You are on page 1of 8

Journal of Alloys and Compounds 656 (2016) 181e188

Contents lists available at ScienceDirect

Journal of Alloys and Compounds


journal homepage: http://www.elsevier.com/locate/jalcom

Synthesis and catalytic performance of hierarchical TiO2 hollow


sphere/reduced graphene oxide hybrid nanostructures
Xiuying Wang a, b, Jing Wang a, Xiaoli Dong a, *, Feng Zhang a, Linge Ma c, Xu Fei b,
Xiufang Zhang a, Hongchao Ma a
a
School of Light Industry and Chemical Engineering, Dalian Polytechnic University, #1 Qinggongyuan, Dalian 116034, PR China
b
Instrumental Analysis Center, Dalian Polytechnic University, #1 Qinggongyuan, Dalian 116034, PR China
c
Experiment Management Center, National Institute of Clean-and-Low-Carbon Energy, Beijing 102209, PR China

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

Article history: Hierarchical TiO2 hollow spheres partially wrapped with reduced graphene oxide were synthesized
Received 28 July 2015 through a facile three-step wet chemistry process. This method involves preparation of amine-modified
Received in revised form TiO2 hollow spheres, assembly of amine-modified TiO2 hollow spheres and graphene oxide via peptide
26 September 2015
bond (eCONHe) formation, and final hydrothermal reduction of graphene oxide. The prepared hybrid
Accepted 28 September 2015
Available online xxx
composites could be used as photocatalysts to reduce 4-nitroaniline with excellent photocatalytic per-
formance. This property is attributed to the efficient nanostructures of the composites, which result in
obvious improvements in interfacial charge separation and transfer efficiency, reactant adsorptivity, and
Keywords:
TiO2 hollow sphere
light harvesting. This study provides guidance for future designs of new graphene-based composites
Graphene with enhanced photocatalytic reduction performance.
Hydrothermal © 2015 Elsevier B.V. All rights reserved.
Photocatalytic reduction

1. Introduction of TiO2 and RGO have recently gained increased research interest,
since the composites present improved photocatalytic behaviors
Semiconductor photocatalyst research has attracted global [19e29]. For example, Fan et al. prepared TiO2 nanospindle/RGO
attention because of its potential application in solar-to-chemical composites via a hydrothermal method, and the products give an
energy conversion. Titanium dioxide (TiO2), the most common impressive photocatalytic enhancement over pure TiO2 [25].
photocatalyst currently available, offers several advantages, Anderson et al. reported that TiO2/thermally RGO composites
including low cost, high chemical stability, safety to the environ- showed superior photocatalytic activity for phenol degradation
ment, and photocorrosion resistance. However, it suffers from rapid [27]. In general, electrons in TiO2/RGO composites transfer from
charge carrier recombination and low light utilization efficiency. To TiO2 to RGO through interfacial charge-transfers, leading to
overcome these limitations, considerable efforts have been devoted improved separation of electrons and holes [28,29]. This phenom-
to improving the charge separation efficiency and light absorption enon reveals the potential use of TiO2/RGO composites for efficient
of TiO2. One of the most efficient ways to do so is to couple TiO2 photocatalytic selective reduction, which is strongly dependent on
with carbonaceous materials (e.g., carbon dots, carbon fibers, car- photogenerated electron transfer from semiconductors under light
bon nanotubes, graphenes, and graphitic carbon nitride) [1e9] and irradiation.
oxide/sulfide semiconductors (e.g., ZnO, WO3, RuO2, Fe2O3, Cu2O, The photocatalytic performance of TiO2 is well known to be
CdS, SnS2, etc.) [10e17]. Among these additives, reduced graphene strongly dependent on its nanostructures. Among the various
oxide (RGO) shows particular promise in improving the charge nanostructures developed thus far, hierarchical nanostructures are
separation efficiency of TiO2 because of its interesting characteris- especially appealing because of their capacity to combine the light-
tics, including superior electrical conductivity, large surface area scattering ability of large particles with the high surface area of
and modified optoelectronic properties [18e20]. The combinations small particles [30]. Increased light scattering enhances light ab-
sorption and utilization, while large surface areas increase the
adsorptivity of reactants [30]. Larger-sized hierarchical structures
also facilitate the removal and recycling of photocatalysts [19].
* Corresponding author.
E-mail address: dongxl@dlpu.edu.cn (X. Dong).
Micro-sized hollow structure is an example of hierarchical

