Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Journal of Colloid and Interface Science 377 (2012) 114–121

Contents lists available at SciVerse ScienceDirect

Journal of Colloid and Interface Science


www.elsevier.com/locate/jcis

ZnO/graphene-oxide nanocomposite with remarkably enhanced


visible-light-driven photocatalytic performance
Benxia Li ⇑, Tongxuan Liu, Yanfen Wang, Zhoufeng Wang
School of Materials Science and Engineering, Anhui University of Science and Technology, Huainan, Anhui 232001, People’s Republic of China

a r t i c l e i n f o a b s t r a c t

Article history: In this work, a high-performance photocatalyst of ZnO/graphene-oxide (ZnO/GO) nanocomposite was
Received 30 December 2011 synthesized via a facile chemical deposition route and used for the photodegradation of organic dye from
Accepted 20 March 2012 water under visible light. The nanocomposite was characterized by X-ray diffraction, X-ray photoelectron
Available online 28 March 2012
spectroscopy, scanning electron microscopy, transmission electron microscopy, Brunauer–Emmett–
Teller N2 adsorption–desorption analysis, and UV–Vis diffusion reflectance spectroscopy. The ZnO/GO
Keywords: nanocomposite consisting of flower-like ZnO nanoparticles anchored on graphene-oxide sheets has a
ZnO
high surface area and hierarchical porosity, which is benefit to the adsorption and mass transfer of dye
Graphene oxide
Nanocomposite
and oxygen species. For the photodegradation of organic dyes under visible light, ZnO/GO nanocomposite
Synthesis exhibited remarkably enhanced photocatalytic efficiency than graphene-oxide sheets and flower-like
Photocatalysis ZnO particles. Moreover, the photocatalytic efficiency of ZnO/GO nanocomposite could be further
Visible light improved by annealing the product in N2 atmosphere. The outstanding photocatalytic performance
was ascribed to the efficient photosensitized electron injection and repressed charge carriers recombina-
tion in the composite with GO as electron collector and transporter, thus leading to continuous genera-
tion of reactive oxygen species for the degradation of methylene blue.
Ó 2012 Elsevier Inc. All rights reserved.

1. Introduction its great potential, the photocatalytic efficiency remains very low
because of the fast recombination of the photogenerated elec-
Currently, organic dyes and their effluents have become one of tron–hole pairs in the single phase semiconductor.
the main sources of water pollution due to the greater demand in The performance of semiconductor photocatalysts is often en-
industry such as textile, paper, and plastic. These organic dyes con- hanced by means of noble metal loading [18–20], ion doping
taminate environment by the release of the toxic, cancerogenic, [21,22], and incorporation of electron-accepting materials [23–
and colored wastewater [1–5]. Most of these dyes escape from tra- 25]. These actions are used to extend the light absorption range
ditional wastewater treatment and persist in water because of or suppress the electron–hole recombination. As a rising star of
their high stability against light, temperature, chemicals, and carbon family, graphene has become the focus of considerable
microbial attack [3–5], whereas photocatalytic degradation of the interest because of its unique electronic properties and other
organic pollutants has opened a new door for the elimination of excellent attributes, such as the large theoretical specific surface
organic dyes in wastewater [6–10]. The photodegradation process area and the high transparency [26–29]. Meanwhile, graphene
utilizes cheaply available semiconductors and leads to complete oxide (GO) is receiving increasing attention because it possesses
mineralization of organic compounds to CO2, water, and mineral the similar properties to graphene as well as the special surface
acids [11]. Thus, semiconductor photocatalysis is deemed to an structures with the introduced hydroxyl and carboxyl groups for
alternative technique to remove various organic pollutants. Until synthesis of GO-containing nanocomposites [30–38]. Particularly,
now, ZnO and TiO2 photocatalysts have been widely used because the fabrication of semiconductor/GO composites has attracted sub-
of its strong oxidizing power, non-toxic nature, and low cost [12– stantial research efforts motivated by the desire to improve the
16]. ZnO is a wide band-gap semiconductor oxide (3.37 eV) with a photocatalytic efficiency [34–38]. The recent studies have revealed
conduction band edge located at approximately the same level as that the composites simultaneously covered three excellent attri-
that of TiO2. More attractively, the electron mobility of ZnO has butes: the increasing adsorptivity of pollutants, extended light
been proven to be higher than that of TiO2 [17]. However, despite absorption range, and efficient charge transportation and separa-
tion [25,39], which are the ideal traits of a photocatalyst we have
been pursuing for. Therefore, it is believed that anchoring well-
⇑ Corresponding author. Fax: +86 554 6668643.
organized ZnO nanostructures on GO sheets can efficiently utilize
E-mail address: libx@mail.ustc.edu.cn (B. Li).

