT 1:100 GT 1:50 GT 1:20 GT 1:10

0.6

60

80

100

120

R O

Time (min)

where, c is the PCP-Na concentration at time t, and k is the apparent first-order rate constant. The values of k for the initial 373 reaction under UV irradiation with different catalysts are shown 374 in Table 2. All degradation curves were fitted well by Eq. (1), with 2 −3 −1 375 Q18 all R values ≥ 0.946. For P25, k was 1.81 × 10 min , and the 376 degradation rate for GT 1:100 was 18.34 times higher than that 377 of P25. The order of photodegradation efficiency of PCP-Na 378 under UV light is as follows: GT 1:100 > GT 1:50 > GT 1:20 > GT 379 1:10 > P25. This result suggests that the GT 1:100 nanocompos380 ite is more efficient than P25 for PCP-Na degradation. In 381 addition, graphene oxide doping can significantly enhance 382 photodegradation efficiency. However, reaction rates decreased 383 with increasing graphene ratio (Table 2). Thus, excessive 384 graphene can be detrimental for the photocatalytic degradation 385 reaction. Although graphene doping can retard the recombina386 tion of electrons and holes, it can also obstruct photon 387 absorption, thereby decreasing degradation efficiency (Wang 388 and Zhang, 2011). 389 Fig. 8b shows the TOC variation in GT catalysts with 390 different graphene ratios. After 120 min of irradiation, 63.4% 391 to 82.9% of TOC was mineralized into CO2 and H2O, which was 392 lower than the decomposition rate of PCP-Na (97.7% to 99.0%). 393 GT 1:100 exhibited the highest mineralization efficiency. The 394 comparatively lower mineralization rate than degradation 395 rate suggests that dechlorination is stepwise and that the 396 dechlorinated intermediates are subsequently oxidized to

P

Fig. 8 – Photodegradation of PCP-Na (A) with graphene–TiO2 and P25 catalyst and first-order fitting curves (inset), and (B) TOC variation.

371 370

t2:4

Table 2 – Apparent first-order rate constant k values for the photocatalytic degradation of PCP-Na with graphene–TiO2 and P25 as catalysts.

t2:5

Catalyst

k (10−3 min−1)

R2

t2:6 t2:7 t2:8 t2:9 t2:10

P25 GT 1:10 GT 1:20 GT 1:50 GT 1:100

1.81 26.94 30.77 32.10 33.19

0.946 0.977 0.984 0.989 0.990

397

2.9. Photocatalytic mechanism

402

A schematic electronic structure and degradation process for PCP-Na by GT was proposed based on the abovementioned characterization and photodegradation results (Fig. 9). Under UV irradiation, the photogenerated electrons can be effectively conducted to the graphene sheet, leaving holes with positive charges on the surface of TiO2 particles. Graphene is an electron acceptor that accelerates the interfacial electron-transfer process. Graphene can strongly hinder the recombination of e−–h+ because of the Schottky barrier effect. As proved by EPR results, where signals with much higher intensity were shown for GT composites, the lifetime of electron–hole pairs can be prolonged, providing more radicals for the degradation of PCP-Na into H2O, CO2 and other intermediates. Thus, the photocatalytic activity of GT could be significantly enhanced. Though the BET surface area of the GT samples increased with increasing graphene content, GT 1:100 showed the highest degradation efficiency other than GT 1:10. This phenomenon can be ascribed to the fact that the degradation was not mainly based on the surface absorption of PCP-Na onto catalysts. Meanwhile, both the PCP-Na decomposition and TOC mineralization corresponded well with the UV–Vis DRS results, where the addition of graphene significantly increased the light absorbance, and GT 1:100 gives the highest intensity under UV light irradiation. However, the GT precursor P25 accumulated along the wrinkles and edges of the graphene oxide sheets during hydrothermal treatment because of the abundant carboxylic acid groups on graphene oxide, as shown in the SEM and TEM results. The inhomogeneous aggregation of TiO2 nanoparticles on graphene creates new e−–h+ recombination centers, thereby decreasing the photocatalytic efficiency.