http://dx.doi.org/10.1016/j.jallcom.2015.09.241
0925-8388/© 2015 Elsevier B.V. All rights reserved.
182 X. Wang et al. / Journal of Alloys and Compounds 656 (2016) 181e188

structures with large size, relatively large surface area, and high 2.2. Characterization
light utilization efficiency [31,32]. Our previous report [33]
demonstrated that micro-sized TiO2 hollow spheres (HSs) exhibit Scanning electron microscopy (SEM) was performed using a
much higher photocatalytic activities than P25. However, we JEOL 7800F microscope, and transmission electron microscopy
believe that the efficiency of these spheres may still be improved to (TEM) was performed on a JEOL 2100 microscope operating at an
a certain extent. Coupling TiO2 HSs to RGO is an effective strategy to acceleration voltage of 200 kV to examine the morphology of the
improve the photocatalytic activity of the resulting material. Zhang samples. Powder X-ray diffraction (XRD) measurements were
et al. prepared TiO2 hollow nanospheres/RGO composites using performed using a Shimadzu XRD-6100 diffractometer with Cu Ka
SiO2 as a template and found that these nanospheres present radiation. Fourier transform infrared spectroscopy (FTIR) was per-
enhanced photocatalytic activity for rhodamine B degradation [34]. formed using a Perkin Elmer Spectrum 10 spectrometer. Photo-
To the best of our knowledge, papers describing micro-sized TiO2 luminescence (PL) spectra were recorded using a Perkin Elmer LS-
HSs coupled to RGO have rarely been reported. Thus, the synthesis 55 spectrometer under 300 nm excitation. UV/visible diffuse
of graphene hybridization with such structures using facile reflection spectra (UVeVis DRS) were obtained over the range of
methods remains challenging endeavors. 300e800 nm via a CARY 100 CONC spectrometer; here, BaSO4 was
Smaller RGO sheets tend to load larger TiO2 particles and used as the reflectance standard. X-ray photoelectron spectroscopy
smaller TiO2 particles tend to load onto larger RGO sheets [35]. (XPS) was performed using a Thermo ESCA Lab250 spectrometer
Thus, considering the relative size of the micro-sized TiO2 HSs and with Al Ka radiation. Raman spectra were recorded on a Horiba HR-
RGO sheets, the final composite material is expected to resemble a 800 system with a 532 nm laser beam. N2 adsorptionedesorption
RGO-wrapped TiO2 HS. To form this structure, strong interactions measurements were performed on a JW-BK222 nitrogen adsorp-
between the components are necessary. Several reports have tion analyzer.
shown that RGO-wrapped metal oxide/sulfide nanocomposites
may be fabricated via electrostatic interactions by using 3- 2.3. Photocatalytic activity
aminopropyltrimethoxysilane (APTMS) [36,37]. In this study,
micro-sized TiO2 HSs partially wrapped with RGO were successfully A 300 W Xe arc lamp (CEL-HXF300) was used as the irradiation
prepared via the formation of peptide bonds (eCONHe) with the source in a typical photocatalytic reaction. Exactly 10 mg of the
assistance of APTMS. Our results show that the hybrid composites photocatalytic materials and 40 mg of ammonium formate were
could be used as photocatalysts to reduce 4-nitroaniline (4-NA) added to 50 mL of 4-nitroaniline solution (10 mg L 1) in a quartz
with excellent photocatalytic performance. The formation mecha- vial. Before illumination, the above suspension was stirred in the
nism of the hybrid structure was investigated, and the reaction dark for 0.5 h to ensure establishment of adsorptionedesorption
scheme of the photocatalytic reduction of 4-NA by TiO2 HS/RGO equilibrium between the sample and the reactant. During the re-
composites was analyzed. action, 5 mL of the sample solution was collected at a specific time
intervals and centrifuged at 12,000 rpm to remove the catalyst