0021-9797/$ - see front matter Ó 2012 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.jcis.2012.03.060
B. Li et al. / Journal of Colloid and Interface Science 377 (2012) 114–121 115

the combinative merits of ZnO and GO to obtain a photocatalyst MKII X-ray photoelectron spectrometer with an exciting source
with superior performance. of Mg Ka. The nitrogen adsorption and desorption isotherms at
This work demonstrated a facile strategy to synthesize ZnO/GO 77 K were measured using a Micromeritics ASAP 2000 system after
nanocomposite consisting of flower-like ZnO nanoparticles an- the sample was degassed in a vacuum at 130 °C overnight. The UV–
chored on GO sheets and its use for photocatalytic degradation of Vis diffuse reflectance spectra (DRS) were recorded on a recording
organic dye in water under visible light. The composition, mor- spectrophotometer of Perkin Elmer Lambda 950.
phology, and microstructure of the as-obtained ZnO/GO nanocom-
posite were characterized. The photocatalytic performance of ZnO/ 2.5. Photocatalytic property test
GO nanocomposite was evaluated by the photodegradation of
methylene blue in water and compared with that of pure flower- The photocatalytic properties of the samples were evaluated by
like ZnO nanoparticles and GO sheets, to highlight the importance photodegradation of methylene blue (MB) in water under visible-
of the anchoring of ZnO nanoparticles on GO sheets for maximum light irradiation from a 300 W Xe light equipped with a 420 nm
utilization of ZnO photocatalyst and GO as electron collector and cutoff filter (CEL-HXF300/CEL-HXUV300, China). In every experi-
transporter in photocatalytic degradation of organic pollutants. ment, 80 mg of photocatalyst was suspended in 100 mL of a
5.0  105 M aqueous solution of MB. Prior to irradiation, the sus-
2. Experimental pension was stirred in the dark for 2 h to achieve an adsorption–
desorption equilibrium between the photocatalyst and MB mole-
2.1. Materials cules. After that, the solution was exposed to the visible-light irra-
diation under magnetic stirring. At given time intervals, 3 mL of
Graphite flakes (200 mesh) were purchased from Qingdao solutions was sampled for analysis of the MB concentration. The
Tianyuan Company. 37% hydrochloric acid (HCl), 98% sulfuric acid photocatalytic degradation process was monitored using a UV–
(H2SO4), 85% phosphoric acid (H3PO4), 30% hydrogen peroxide Vis spectrophotometer (Shimadzu UV2550) to record the charac-
(H2O2), potassium permanganate (KMnO4), zinc chloride (ZnCl2), teristic absorption at 665 nm.
sodium hydroxide (NaOH), and absolute ethanol were purchased
from Shanghai Sinopharm Chemical Reagent Co. Ltd. All chemicals 3. Results and discussion
were analytical grade without further purification. Deionized
water was used throughout this study. 3.1. The formation mechanism of ZnO/GO composite