403

E

T

C

E

R

R

N C O

U

t2:1 t2:2 t2:3 Q2

open the aromatic ring (Liu et al., 2011a). However, after 2 hr of reaction, the decreasing tendency of TOC was apparent in all GT cases. The continuous decrease in TOC can be attributed to the decomposition of small molecule acids, such as formic acid and acetic acid (Zhang et al., 2012b).

D

372

Please cite this article as: Zhang, Y., et al., Graphene TiO2 nanocomposites with high photocatalytic activity for the degradation of sodium pentachlorophenol, J. Environ. Sci. (2014), http://dx.doi.org/10.1016/j.jes.2014.08.011

398 399 400 401

404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432

8

J O U RN A L OF E N V I RO N ME N TA L S CIE N CE S X X (2 0 1 4 ) XXX –XXX

UV Electron e-

Hole h+

TiO2

Graphene UV PCP-Na

PCP-Na

Graphere

Graphere

F

H2O, CO2 and other mineralization

H2O, CO2 and other mineralization

O

PCP-Na

449 448

Acknowledgments

450 451

This work was supported by the National Natural Science Foundation of China (No. 41371472).

45 2

REFERENCES

453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472

An, X.Q., Yu, J.C., 2011. Graphene-based photocatalytic composites. Rsc. Adv. 1 (8), 1426–1434. Geng, W., Liu, H.X., Yao, X.J., 2013. Enhanced photocatalytic properties of titania–graphene nanocomposites: a density functional theory study. Phys. Chem. Chem. Phys. 15 (16), 6025–6033. Gopal, N.O.,Lo, H.H.,Ke, S.C., 2008. Chemical state and environment of boron dopant in B, N-codoped anatase TiO2 nanoparticles: an avenue for probing diamagnetic dopants in TiO2 by electron paramagnetic resonance spectroscopy. J. Am. Chem. Soc. 130 (9), 2760–2765. Homa, B., Andersson, M., Inganäs, O., 2009. Photogenerated charge carrier transport and recombination in polyfluorene/fullerene bilayer and blend photovoltaic devices. Org. Electron. 10 (3), 501–505. Kamegawa, T., Yamahana, D., Yamashita, H., 2010. Graphene coating of TiO2 nanoparticles loaded on mesoporous silica for enhancement of photocatalytic activity. J. Phys. Chem. C 114 (35), 15049–15053.

445 446

C

444

E

443

R

442

R

441

O

440

C

439

N

438

U

437

P

447

In this study, a series of GT nanocomposites of different carbon ratios were fabricated via a facile hydrothermal treatment using commercial P25 and graphene oxide as precursors. The asprepared composites exhibited high surface area, pore volume, and irregularity on the catalyst surface. Doping of C into TiO2 provided excellent electrical and optical properties to the catalysts, with graphene as the electron acceptor. Thus, the lifetime of photogenerated electron‒hole pairs was significantly prolonged. The photocatalytic activity of electron‒hole pairs for the degradation and mineralization of PCP-Na was much higher than that of commercial P25 under UV irradiation. However, excessive graphene in the nanocomposites decreased the photocatalytic efficiency.