completely. Afterward, the solution was analyzed on a UVevisible
2. Experimental section light (UVevis) spectrophotometer (Shimadzu, UV-1600 PC). The
entire experiment was performed under N2 bubbling.
2.1. Preparation of TiO2 HS/RGO composites
3. Results and discussion
The TiO2 HS/RGO composites were fabricated via a facile self-
assembly method followed by reduction of GO to RGO. TiO2 HSs The morphology of the as-synthesized TiO2 HS was character-
were prepared using a facile hydrothermal method based on pre- ized through SEM, as shown in Fig. 1a and b. The hierarchical ma-
vious literature [33,38]. First, 1.6 g of titanium sulfate [Ti(SO4)2] and terial is approximately 3 mm in diameter and presents a hollow
0.64 g of resorcinol were added to 12 mL of deionized water. After spherical morphology consisting of small-sized nanoparticles, the
stirring for 20 min, 1.6 mL of 35 wt% formaldehyde was quickly formation mechanism of which has been discussed previously in
added to the solution. The mixture was transferred to a 20 mL the literature [33,38]. The TiO2 HS shell is partially wrapped by RGO
Teflon-lined stainless steel autoclave, heated to 85  C for 48 h, and sheets after TiO2eRGO coupling (Fig. 1c and d). This SEM result
then washed several times with deionized water. After drying at suggests tight interfacial contact between TiO2 HS and the RGO
85  C for a few hours, the intermediate product was annealed at sheet, which promotes rapid charge transfer across the interface.
450  C for 3 h in air to produce TiO2. Hydrothermal treatment barely affected the morphology of the
To synthesize the TiO2 HS/RGO composites, 200 mg of TiO2 HS TiO2 HSs, thereby demonstrating that the simple synthesis
was added into 100 mL of ethanol with stirring for 20 min. Then, approach facilitates maintenance of the original TiO2 HS
0.5 mL of APTMS was injected into this solution, and the mixture morphology. To confirm the synthesized sample texture, TEM
was heated at 60  C for 4 h to form amine-modified TiO2 HS (ATHS). characterization was performed (Fig. 2). The low magnification TEM
Thereafter, ATHS was collected, washed with ethanol for several image (Fig. 2a) cannot reveal the surface change in TiO2 HSs
times, and then dried at 60  C for a few hours. A GO (prepared by a because of the considerable difference in size of the RGO sheets and
modified Hummers method [39]) suspension (0.5 mg mL 1) was the TiO2 HSs. However, high magnification TEM images (Fig. 2bef)
added to an ATHS dispersion at GO to TiO2 weight ratios of 0.01:1, clearly show that nanoparticles on the TiO2 HSs are anchored on
0.03:1, and 0.05:1 under vigorous stirring. After mixing for 30 min, the two-dimensional (2D) transparent sheet structure with wrin-
the mixture was centrifuged and denoted as ATHSeGO. ATHSeGO kles; this finding implies that considerable interfacial contact be-
with different weight ratios of GO was subsequently dispersed in tween TiO2 HS and RGO sheet was achieved. The SEM and TEM
17 mL of deionized water and then transferred to a 20 mL Teflon- studies thus far show intimate interfacial contact between TiO2 HS
lined stainless steel autoclave. After heating in an electric oven at and the RGO sheet, which could result in better separation of
120  C for 12 h, the autoclave was cooled down naturally. Finally, photoinduced charge carriers and more efficient charge transfer
the products were collected, washed with deionized water, and within the composite structure.
then dried at 60  C for a few hours to obtain the TiO2 HS/RGO The typical XRD patterns of the TiO2 HS/RGO composites with
composites with 1%, 3%, and 5% weight addition ratios of RGO. various amounts of RGO are shown in Fig. 3. No significant
X. Wang et al. / Journal of Alloys and Compounds 656 (2016) 181e188 183