2.2. Preparation of graphene oxide (GO) sheets In this work, ZnO/GO nanocomposite was fabricated by a two-
step aqueous-solution route. GO sheets were firstly prepared using
The graphene oxide was synthesized by chemical oxidation and an improved Hummers’ method, and ZnO nanoparticles were then
exfoliation of graphite flakes using an improved Hummers’ method anchored on GO sheets via a facile reaction between Zn2+ and OH
[40]. Graphite flakes (1.5 g) and KMnO4 (9.0 g) were successively ions in aqueous solution. Fig.1 illustrates the fabrication process
added into a 9:1 mixture of concentrated H2SO4/H3PO4 and formation mechanism of ZnO/GO composite. It has been
(180:20 mL) under continuous stirring. The reaction was heated proved that the surfaces of the chemically exfoliated GO sheets
at 50 °C and stirred for 12 h. Then, the reaction was cooled to room are covered by a large number of hydroxyl, carboxyl, and epoxy
temperature and poured onto ice (200 mL) with 30% H2O2 (3 mL). groups that are introduced on GO sheets due to oxidation proce-
The mixture was then repeatedly centrifuged and washed in suc- dures [41,42]. These functional groups can act as anchor sites to
cession with water, 30% HCl solution, and ethanol. The obtained so- enable the subsequent in situ formation of ZnO nanoparticles on
lid product of GO was finally dried at 80 °C. GO sheets. The formation of ZnO/GO composite undergoes the fol-
lowing two distinctive stages: (i) when dissolving ZnCl2 into GO
2.3. Preparation of ZnO/GO nanocomposite suspension, Zn2+ ions will be adsorbed onto the surfaces of GO
sheets due to their bonding with the O atoms of the negatively
To synthesize ZnO/GO composite, 0.1 g dried GO precipitate charged oxygen-containing functional groups via electrostatic
was dispersed in 40 mL of water to form GO suspension by ultra- force. (ii) After the addition of NaOH, the crystal growth units of
sonication, in which a further exfoliation of GO was achieved. Then, ZnðOHÞ24 and ZnO
2
may combine with the functional groups of
zinc chloride (ZnCl2, 1.0 mmol, 0.1364 g) and sodium hydroxide GO sheets via intermolecular hydrogen bonds or coordination
(NaOH, 10.0 mmol, 0.4000 g) were successively dissolved in the bonds, acting as anchor sites for ZnO nanoparticles. With heating
above GO suspension. The mixture was sealed in a glass bottle at 90 °C, a large number of ZnO nuclei are formed in a short time
(60 mL), kept static at 90 °C for 6 h, and then cooled to room tem- due to the hydrolysis reaction of ZnðOHÞ24 . Finally, ZnO/GO nano-
perature naturally. Finally, the composite was filtered, washed sev- composite is obtained. The in situ formation of ZnO nanoparticles
eral times with distilled water and ethanol, and dried at 80 °C for in return caused the exfoliation of the lamellar GO.
24 h. For comparison, flower-like ZnO nanoparticles were synthe-
sized by a similar method without adding GO. In addition, the 3.2. XRD and XPS analysis
ZnO/GO product prepared via aqueous-solution route at 90 °C
was further annealed at 400 °C and in N2 for 2 h. Fig. 2 shows the XRD patterns of ZnO nanoparticles, ZnO/GO
nanocomposite, and GO sheets, respectively. All the diffraction
2.4. Characterizations of the samples peaks in the XRD pattern of ZnO nanoparticles are consistent with
the hexagonal phase wurtzite ZnO (JCPDS No. 36-1451). The dif-
The products were characterized by X-ray diffraction (XRD, Shi- fraction peak at around 2h = 8° in the XRD pattern of GO sheets be-
madzu XRD-6000, Cu Ka radiation), field-emission scanning elec- longs to the (0 0 1) reflection of GO, and the interlayer spacing
tron microscopy (FE-SEM, Sirion200), and transmission electron (0.95 nm) is much larger than that of graphite (about 0.37 nm) ow-
microscopy/high resolution transmission electron microscopy ing to the introduction of oxygen-containing functional groups on
(TEM/HRTEM, H-7650 and JEOL-2010). X-ray photoelectron spec- the graphite sheet surfaces [43,44]. The diffraction peaks of ZnO/
troscopy (XPS) measurements were performed on a VGESCALAB GO nanocomposite are similar to those of hexagonal (wurtzite)
116 B. Li et al. / Journal of Colloid and Interface Science 377 (2012) 114–121

Fig. 1. Illustration of the fabrication process and formation mechanism for ZnO/GO nanocomposite. (1) Oxidation of graphite to graphite oxide with larger interlayer spacing.
(2) GO sheets from the ultrasonic exfoliation of graphite oxide. (3) Adsorption and bonding of Zn2+ ions onto the GO sheets. (3) The nucleation and growth of ZnO crystallites,
resulting in ZnO/GO nanocomposite.

Fig. 2. (a) XRD patterns of ZnO nanoparticles, ZnO/GO nanocomposite and GO; (b–d) XPS spectra of ZnO/GO nanocomposite.

ZnO. Nevertheless, the (0 0 1) diffraction peak at 8° of GO dropped survey spectrum (Fig. 2b) shows the presence of Zn and O as well
profoundly to an almost undetectable level, which may indicate as C. The higher-resolution spectrum of Zn2p3/2 in Fig. 2c shows
that the regular layered structure of GO has been destroyed and that the level of Zn2p3/2 is 1021.68 eV. The C1s XPS spectrum of
exfoliated GO sheets are formed due to the growth of ZnO nano- ZnO/GO (Fig. 2d) clearly indicates a considerable degree of graph-
crystals [31]. The XRD patterns demonstrate that the ZnO/GO ene oxide with three main components that correspond to carbon
nanocomposite can be obtained in the present system. Further evi- atoms in different functional groups: the non-oxygenated CAC
dence for the composition of the ZnO/GO nanocomposite was ob- bond (284.79 eV), the CAO of epoxy and hydroxyl (286.15 eV),
tained by the X-ray photoelectron spectra (XPS, Fig. 2b–d). The and the carbonyl C@O (288.35 eV) [35–37]. The small peak at
B. Li et al. / Journal of Colloid and Interface Science 377 (2012) 114–121 117

290.71 eV can be ascribed to the p–p satellite of the aromatic con- flower-like ZnO particles were mainly located at the edge of the
jugated domains modified with carboxylic acid, hydroxyl, and GO sheets, which might result from the aggregation of ZnO parti-
epoxide groups in graphene oxide [45,46]. Therefore, the peak at cles that confined their efficient dispersion on GO sheets. Fig. 4d
290.71 eV suggests the existence of a partially modified p structure shows HRTEM image of a ZnO fragment marked with the circle
in the ZnO/GO nanocomposite. in Fig. 4c, which exhibits well-resolved two-dimensional lattice
fringes with the spacings of 0.52 nm and 0.28 nm, in good agree-
ment with the interplanar spacings of {0 0 0 1} and {0 1 1 0} planes
3.3. SEM and TEM images of hexagonal (wurtzite) ZnO.