436

D

435

Khalid, N.R., Ahmed, E., Hong, Z., Sana, L., Ahmed, M., 2013. Enhanced photocatalytic activity of graphene–TiO2 composite under visible light irradiation. Curr. Appl. Phys. 13 (4), 659–663. Kumar, C.P., Gopal, N.O., Wang, T.C., Wong, M.S., Ke, S.C., 2006. EPR investigation of TiO2 nanoparticles with temperaturedependent properties. J. Phys. Chem. B 110 (11), 5223–5229. Li, K.X., Xiong, J.J., Chen, T., Yan, L.S., Dai, Y.H., Song, D.Y., et al., 2013. Preparation of graphene/TiO2 composites by nonionic surfactant strategy and their simulated sunlight and visible light photocatalytic activity towards representative aqueous POPs degradation. J. Hazard. Mater. 250–251, 19–28. Liu, J.W., Han, R., Wang, H.T., Zhao, Y., Lu, W.J., Wu, H.Y., et al., 2011a. Degradation of PCP-Na with La–B co-doped TiO2 series synthesized by the sol–gel hydrothermal method under visible and solar light irradiation. J. Mol. Catal. A Chem. 344 (1–2), 145–152. Liu, J.W., Han, R., Zhao, Y., Wang, H.T., Lu, W.J., Yu, T.F., et al., 2011b. Enhanced photoactivity of V–N codoped TiO2 derived from a two-step hydrothermal procedure for the degradation of PCP-Na under visible light irradiation. J. Phys. Chem. C 115, 4507–4515. Long, R., 2013. Understanding the electronic structures of graphene quantum dot hysisorption and chemisorption onto the TiO2 (110) surface: a first-principles calculation. ChemPhysChem 14 (3), 579–582. Murugan, A.V., Muraliganth, T., Manthiram, A., 2009. Rapid, facile microwave-solvothermal synthesis of graphene nanosheets and their polyaniline nanocomposites for energy strorage. Chem. Mater. 21 (21), 5004–5006. Nguyen-Phan, T.D., Pham, V.H., Shin, E.W., Pham, H.D., Kim, S., Chung, J.S., et al., 2011. The role of graphene oxide content on the adsorption-enhanced photocatalysis of titanium dioxide/ graphene oxide composites. Chem. Eng. J. 170 (1), 226–232. Ribbens, S., Caretti, I., Beyers, E., Zamani, S., Vinck, E., Van Doorslaer, S., et al., 2011. Unraveling the photocatalytic activity of multiwalled hydrogen trititanate and mixed-phase anatase/ trititanate nanotubes: a combined catalytic and EPR study. J. Phys. Chem. C 115, 2302–2313. Savitzky, A., Golay, M.J.E., 1964. Smoothing and differentiation of data by simplified least squares procedures. Anal. Chem. 36 (8), 1627–1639. Shen, J.F., Shi, M., Yan, B., Ma, H.W., Li, N., Ye, M.X., 2011. Ionic liquid-assisted one-step hydrothermal synthesis of TiO2-reduced graphene oxide composites. Nano Res. 4 (8), 795–806. Shkrob, I.A., Marin, T.W., Chemerisov, S.D., Sevilla, M.D., 2011. Mechanistic aspects of photooxidation of polyhydroxylated molecules on metal oxides. J. Phys. Chem. C 115 (11), 4642–4648.

E

3. Conclusions

T

434 433

R O

Fig. 9 – Schematic of the PCP-Na photocatalytic degradation under UV irradiation with graphene–TiO2.

Please cite this article as: Zhang, Y., et al., Graphene TiO2 nanocomposites with high photocatalytic activity for the degradation of sodium pentachlorophenol, J. Environ. Sci. (2014), http://dx.doi.org/10.1016/j.jes.2014.08.011

473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521

9

J O U RN A L OF E N V I RO N ME N TA L S CI EN CE S X X (2 0 1 4 ) XX X–XXX

D

P

R O

O

F

Zhang, Y.P., Pan, C.X., 2011. TiO2/graphene composite from thermal reaction of graphene oxide and its photocatalytic activity in visible light. J. Mater. Sci. 46 (8), 2622–2626. Zhang, H., Lü, X.J., Li, Y.M., Wang, Y., Li, J.H., 2010a. P25–graphene composite as a high performance photocatalyst. ACS Nano 4 (1), 380–386. Zhang, X.Y., Li, H.P., Cui, X.L., Lin, Y.H., 2010b. Graphene/TiO2 nanocomposites: synthesis, characterization and application in hydrogen evolution from water photocatalytic splitting. J. Mater. Chem. 20 (14), 2801–2806. Zhang, Y., Tang, Z.R., Fu, X., Xu, Y.J., 2010c. TiO2-graphene nanocomposites for gas-phase photocatalytic degradation of volatile aromatic pollutant: is TiO2-graphene truly different from other TiO2-carbon composite materials? ACS Nano 4 (12), 7303–7314. Zhang, Y.H., Tang, Z.R., Fu, X., Xu, Y.J., 2011. Engineering the unique 2D mat of graphene to achieve graphene–TiO2 nanocomposite for photocatalytic selective transformation: what advantage does graphene have over its forebear carbon nanotube? ACS Nano 5 (9), 7426–7435. Zhang, N., Zhang, Y.H., Xu, Y.J., 2012a. Recent progress on graphene-based photocatalysts: current status and future perspectives. Nanoscale 4 (19), 5792–5813. Zhang, Y.X., Wu, H.Y., Zhang, J., Wang, H.T., Lu, W.J., 2012b. Enhanced photodegradation of pentachlorophenol by single and mixed cationic and nonionic surfactants. J. Hazard. Mater. 221–222, 92–99. Zhang, Y.X., Zhao, Y., Zhu, Y., Wu, H.Y., Wang, H.T., Lu, W.J., 2012c. Adsorption of mixed cationic–nonionic surfactant and its effect on bentonite structure. J. Environ. Sci. China 24 (8), 1525–1532. Zhu, J.X., Zhu, L.Z., Zhu, R.L., Tian, S.L., Li, J.W., 2009. Surface microtopography of surfactant modified montmorillonite. Appl. Clay Sci. 45 (1–2), 70–75.