Fig. 1. SEM images of (aeb) pure TiO2 HS and (ced) TiO2 HS/3% RGO composite.

difference was observed among the patterns obtained, and all of attributed to the Raman modes of B1g(1), A1g(1)/B1g(2), and Eg(2)
the peaks could readily be indexed to anatase TiO2. Sharp peaks of anatase TiO2, which is consistent with previous reports [40,42].
located at 25.3 , 37.8 , 48.1, 53.9 , 55.0 , 62.7, 68.7, 70.3 , and This result also confirms the successful preparation of TiO2 HS/RGO
75.1 were respectively indexed to the (101), (004), (200), (105), composites.
(211), (204), (116), (220), and (215) crystal planes of anatase TiO2 The synthesis of TiO2 HS/RGO composites (Scheme 1) typically
(JCPDS Card No. 21-1272). No diffraction peak at 11.0 , which is involves three steps: (1) fabrication of amine-modified TiO2 HS by
ascribed to GO (Supporting Information Fig. S1), was observed, using APTMS, which is widely used to introduce amine groups
which implies that GO was reduced. However, no characteristic (eNH2) to various nanomaterials [26,34]; (2) assembly of TiO2 HS
RGO diffraction peaks were detected in the patterns obtained. This with GO via peptide bond (eCONHe) formation between eCOOH
phenomenon could be attributed to the low RGO weight in the and eNH2; and (3) reduction of GO to RGO. This process may be
samples and overlapping of the main characteristic peak of RGO (at demonstrated by the IR results. As shown in Fig. 6, TiO2 HSs present
25.0 ) with the (101) peak of anatase TiO2 [26]. a peak at 801 cm 1, which can be assigned to the stretching vi-
To better understand the structures for the materials, N2 brations of TieOeTi bonds [43]. Bands at 1135 and 1054 cm 1,
adsorptionedesorption isotherms were performed. Fig. 4 gives the which correspond to CeO stretching vibrations, as well as bands at
N2 adsorptionedesorption isotherms and the corresponding pore 2925 (asymmetric stretching), 2863 (symmetric stretching), 1463,
size distribution for TiO2 HSs and TiO2 HS/3% RGO composites. Both and 1387 cm 1 (scissoring, deformation), which correspond to CH2
of the N2 adsorptionedesorption isotherms are typical type IV groups, were also observed. These peaks are attributed to the
curves, indicating the existence of mesopores in the materials. The preparation of TiO2 HS via calcination of the phenol-formaldehyde
mesopores in both TiO2 HSs and TiO2 HS/3% RGO composites are resin template [33,38]. When TiO2 HS was treated with APTMS, two
about 4 nm in diameter. The BrunauereEmmetteTeller (BET) sur- new peaks appeared. The characteristic peak at 1127 cm 1 could be
face area of TiO2 HSs and TiO2 HS/3% RGO composites are attributed to SieO stretching vibrations, and the typical peak at
81.2 m2 g 1 and 84.2 m2 g 1, respectively. On the basis of above 1531 cm 1 could correspond to amine groups; these peaks indicate
studies, it is concluded that TiO2 HS/3% RGO composites remain the that the amine group (eNH2) was successfully introduced to the
mesoporous hollow nanostructures of raw TiO2 HSs. TiO2 HS. The stretching vibrations of C]O (1720 cm 1) were
Raman spectroscopy is a popular and powerful tool used to observed in bare GO, and this peak indicates the presence of
characterize the crystalline quality for carbonaceous materials. The carboxyl groups (eCOOH). The distinct ATHSeGO peak at
intensity ratio of the D (~1350 cm 1) and G (~1590 cm 1) bands, ID/ 1585 cm 1 was assigned to the presence eCONHe groups, and the
IG, can be used to indicate the relative concentration of local defects peak ascribed to carboxyl groups became much weaker than that of
or disorders (particularly sp3-hybridized defects) compared with bare GO. These phenomena demonstrate the reaction between the
the sp2-bonded carbon atoms in a 2D hexagonal lattice [40]. Fig. 5 carboxyl groups (eCOOH) of GO and amine groups (eNH2) of TiO2
shows the Raman spectra of GO and the TiO2 HS/RGO composites. HS. In the IR spectrum of the TiO2 HS/RGO composites obtained
The ID/IG ratio for GO was 0.97, while that for the TiO2 HS/RGO after hydrothermal treatment, the peak ascribed to the carboxyl
composites was 1.05; this result suggests the reduction of GO to groups disappeared, which indicates complete reduction of GO to
RGO and a slight increase in the defect concentration in the com- RGO.
posites [41]. Peaks located at 396, 517, and 639 cm 1 were To confirm RGO formation during the hydrothermal process,
184 X. Wang et al. / Journal of Alloys and Compounds 656 (2016) 181e188

Fig. 2. TEM images of TiO2 HS/3% RGO composite.

Fig. 3. XRD patterns of (a) pure TiO2 HS, (b) TiO2 HS/1% RGO composite, (c) TiO2 HS/3% Fig. 4. N2 adsorptionedesorption isotherms and pore size distribution (inset) for TiO2
RGO composite and (d) TiO2 HS/5% RGO composite. HS and TiO2 HS/3% RGO composite.
X. Wang et al. / Journal of Alloys and Compounds 656 (2016) 181e188 185

Fig. 5. Raman spectra of (a) GO and (b) TiO2 HS/3% RGO composite.

Scheme 1. Schematic illustration of the formation of TiO2 HS/RGO composites.

