DOI:
10.1039/C5RA26096E
(Paper)
RSC Adv., 2016,
6, 2479-2488
Enhanced hydrogen production by carbon-doped TiO2 decorated with reduced graphene oxide (rGO) under visible light irradiation†
Received
7th December 2015
, Accepted 17th December 2015
First published on 22nd December 2015
Abstract
Enhancing visible light utilization by photocatalysts, avoiding electron–hole recombination, and facilitating charge transfer are three major challenges to the success of sustainable photocatalytic systems. In our study, carbon-doped TiO2 was synthesized with decoration of reduced graphene oxide (C-TiO2/rGO) to form a hybrid nanocomposite that exhibits excellent photocatalytic activity and longevity. Morphology, chemical and colloidal stability, crystallinity, surface compositions and band structures were systematically assessed. The results revealed that the hybrid C-TiO2/rGO had a band gap of 2.2 ± 0.2 eV and crystallite sizes of 0.9–2 nm in diameter. Transmission electron microcopy (TEM) images showed that C-TiO2 particles attached to the carbon sheet of rGO. Under irradiation of 135 mW cm−2 at 400–690 nm with methanol as electron donor, C-TiO2 and C-TiO2/rGO yielded incredibly high H2 production rates of 0.67 ± 0.12 to 1.50 ± 0.2 mmol g−1 h−1, respectively, which were greater than those of other titanium hybrid catalysts such as C-TiO2/Pt. rGO not only greatly improved the photocatalytic activity but also led to greater stability of H2 production compared to C-TiO2. This work lays groundwork toward the design of novel visible light-driven photocatalytic systems for harnessing solar energy and environmental applications.
1. Introduction
Visible light-driven photocatalytic hydrogen (H2) production presents an appealing approach to harness solar energy and potentially tackles many environmental issues such as wastewater treatment.1–3 Efficient photocatalytic H2 production requires efficient and stable photocatalysts that could maintain excellent photocatalytic activity and longevity. Most traditional photocatalysts (e.g., TiO2) have relatively large band gaps (>3 eV) and thus can only capture ultraviolet (UV) irradiation,4 which only accounts for about 5% of solar irradiation. To effectively capture visible light from the solar spectrum, photocatalytic materials should have band gaps in the 1.6–1.9 eV range.5 Thus, band engineering is the typical strategy to broaden visible light utilization, while retaining excellent electron–hole separation and stability of photocatalytic reactions.6
Elemental doping is one of the band engineering methods that incorporate foreign metal and nonmetal ions into catalyst synthesis.7–9 Metal ions such V, Ni, Cr, Au, Ag, Mo, Fe, Sn, and Mn were previously used.10–13 Non-metals such as C, N, and S have been widely used to shift the valence band edge upward and thus narrow band gaps.14–16 For instance, Asahi et al. studied the substitutional doping of C, N, F, P, and S for O in anatase TiO2.17 Chen and co-workers used X-ray photoelectron spectroscopy (XPS) to demonstrate the presence of additional electronic states above the valence band edge of C-, N-, and S-doped TiO2.8,18 The additional electron density of states explained the red shifted absorption as observed in the “shoulder” and “tail-like” features in the UV-vis spectra of these modified photocatalysts. Particularly, carbon doping has proved to be effective in narrowing band gap of TiO2.8,14,17,19,20 Particularly, carbon-doped TiO2 (C-TiO2) was shown to be the best in narrowing energy bandgap.21 C-TiO2 nanotubes also displayed a high photoactivity for water splitting and utilization of solar energy up to the visible to infrared region, due to band gap reduction and the new intragap band formation. However, the current challenges for these doped Ti-based photocatalysts are the rapid electron–hole recombination,22 followed by quantum yield,23 stability,23,24 and synthesis cost. One of the strategies for preventing the recombination is to build heterojunctions containing anion-doped titania and to coat or decorate metal or semiconductor nanomaterials or organic dye photosensitizers to facilitate electron–hole separation.25,26