FESEM images in Fig. 3a–d show the morphologies and micro-


structures of GO sheets, pure ZnO nanoparticles, and ZnO/GO 3.4. N2 adsorption/desorption measurement
nanocomposite, respectively. From the FESEM image in Fig. 3a,
the layered structure of the stacked GO sheets can be seen, and Due to the high surface area of graphene sheets (calculated va-
there are many wrinkles through all the surfaces of GO sheets. lue: 2630 m2 g1), a higher surface area of ZnO/GO nanocomposite
The solid GO sample is severely agglomerated because of its high than pure ZnO is expected. The N2 adsorption–desorption iso-
specific surface area. Fig. 3b shows an FESEM image of pure ZnO therms at 77 K and pore-size distribution plot calculated by the
nanoparticles, and it reveals that the detailed morphology of ZnO Barrett–Joyner–Halenda (BJH) method of ZnO/GO nanocomposite
is flower-like aggregates with diameters about 1 lm, assembled are shown in Fig. 5. The isotherms (Fig. 5a) present a reverse ‘‘S’’
by many densely arranged nanosheets. In the presence of GO mate- shape, which is identified as type IV and characteristic of mesopor-
rial, ZnO/GO nanocomposite with flower-on-sheet morphology ous structures [47,48] From the adsorption branch of the iso-
(Fig. 3c and d) was obtained under similar experimental condition therms, the specific surface area of 234.0561 m2 g1 is calculated
for preparing flower-like ZnO nanoparticles. As shown in Fig. 3c, for ZnO/GO nanocomposite through a multi-point Brunauer–Em-
GO nanosheets were well decorated by the flower-like ZnO parti- mett–Teller (BET) method [49]. The distributions of pore sizes
cles. Moreover, some ZnO particles entered into the interlayers of shown in Fig. 5b indicate that there are two types of pores in the
GO sheets to form a sandwich-like composite structure (Fig. 3d), ZnO/GO nanocomposite. One type of them is concentrated in the
which can prevent the stacking of GO sheets, and thus avoid the range of typical mesoporous structure of 2.8 nm, which is probably
loss of their high active surface area. attributed to the interspaces between the layers of GO sheets. The
Fig. 4 shows TEM images of GO sheets and ZnO/GO composite. It other type of pores with larger sizes distributes around 65.8 nm,
is observed from Fig. 4a that the GO sheets present a crumpled sur- which presumably arises from the spaces between nanosheets in
face, which is consistent with the observation from FESEM image. the flower-like ZnO particles [12]. The average pore diameter of
Fig. 4b and c show TEM images with different sizes of ZnO/GO 16.6949 nm is calculated for ZnO/GO nanocomposite using a Bar-
nanocomposite. The light-gray thin films are the GO sheets, and ret–Joyner–Halenda (BJH) model. The cumulative volume of pores
the dark regions on the GO background are due to the presence with diameters between 1.7 and 300 nm is 0.9743 cm3 g1. For
of ZnO particles. It can be clearly seen in Fig. 4b that the exfoliated comparison, the BET surface area of pure ZnO with flower-like
GO sheet was decorated with ZnO aggregates with sizes of 0.5– morphology was calculated to be 25.1617 m2 g1. The higher spe-
1 lm, whereas some of the flower-like ZnO microstructures on cific surface area and hierarchical porosity of ZnO/GO nanocom-
GO sheets were crushed into fragments because of the ultrasonic posite provide the possibility for the efficient adsorption and
treatment before TEM observation. Moreover, the loaded mass transfer of the degradable organic molecules and hydroxyl

Fig. 3. FESEM images of (a) GO sheets, (b) pure ZnO, (c and d) ZnO/GO nanocomposite.
118 B. Li et al. / Journal of Colloid and Interface Science 377 (2012) 114–121

Fig. 4. TEM images of (a) GO sheets, (b and c) ZnO/GO nanocomposite, (d) HRTEM image of a ZnO fragment marked with the circle in (c).

Fig. 5. (a) N2 adsorption–desorption isotherms at 77 K and (b) pore diameter distribution of ZnO/GO nanocomposite.

radicals in photochemical reaction, and a superior photocatalytic


performance of ZnO/GO nanocomposite should be expected.