XPS measurements were conducted (Fig. 7). The high-resolution C an electron transfer channel, which improves photoinduced charge
1s spectrum of GO (Fig. 7a) displayed three peaks at 284.7, 286.8, separation rates during the photocatalytic process [40].
and 288.9 eV, which were respectively assigned to sp2-hybridized The photocatalytic activities of TiO2 HS/RGO composites were
carbon (C]C), hydroxyl carbon (CeO), and carboxyl carbon (O] evaluated by reducing 4-nitroaniline (4-NA) in water under simu-
CeO) [44]. The significant loss of oxygen-containing functional lated sunlight irradiation. In the photocatalytic reaction system,
groups in the TiO2 HS/RGO composites was further observed ammonium formate was used as a photogenerated hole quencher
(Fig. 7c). These XPS studies suggest that GO is sufficiently reduced and graphene photodegradation inhibitor [36]. N2 purging pro-
to RGO after hydrothermal treatment, which is in considerable vided an anaerobic atmosphere to ensure that the aromatic nitro
agreement with the IR analysis. The survey spectrum of the TiO2 organics do not undergo oxidation. Fig. 8 shows the photocatalytic
HS/RGO composites showed the presence of Ti 2p, C 1s, and O 1s, as activities of P25, TiO2 HSs, and TiO2 HS/RGO composites with
shown in Fig. 7b, thereby implying that TiO2 HS was successfully various amounts of RGO. TiO2 HSs exhibited higher photocatalytic
coupled with RGO. Comparison of the high-resolution Ti 2p spectra activities than P25, which is attributed to the unique hierarchical
of the TiO2 HS/RGO composites and pure TiO2 HSs (Fig. 7d) showed nanostructure of the TiO2 HSs, similar to our previous report [33].
a slight shift, indicating strong interactions at the TiO2 HS and RGO Furthermore, all of the photocatalytic activities of the TiO2 HS/RGO
interface. This intense interaction likely results in the formation of composites were considerably higher than that of the TiO2 HSs.
186 X. Wang et al. / Journal of Alloys and Compounds 656 (2016) 181e188

Fig. 8. Photocatalytic selective reduction of 4-NA to PPD over P25, bare TiO2 HS, and
Fig. 6. FTIR spectra of (a) GO, (b) pure TiO2 HS, (c) amine-modified TiO2 HS (ATHS), (d) the TiO2 HS/RGO composites with different amounts of RGO under simulated sunlight
ATHSeGO and (e) TiO2 HS/3% RGO composite. irradiation.

Among the materials studied, the TiO2 HS/3% RGO composites


exhibited the highest photocatalytic activity for 4-NA reduction. characterization was performed. DRS results (Fig. 9) revealed that
The product for this photocatalytic reaction was characterized by introduction of RGO improves the light absorption intensity in the
absorption spectra. As shown in Fig. S2, the absorption bands at UV and visible light regions as compared with that obtained on the
240 nm and 305 nm were gradually increased along with the bare TiO2 HSs. Absorption intensity also increased with increasing
increasing irradiation time, while the absorption band at 380 nm RGO loading. Adsorptivity measurements (Supporting Information
was obviously decreased. According to previous report [45], the Fig. S3) suggested that introduction of RGO could increase the
absorption bands at about 240 nm and 305 nm were attributed to adsorptivity of reactants because of pep conjugation between ar-
p-phenylenediamine (PPD). omatic nitro organics and the aromatic regions of RGO [29], rather
To understand differences in the photocatalytic performance of than negligible variations in surface area resulting from the low
the TiO2 HSs and TiO2 HS/RGO composites better, further RGO content. This increased TiO2 HS/RGO composite adsorptivity

Fig. 7. XPS spectra of (a) C 1s for GO, (b) TiO2 HS/3% RGO composite, (c) C 1s for TiO2 HS/3% RGO composite, (d) Ti 2p for TiO2 HS/3% RGO composite and TiO2 HS, respectively.
X. Wang et al. / Journal of Alloys and Compounds 656 (2016) 181e188 187