E

T

Sing, K.S.W., Everett, D.H., Haul, R.A.W., Moscou, L., Pierotti, R.A., Rouquerol, J., et al., 1985. Reporting physisorption data for gas solid systems with special reference to the determination of surface-area and porosity (Recommendations 1984). Pure Appl. Chem. 57 (4), 603–619. Song, J.L., Yin, Z.Y., Yang, Z.J., Amaladass, P., Wu, S.X., Ye, J., et al., 2011. Enhancement of photogenerated electron transport in dye-sensitized solar cells with introduction of a reduced graphene oxide–TiO2 junction. Chem. A. Eur. J. 17 (39), 10832–10837. Venkatachalam, N., Palanichamy, M., Arabindoo, B., Murugesan, V., 2007. Enhanced photocatalytic degradation of 4-chlorophenol by Zr4+ doped nano TiO2. J. Mol. Catal. A Chem. 266 (1–2), 158–165. Wang, F., Zhang, K., 2011. Reduced graphene oxide–TiO2 nanocomposite with high photocatalystic activity for the degradation of rhodamine B. J. Mol. Catal. A Chem. 345 (1–2), 101–107. Wang, P., Wang, J., Wang, X., Yu, H., Yu, J., Lei, M., et al., 2013. One-step synthesis of easy-recycling TiO2–rGO nanocomposite photocatalysts with enhanced photocatalytic activity. Appl. Catal. B Environ. 132, 452–459. Xiang, Q.J., Yu, J.Q., Jaroniec, M., 2011a. Preparation and enhanced visible-light photocatalytic H-2-production activity of graphene/C3N4 composites. J. Phys. Chem. C 115, 7355–7363. Xiang, Q.J., Yu, J.G., Jaroniec, M., 2011b. Enhanced photocatalytic H2-production activity of graphene-modified titania nanosheets. Nanoscale 3 (9), 3670–3678. Xie, J., Bian, L., Yao, L., Hao, Y.J., Wei, Y., 2013. Simple fabrication of mesoporous TiO2 microspheres for photocatalytic degradation of pentachlorophenol. Mater. Lett. 91, 213–216. Yang, M.Q.,Zhang, N.,Xu, Y.J., 2013. Synthesis of fullerene-, carbon nanotube-, and graphene-TiO2 nanocomposite photocatalysts for selective oxidation: A comparative study. Acs Appl. Mater. Int. 5 (3), 1156–1164.

C

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

U

N C O

R

R

E

591

Please cite this article as: Zhang, Y., et al., Graphene TiO2 nanocomposites with high photocatalytic activity for the degradation of sodium pentachlorophenol, J. Environ. Sci. (2014), http://dx.doi.org/10.1016/j.jes.2014.08.011

556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590

Graphene TiO2 nanocomposites with high photocatalytic activity for the degradation of sodium pentachlorophenol.

A series of graphene-TiO2 photocatalysts was synthesized by doping TiO2 with graphene oxide via hydrothermal treatment. The photocatalytic capability ...
2MB Sizes 0 Downloads 7 Views