3.5. UV–Vis diffuse reflectance spectra and photocatalytic performance

Fig. 6 shows UV–Vis diffuse reflectance spectra of pure ZnO and


ZnO/GO nanocomposite. ZnO sample shows the characteristic
spectrum with its fundamental absorption edge rising at 400 nm,
while the ZnO/GO composite shows absorption in the whole visible
region. Because ZnO and GO in the ZnO/GO composite are different
two phases, their band-gap energies were not changed [50]. The
absorption edge of ZnO can also be detected in the UV–Vis spec-
trum of ZnO/GO sample. Besides, an intense and broad background
absorption in visible region was observed due to the presence of
GO.
The photocatalytic performance of ZnO/GO nanocomposite was
evaluated by examining the photodegradation of methylene blue
(MB) as a representative pollutant under visible-light irradiation
from a 300 W Xe lamp. Prior to irradiation, the photocatalytic Fig. 6. UV–Vis diffuse reflection spectra of pure ZnO and ZnO/GO nanocomposite.
B. Li et al. / Journal of Colloid and Interface Science 377 (2012) 114–121 119

Fig. 7. (a) The time-dependent absorption spectra of MB solution (5.0  105 mol/L, 100 mL) in the presence of ZnO/GO nanocomposite (80 mg) and under visible-light
irradiation, and the inset is the color-change sequence of MB solution during this process. (b) Photodegradation of MB over photocatalyst-free solution (blank), GO sheets,
flower-like ZnO particles, ZnO/GO nanocomposite, and annealed ZnO/GO, respectively.

Fig. 8. (a) Kinetic linear simulation curves and (b) the reaction rates of photocatalytic degradation of MB over photocatalyst-free solution (blank), GO sheets, flower-like ZnO
particles, ZnO/GO nanocomposite, and annealed ZnO/GO, respectively.

reaction system was magnetically stirred in the dark for 2 h to photocatalytic efficiency, and 98.1% of MB was photodegraded
reach the adsorption/desorption equilibrium of MB on the surface from the aqueous solution after visible-light irradiation for
of the photocatalyst. Fig. 7a shows the UV–Vis absorption spec- 60 min. Moreover, the photocatalytic efficiency of ZnO/GO nano-
trum of the aqueous solution of MB (initial concentration, composite could be further improved by annealing the product in
5.0  105 mol/L; 100 mL) with 80 mg of ZnO/GO nanocomposite N2 atmosphere. The time-dependent absorption spectra of MB
obtained at low temperature as photocatalyst and exposure to solution in the presence of annealed ZnO/GO (80 mg) and under
the visible light for various durations. The characteristic absorption visible-light irradiation are shown in Fig. S1 (Supporting informa-
of MB at 665 nm decreases rapidly with extension of the exposure tion), which demonstrates that the characteristic absorption of
time, and almost disappears after about 60 min. Further exposure
leads to no absorption peak in the whole spectrum. The color-
change sequence in the MB solution during this process is shown
in the inset of Fig. 7a, from which it is clear that the intense blue
color of the initial solution gradually disappears with increasingly
longer exposure times. To demonstrate the synergy-induced
enhancement of the photocatalytic efficiency of ZnO/GO nanocom-
posite, contrastive experiments were performed using flower-like
ZnO particles and GO sheets as photocatalyst for the photodegra-
dation of MB, respectively. The results of the MB degradation in a
series of experimental conditions are summarized in Fig. 7b, where
C0 and Ct are the initial concentration after the equilibrium adsorp-
tion and the residual concentration of MB, respectively. Without
any photocatalyst (blank), there was hardly any degradation of
MB solution under visible-light irradiation. In the presence of GO
sheets, MB solution was just slightly degraded after 100 min. By
comparison, the degradation rate of MB in the presence of flow-
er-like ZnO particles is 54.3% after 100 min of irradiation. However,
Fig. 9. Cycling runs in the photocatalytic degradation of MB in the presence of ZnO/
ZnO/GO nanocomposite was found to exhibit very prominent GO nanocomposite (80 mg) and under visible-light irradiation.
120 B. Li et al. / Journal of Colloid and Interface Science 377 (2012) 114–121

Fig. 10. Possible mechanism of photosensitized degradation of dye over ZnO/GO nanocomposite under visible light.