photogenerated electronehole pairs is efficiently inhibited after


coupling TiO2 HS with RGO. These results further imply that effi-
cient electron transfer for reduction of aromatic nitro organics to
amino organics is achieved over the TiO2 HS/RGO composites,
which is particularly important in the reduction reaction driven by
energized electrons.
Based on the results described above, incorporation of RGO into
the TiO2 HSs could significantly improve light absorption, adsorp-
tivity toward the reactants, and transfer of photogenerated elec-
trons. These positive factors result in significant enhancements in
the photocatalytic activity of the TiO2 HS/RGO composites for
reducing 4-NA. A schematic model for the photocatalytic reduction
of 4-NA to PPD over TiO2 HS/RGO composites is proposed in
Scheme 2. Under simulated sunlight irradiation, TiO2 is photoex-
cited to generate electronehole pairs. The photoexcited electrons,
after migrating from the bulk to the TiO2 surface, are transferred to
Fig. 9. Diffuse reflectance spectra of TiO2 HS and the TiO2 HS/RGO composites with the RGO sheets by a percolation mechanism [47], while the pho-
different amounts of RGO. togenerated holes are trapped by the quenching agent, ammonium
formate. Since the 2D planar p-conjugation structure endows RGO
with excellent electron conductivity, the photoexcited electrons
could easily transfer to the RGO framework and drive the reduction
reaction of 4-NA adsorbed on the TiO2 HS/RGO composites. The
special hollow structures of the catalysts improve the efficiency of
the photoreduction reaction because of their ability to multiscatter
incident light, leading to higher light utilization. The TiO2 HS/5%
RGO composites have higher light absorption and reactant
adsorptivity but lower photoactivity than the TiO2 HS/3% RGO
composites. This result is attributed to the excessive addition of
RGO, which decreases the amount of the primary photoactive
ingredient, TiO2, and reduces the contact surface of the component
with light irradiation [48]. Hence, introduction of an appropriate
amount of RGO to TiO2 HS is crucial to achieve optimal synergistic
effects between RGO and the TiO2 HSs and obtain the best photo-
catalytic performance of the TiO2/RGO composites.
The recycling behavior of TiO2 HS/3% RGO composites was also
tested since it is important for the practical application. The recy-
cling test was performed with five cycles. As shown in Fig. 11, it is
Fig. 10. PL spectra of TiO2 HS and TiO2 HS/3% RGO composite.
interesting that the photocatalytic activity of TiO2 HS/3% RGO
composites in the second cycle is slight higher than that in the first
promotes the photocatalytic reduction of aromatic nitro organics cycle. This phenomenon is also found on ordered mesoporous TiO2
on the photocatalyst surface. PL spectra have been widely used to for photocatalytic hydrogen generation recycling test [49]. It maybe
study the separation efficiency of electronehole pairs photo- ascribed to the mesoporous feature of TiO2 HSs, which can lead to
generated from semiconductor/graphene composites [24,37,46]. adsorbing some species. After the first light irradiation, these spe-
The PL results in Fig. 10 clearly show that introduction of RGO leads cies may disappear, resulting in exposure of more active sites. We
to TiO2 HS PL quenching, which indicates that recombination of will further study on this issue in our next work. From the third
cycle, the photocatalytic activity of TiO2 HS/3% RGO composites
begins to decrease gradually. After five cycles, the photocatalytic
activity of TiO2 HS/3% RGO composites is still higher than that of
pure TiO2 HSs. Furthermore, the nanostructures of TiO2 HS/3% RGO
composites barely changed even after five cycles (Supporting In-
formation Fig. S4), implying the good stability of the composites
during photocatalytic reaction.

4. Conclusions

In summary, we have demonstrated an efficient approach to


partially wrap hierarchical TiO2 HSs with RGO via peptide bond
(eCONHe) formation. The obtained TiO2 HS/RGO composites
exhibited high photocatalytic activity for the reduction of 4-NA,
which is ascribed to increase in light absorption, improvement in
reactant adsorptivity, and enhancement of photogenerated elec-
Scheme 2. Schematic illustration of the photocatalytic reduction process of 4-NA to
tron transfer. Our approach for preparing TiO2 HS/RGO composites
PPD over the TiO2 HS/RGO composites under simulated sunlight irradiation with the presents new possibilities for exploiting suitable photocatalysts for
addition of ammonium formate as quencher for photogenerated holes and N2 purge. selective organic reduction transformation.
188 X. Wang et al. / Journal of Alloys and Compounds 656 (2016) 181e188

Fig. 11. Recycling photocatalytic reduction of 4-NA over TiO2 HS/3% RGO composite under simulated sunlight irradiation.