MB almost disappears after 40 min of the exposure time. As shown self-degradation or degradation by the reactive oxidation species.
in Fig. 7b, 97.9% of MB was photodegraded from the aqueous solu- Such a photocatalytic degradation process for dyes has been dem-
tion in the presence of the annealed ZnO/GO after only 40 min of onstrated to be an efficient approach to remove textile dyestuff
visible-light irradiation. The kinetic linear simulation curves of from aquatic environments [37,38,55,56]. Considering the wide
photocatalytic degradation of MB with photocatalyst-free solution band-gap (3.37 eV) of ZnO, degradation of MB or MO under visible
(blank), GO sheets, flower-like ZnO particles, ZnO/GO nanocom- light in the present case should be conducted following the dye-
posite, and annealed ZnO/GO are presented in Fig. 8a, respectively. excitation mechanism, as illustrated in Fig. 10. GO sheets with
It is clear that the curve with irradiation time as abscissa and ln(Ct/ large specific surface area and numerous oxygenic groups allow
C0) as the vertical ordinate is close to a linear curve. The values of k good access of dye molecules to their surfaces in photocatalytic
(Fig. 8b) for photocatalyst-free solution (blank), GO sheets, flower- system. For example, MB was firstly excited to MB, followed by
like ZnO particles, ZnO/GO nanocomposite, and annealed ZnO/GO electron transfer to the conduction band of ZnO via the transporta-
are 0.00072, 0.0024, 0.0077, 0.064, and 0.099 min1, respectively, tion in GO sheets, then the photoelectrons were captured by sur-
indicating that the ZnO/GO nanocomposite has remarkably en- face-adsorbed O2 to generate the reactive oxidation species (such
hanced photocatalytic activity, which may be attributed to the as  OH;O2 ). Finally, the MB
+
radical was degraded itself or by
interactions between the GO sheets and ZnO nanoparticles. Never- the formed reactive oxidation species. Such an efficient electron-
theless, the photocatalytic activity of the ZnO/GO nanocomposite transfer process with GO as electron collector and transporter is
obtained via aqueous-solution route at low temperature has not responsible for the enhanced photocatalytic performance under
achieved the best result because the surface of the product was visible light. As for GO sheets used as photocatalyst, although the
presented by a number of OH groups, which is believed to quench efficient electron-transfer process can be occurred between MB
photogenerated chargers. After annealed at 400 °C and in N2, the and GO, a very slow reaction rate is caused by the electron accu-
OH groups were eliminated. Therefore, the annealed ZnO/GO mulation on GO surface and the rapid electron-MB+ recombina-
exhibited further improved photocatalytic efficiency. To demon- tion [56].
strate the photocatalytic effect of ZnO/GO nanocomposite toward
a more stable dye, the photodegradation of methyl orange (MO) 4. Conclusion
in water under the visible-light irradiation was carried out. The
time-dependent absorption spectra of MO solution in the presence In conclusion, the ZnO/GO nanocomposite was successfully syn-
of ZnO/GO nanocomposite (Fig. S2, Supporting information) indi- thesized via a facile chemical deposition route at low temperature
cate that the characteristic absorption of MO almost disappeared and its use for the photodegradation of organic dye from water un-
after about 80 min, and the color of MO solution changed gradually der visible light was investigated. The ZnO/GO nanocomposite is
from yellow to colorless after irradiation for 80 min. composed of flower-like ZnO nanoparticles anchored on graph-
The stability of photocatalyst during photocatalytic reaction is a ene-oxide sheets. For the photodegradation of organic dyes from
crucial factor for the practical applications. To test the reusability water under visible light, ZnO/GO nanocomposite exhibits much
of ZnO/GO nanocomposite in MB photodegradation, five cycles of higher photocatalytic efficiency than GO sheets and flower-like
the photocatalytic experiment for ZnO/GO nanocomposite were ZnO particles. The enhanced photocatalytic performance of ZnO/
carried out. As shown in Fig. 9, MB could be totally decomposed GO nanocomposite can be attributed to the efficient photosensi-
in each cycle and the ZnO/GO photocatalyst exhibited almost no tized electron injection and repressed electron recombination
change in its photocatalytic activity during the repeated photocat- due to the electron-transfer process with GO as electron collector
alytic experiments. Thus, the higher photocatalytic activity and and transporter. These features make the ZnO/GO composite an
reusability of ZnO/GO nanocomposite were beneficial for its appli- excellent candidate for applications relating to a number of envi-
cation as a photocatalyst. ronmental issues. The preparation method may be extended to fab-
The photocatalytic degradation of organic dyes by semiconduc- ricate more graphene-based composites for a variety of
tor under visible-light irradiation generally involves two mecha- applications, such as catalysts, gas sensors, and nanoelectronic
nisms. The first is based on the excitation of the semiconductor, devices.
which involves excitation of the semiconductor by light irradiation
to form photogenerated electrons in the conduction band and Acknowledgments
holes in the valence band, and the subsequent chemical reactions
with the surrounding media after the photogenerated charges Financial supports from the National Natural Science Founda-
move to the particle surface [51]. The other mechanism is based tion of China (21001003), Natural Science Foundation of Anhui
on the excitation of dye [52–54], in which the dye acts as a sensi- Province of China (10040606Q15), Natural Science Research for
tizer of visible light as well as injects excited electrons to an elec- Colleges, and Universities of Anhui Province of China (KJ2010
tron acceptor to become a cationic dye radical (dye+), followed by A101) are acknowledged.
B. Li et al. / Journal of Colloid and Interface Science 377 (2012) 114–121 121