Acknowledgments [19] G. Lui, J.Y. Liao, A.S. Duan, Z.S. Zhang, M. Fowler, A.P. Yu, J. Mater. Chem. A 1
(2013) 12255e12262.
[20] H. Liu, T. Lv, Z.F. Zhu, J. Mol. Catal. A Chem. 404 (2015) 178e185.
The research was supported by the National Natural Science [21] J.F. Shen, B. Yan, M. Shi, H.W. Ma, N. Li, M.X. Ye, J. Mater. Chem. 21 (2011)
Foundation of China (Grant No. 21476033, 21401016, and 3415e3421.
21204007), Cultivation Program for Excellent Talents of Science and [22] X. Pan, Y. Zhao, S. Liu, C.L. Korzeniewski, S. Wang, Z.Y. Fan, ACS. Appl. Mater.
Interfaces 4 (2012) 3944e3950.
Technology Department of Liaoning Province (No. 201402610) and [23] X.L. Gou, Y.L. Cheng, B. Liu, B.J. Yang, X.B. Yan, Eur. J. Inorg. Chem. 13 (2015)
Science and Technology Research Project of the Education 2222e2228.
Department of Liaoning Province (No. L2014225). [24] L. Yuan, Q.Q. Yu, Y.H. Zhang, Y.J. Xu, RSC Adv. 4 (2014) 15264e15270.
[25] H.T. Wu, J. Fan, E.Z. Liu, X.Y. Hu, Y.N. Ma, X. Fan, Y.Y. Li, C.N. Tang, J. Alloys
Compd. 623 (2015) 298e303.
Appendix A. Supplementary data [26] J.S. Lee, K.H. You, C.B. Park, Adv. Mater. 24 (2012) 1084e1088.
[27] H. Adamu, P. Dubey, J.A. Anderson, Chem. Eng. J. 284 (2016) 380e388.
[28] S. Umrao, S. Abraham, F. Theil, S. Pandey, V. Ciobota, P.K. Shukla, C.J. Rupp,
Supplementary data related to this article can be found at http:// S. Chakraborty, R. Ahuja, J. Popp, B. Dietzek, A. Srivastava, RSC Adv. 4 (2014)
dx.doi.org/10.1016/j.jallcom.2015.09.241. 59890e59901.
[29] T.W. Lu, R.B. Zhang, C.Y. Hu, F. Chen, S.W. Duo, Q.H. Hu, Phys. Chem. Chem.
Phys. 15 (2013) 12963e12970.
References [30] Y. Li, Z.Y. Fu, B.L. Su, Adv. Funct. Mater. 22 (2012) 4634e4667.
[31] S. Dadgostar, F. Tajabadi, N. Taghavinia, ACS Appl. Mater. Interfaces 4 (2012)
[1] G.W. Cui, W.L. Wang, M.Y. Ma, M. Zhang, X.Y. Xia, F.Y. Han, X.F. Shi, Y.Q. Zhao, 2964e2968.
Y.B. Dong, B. Tang, Chem. Commun. 49 (2013) 6415e6417. [32] J.F. Qian, P. Liu, Y. Xiao, Y. Jiang, Y.L. Cao, X.P. Ai, H.X. Yang, Adv. Mater. 21
[2] M.Y. Sun, S.N. Qu, W.Y. Ji, P.T. Jing, D. Li, L. Qin, J.S. Cao, H. Zhang, J.L. Zhao, (2009) 3663e3667.
D.Z. Shen, Phys. Chem. Chem. Phys. 17 (2015) 7966e7971. [33] J. Wang, X.Y. Wang, X.L. Dong, X.F. Zhang, H.C. Ma, X. Fei, RSC Adv. 4 (2014)
[3] J.Q. Pan, Y.Z. Sheng, J.X. Zhang, J.M. Wei, P. Huang, X. Zhang, B.X. Feng, J. Mater. 59503e59507.
Chem. A 2 (2014) 18082e18086. [34] J. Zhang, Z.P. Zhu, Y.P. Tang, X.L. Feng, J. Mater. Chem. A 1 (2013) 3752e3756.
[4] Z.F. Wang, K. Yoshinaga, X.R. Bu, M. Zhang, J. Hazard. Mater. 290 (2015) [35] H.I. Kim, G.H. Moon, D. Monllor-Satoca, Y. Park, W. Choi, J. Phys. Chem. C 116
134e141. (2012) 1535e1543.