Appendix A. Supplementary material [24] J.B. Mu, C.L. Shao, Z.C. Guo, Z.Y. Zhang, M.Y. Zhang, P. Zhang, B. Chen, Y.C. Liu,
ACS Appl. Mater. Interfaces 3 (2011) 590–596.
[25] H. Zhang, X.J. Lv, Y.M. Li, Y. Wang, J.H. Li, ACS Nano 4 (2010) 380–386.
Supplementary data associated with this article can be found, in [26] A.K. Geim, K.S. Novoselov, Nat. Mater. 6 (2007) 183–191.
the online version, at http://dx.doi.org/10.1016/j.jcis.2012.03.060. [27] C. Lee, X.D. Wei, J.W. Kysar, J. Hone, Science 321 (2008) 385–388.
[28] L.Y. Jiao, L. Zhang, X.R. Wang, G. Diankov, H.J. Dai, Nature 458 (2009) 877–880.
[29] M.J. Allen, V.C. Tung, R.B. Kaner, Chem. Rev. 110 (2010) 132–145.
[30] Y.Y. Liang, Y.G. Li, H.L. Wang, J.G. Zhou, J. Wang, T. Regier, H.J. Dai, Nat. Mater.
References 10 (2011) 780–786.
[31] J.W. Zhu, G.Y. Zeng, F.D. Nie, X.M. Xu, S. Chen, Q.F. Han, X. Wang, Nanoscale 2
[1] I. Arslan, I.A. Balcioglu, T. Tuhkanen, D. Bahnemann, J. Environ. Eng. 126 (2000) (2010) 988–994.
903–911. [32] J. Wang, Z. Gao, Z.S. Li, B. Wang, Y.X. Yan, Q. Liu, T. Mann, M.L. Zhang, Z.H. Jiang,
[2] T. Robinson, G. Mcmullan, R. Marchant, P. Nigam, Bioresour. Technol. 77 (2001) J. Solid State Chem. 184 (2011) 1421–1427.
247–255. [33] Y.L. Chen, Z.A. Hu, Y.Q. Chang, H.W. Wang, Z.Y. Zhang, Y.Y. Yang, H.Y. Wu, J.
[3] E. Forgacs, T. Cserhati, G. Oros, Environ. Int. 30 (2004) 953–971. Phys. Chem. C 115 (2011) 2563–2571.
[4] H.S. Rai, M.S. Bhattacharyya, J. Singh, T.K. Bansal, P. Vats, U.C. Banerjee, Crit. [34] Y. Yang, L.L. Ren, C. Zhang, S. Huang, T.X. Liu, ACS Appl. Mater. Interfaces 3
Rev. Environ. Sci. Technol. 35 (2005) 219–238. (2011) 2779–2785.
[5] V.K. Gupta, Suhas, J. Environ. Manage. 90 (2009) 2313–2342. [35] G.D. Jiang, Z.F. Lin, C. Chen, L.H. Zhu, Q. Chang, N. Wang, W. Wei, H.Q. Tang,
[6] M.R. Hoffmann, S.T. Martin, W.Y. Choi, D.W. Bahnemannt, Chem. Rev. 95 Carbon 49 (2011) 2693–2701.
(1995) 69–96. [36] Y.H. Ng, A. Iwase, A. Kudo, R. Amal, J. Phys. Chem. Lett. 1 (2010) 2607–2612.
[7] P.A. Pekakis, N.P. Xekoukoulotakis, D. Mantzavinos, Water Res. 40 (2006) [37] B.J. Li, H.Q. Cao, J. Mater. Chem. 21 (2011) 3346–3349.
1276–1286. [38] Z.G. Xiong, L.L. Zhang, X.S. Zhao, Chem. Eur. J. 17 (2011) 2428–2434.
[8] C.C. Chen, W.H. Ma, J.C. Zhao, Chem. Soc. Rev. 39 (2010) 4206–4219. [39] N.L. Yang, J. Zhai, D. Wang, Y.S. Chen, L. Jiang, ACS Nano 4 (2010) 887–894.
[9] M.N. Chong, B. Jin, C.W.K. Chow, C. Saint, Water Res. 44 (2010) 2997–3027. [40] D.C. Marcano, D.V. Kosynkin, J.M. Berlin, A. Sinitskii, Z.Z. Sun, A. Slesarev, L.
[10] J.G. Hou, Z. Wang, S.Q. Jiao, H.M. Zhu, J. Hazard. Mater. 192 (2011) 1772–1779. Alemany, W. Lu, J.M. Tour, ACS Nano 4 (2010) 4806–4814.
[11] A.L. Linsebigler, G.Q. Lu, J.T. Yates, Chem. Rev. 95 (1995) 735–758. [41] S. Stankovich, D.A. Dikin, G.H.B. Dommett, K.M. Kohlhaas, E.J. Zimney, E.A.
[12] B.X. Li, Y.F. Wang, J. Phys. Chem. C 114 (2010) 890–896. Stach, R.D. Piner, S.T. Nguyen, R.S. Ruoff, Nature 442 (2006) 282–286.
[13] A. Umar, M.S. Chauhan, S. Chauhan, R. Kumar, G. Kumar, S.A. Al-Sayari, S.W. [42] J.T. Paci, T. Belytschko, G.C. Schatz, J. Phys. Chem. C 111 (2007) 18099–18111.
Hwang, A. Al-Hajry, J. Colloid Interface Sci. 363 (2011) 521–528. [43] J.F. Shen, M. Shi, N. Li, B. Yan, H.W. Ma, Y.Z. Hu, M.X. Ye, Nano Res. 3 (2010)
[14] J. Becker, K.R. Raghupathi, J.S. Pierre, D. Zhao, R.T. Koodali, J. Phys. Chem. C 115 339–349.
(2011) 13844–13850. [44] Y. Yang, T.X. Liu, Appl. Surf. Sci. 257 (2011) 8950–8954.
[15] J.Q. Wang, Y.Y. He, J. Tao, J. He, W.J. Zhang, S.J. Niu, Z.Y. Yan, Chem. Commun. [45] E.C. Onyiriuka, Chem. Mater. 5 (1993) 798–801.
46 (2010) 5250–5252. [46] X.Z. Zhou, X. Huang, X.Y. Qi, S.X. Wu, C. Xue, F.Y.C. Boey, Q.Y. Yan, P. Chen, H.
[16] L.W. Zhu, L. Gu, Y. Zhou, S.L. Cao, X.B. Cao, J. Mater. Chem. 21 (2011) 12503– Zhang, J. Phys. Chem. C 113 (2009) 10842–10846.
12510. [47] A. Vinu, D.P. Sawant, K. Ariga, M. Hartmann, S.B. Halligudi, Microporous
[17] K. Keis, C. Bauer, G. Boschloo, A. Hagfeldt, K. Westermark, H. Rensmo, H. Mesoporous Mater. 80 (2005) 195–203.
Siegbahn, J. Photochem. Photobiol., A 148 (2002) 57–64. [48] Y.Y. Li, J.P. Liu, X.T. Huang, G.Y. Li, Cryst. Growth Des. 7 (2007) 1350–1355.
[18] Y. Liu, L.F. Chen, J.C. Hu, J.L. Li, R. Richards, J. Phys. Chem. C 114 (2010) 1641– [49] S. Brunauer, P.H. Emmett, E. Teller, J. Am. Chem. Soc. 60 (1938) 309–319.
1645. [50] T.G. Xu, L.W. Zhang, H.Y. Cheng, Y.F. Zhu, Appl. Catal., B 101 (2011) 382–387.
[19] L.L. Sun, D.X. Zhao, Z.M. Song, C.X. Shan, Z.Z. Zhang, B.H. Li, D.Z. Shen, J. Colloid [51] A. Fujishima, X.T. Zhang, D.A. Tryk, Surf. Sci. Rep. 63 (2008) 515–582.
Interface Sci. 363 (2011) 175–181. [52] P.V. Kamat, Chem. Rev. 93 (1993) 267–300.
[20] J.W. Chiou, S.C. Ray, H.M. Tsai, C.W. Pao, F.Z. Chien, W.F. Pong, C.H. Tseng, J.J. [53] J.N. Clifford, E. Palomares, M.K. Nazeeruddin, R. Thampi, M. Gratzel, J.R.
Wu, M.H. Tsai, C.H. Chen, H.J. Lin, J.F. Lee, J.H. Guo, J. Phys. Chem. C 115 (2011) Durrant, J. Am. Chem. Soc. 126 (2004) 5670–5671.
2650–2655. [54] M. Zhang, C.C. Chen, W.H. Ma, J.C. Zhao, Angew. Chem., Int. Ed. 47 (2008)
[21] L. Yu, S. Yuan, L.Y. Shi, Y. Zhao, J.H. Fang, Microporous Mesoporous Mater. 134 9730–9733.
(2010) 108–114. [55] D. Zhao, C.C. Chen, Y.F. Wang, W.H. Ma, J.C. Zhao, T. Rajh, L. Zang, Environ. Sci.
[22] H.C. Qin, W.Y. Li, Y.J. Xia, T. He, ACS Appl. Mater. Interfaces 3 (2011) 3152– Technol. 42 (2008) 308–314.
3156. [56] Z.G. Xiong, L.L. Zhang, J.Z. Ma, X.S. Zhao, Chem. Commun. 46 (2010) 6099–
[23] K. Woan, G. Pyrgiotakis, W. Sigmund, Adv. Mater. 21 (2009) 2233–2239. 6101.

You might also like