[5] N. Mohaghegh, M. Faraji, F. Gobal, M.R. Gholami, RSC Adv. 5 (2015) [36] Z. Chen, S.Q. Liu, M.Q. Yang, Y.J. Xu, ACS Appl. Mater. Interfaces 5 (2013)
44840e44846. 4309e4319.
[6] M.Q. Yang, N. Zhang, Y.J. Xu, ACS Appl. Mater. Interfaces 5 (2013) 1156e1164. [37] J.Y. Zhu, J.H. He, ACS Appl. Mater. Interfaces 4 (2012) 1770e1776.
[7] K.M. Cho, K.H. Kim, H.O. Choi, H.T. Jung, Green Chem. 17 (2015) 3972e3978. [38] F. Zhang, Y. Zhang, S.Y. Song, H.J. Zhang, J. Power Sources 196 (2011)
[8] J. Hu, H.S. Li, Q. Wu, Y. Zhao, Q.Z. Jiao, Chem. Eng. J. 263 (2015) 144e150. 8618e8624.
[9] J.Y. Lei, Y. Chen, F. Shen, L.Z. Wang, Y.D. Liu, J.L. Zhang, J. Alloys Compd. 631 [39] N.I. Kovtyukhova, P.J. Ollivier, B.R. Martin, T.E. Mallouk, S.A. Chizhik,
(2015) 328e334. E.V. Buzaneva, A.D. Gorchinskiy, Chem. Mater. 11 (1999) 771e778.
[10] F. Kayaci, S. Vempati, C. Ozgit-Akgun, I. Donmez, N. Biyikliab, T. Uyar, Nano- [40] X.R. Cao, G.H. Tian, Y.J. Chen, J. Zhou, W. Zhou, C.G. Tian, H.G. Fu, J. Mater.
scale 6 (2014) 5735e5745. Chem. A 2 (2014) 4366e4374.
[11] W. Niu, G. Wang, X.D. Liu, J. Tang, X.G. Bi, Int. J. Electrochem. Sci. 10 (2015) [41] J. Zhou, G.H. Tian, Y.J. Chen, X.Y. Meng, Y.H. Shi, X.R. Cao, K. Pan, H.G. Fu, Chem.
8513e8521. Commun. 49 (2013) 2237e2239.
[12] M.T. Uddin, O. Babot, L. Thomas, C. Olivier, M. Redaelli, M.D. Arienzo, [42] Q.J. Xiang, J.G. Yu, M. Jaroniec, Nanoscale 3 (2011) 3670e3678.
F. Morazzoni, W. Jaegermann, N. Rockstroh, H. Junge, T. Toupance, J. Phys. [43] H. Zhang, X.J. Lv, Y.M. Li, Y. Wang, J.H. Li, ACS Nano 4 (2010) 380e386.
Chem. C 119 (2015) 7006e7015. [44] L.C. Liu, Z. Liu, A.N. Liu, X.R. Gu, C.Y. Ge, F. Gao, L. Dong, ChemSusChem 7
[13] Z.Y. Lin, P. Liu, J.H. Yan, G.W. Yang, J. Mater. Chem. A 3 (2015) 14853e14863. (2014) 618e626.
[14] L.M. Liu, W.Y. Yang, W.Z. Sun, Q. Li, J.K. Shang, ACS Appl. Mater. Interfaces 7 [45] C.Y. Chiu, P.J. Chung, K.U. Lao, C.W. Liao, M.H. Huang, J. Phys. Chem. C 116
(2015) 1465e1476. (2012) 23757e23763.
[15] K. Ma, O. Yehezkeli, D.W. Domaille, H.H. Funke, J.N. Cha, Angew. Chem. Int. Ed. [46] Y.H. Zhang, N. Zhang, Z.R. Tang, Y.J. Xu, Phys. Chem. Chem. Phys. 14 (2012)
54 (2015) 11490e11494. 9167e9175.
[16] G.H. Li, L. Wu, F. Li, P.P. Xu, D.Q. Zhang, H.X. Li, Nanoscale 5 (2013) 2118e2125. [47] X. Wang, L.J. Zhi, K. Müllen, Nano Lett. 8 (2008) 323e327.
[17] Z.Y. Zhang, C.G. Shao, X.H. Li, Y.Y. Sun, M.Y. Zhang, J.B. Mu, P. Zhang, Z.C. Guo, [48] Y.H. Zhang, Z.R. Tang, X.Z. Fu, Y.J. Xu, ACS Nano 4 (2010) 7303e7314.
Y.C. Liu, Nanoscale 5 (2013) 606e618. [49] W. Zhou, W. Li, J.Q. Wang, Y. Qu, Y. Yang, Y. Xie, K.F. Zhang, L. Wang, H.G. Fu,
[18] D. Chen, H. Zhang, Y. Liu, J.H. Li, Energy Environ. Sci. 6 (2013) 1362e1387. D.Y. Zhao, J. Am. Chem. Soc. 136 (2014) 9280e9283.

You might also like