Carbonaceous nanomaterials such as carbon nanotubes,27 graphene or graphene oxide (GO)28,29 have been integrated in the synthesis of novel nanocomposites with improved performance of photocatalysts,30,31 fuel cells,32 and batteries,33 where charge separation and electron transport are dominant operating principles. Graphene, for instance, provides a unique two-dimensional (2-D) platform for electron transport with a high specific surface area (up to 2630 m2 g−1) with exceptional chemical and mechanical stability, electrical conductivity and electron mobility (∼200
000 cm2 V−1 s−1).34 Anchoring photocatalysts onto graphitic nanostructures could also cause a red-shift in the absorption spectrum,35,36 as well as prevention of catalyst aggregation.37 Reduced graphene oxide (rGO) greatly restores a high degree of the sp2 bonding structure inherent to pristine graphene and yields a conductive 2-D carbon mat that can shuttle charges between the active nanostructured materials.38 Previous work employed rGO as a supporting matrix for TiO2 and achieved good photocatalytic performance under UV-vis irradiation.39 However, the roles of rGO in visible-light-driven photocatalytic H2 evolution as well as in photochemical or colloidal stability of hybrid nanostructures of doped titania have not been well elucidated (see the summary of recent studies in Table 1). As opposed to doping rare earth elements, noble or transition metals such as Pt, Pd, Ru and Rh, non-metal co-catalysts (e.g., nitrogen-doped GO40) are shown to be more sustainable because carbon or nitrogen elements are earth abundant,41 and some heavy metal dopants such as Cr and Cd are toxic and harmful to ecosystems and the environment, which limits their wide application.42
Table 1 Comparison of literature on visible light-driven H2 production by titania based catalysts with or without graphenea
Catalyst |
Light source (nm) |
Electron donor |
Light intensity (mW cm−2) |
Average H2 productivity |
Longevity of each reaction cycle (h) |
Ref. |
μmol L−1 h−1 |
mmol g−1 h−1 |
μmol m−2 h−1 |
G denotes graphene. rGO denotes reduced graphene oxide. GSs denotes graphene sheets. |
GSs/TiO2 |
Xe lamp |
Na2S/Na2SO3 |
150 |
43 |
0.086 |
NA |
NA |
39 |
TiO2/G |
UV-vis |
Na2S/Na2SO3 |
80 |
27 |
0.108 |
NA |
5 |
25 |
TiO2/G |
Xe lamp |
Methanol |
NA |
11 133 |
6.68 |
119.07 |
9 |
43 |
TiO2/rGO |
UV-vis |
Methanol |
NA |
740 |
0.74 |
NA |
3 (4 runs) |
44 |
TiO2/rGO |
>320 |
Methanol |
205 |
NA |
NA |
NA |
NA |
45 |
Cu2O-TiO2/rGO |
Xe lamp |
Glycerol |
NA |
NA |
110.968 |
NA |
8 h |
46 |
Eosin Y–Pt–N-TiO2 |
>420 |
TEOA |
NA |
1000 |
0.8 |
8.7 |
NA |
47 |
g-C3N4/N-TiO2 nanofibers |
Xe lamp |
Methanol |
NA |
3572.52 |
8.93 |
NA |
NA |
48 |
Ce–N-TiO2 |
Tungsten halogen lamp |
Methanol |
NA |
8240 |
1.03 |
NA |
NA |
49 |
Ce–B-TiO2 |
Tungsten halogen lamp |
Methanol |
NA |
6334.5 |
0.79 |
NA |
NA |
49 |
Ce–C-TiO2 |
Tungsten halogen lamp |
Methanol |
NA |
5407.5 |
0.68 |
NA |
NA |
49 |
Ce–S-TiO2 |
Tungsten halogen lamp |
Methanol |
NA |
1287.5 |
0.16 |
NA |
NA |
49 |
N-TiO2/N-GO |
Hg lamp |
Methanol |
NA |
906.18 |
0.996 |
NA |
NA |
50 |
Au/N-TiO2 |
Xe lamp |
Methanol |
NA |
8252 |
8.252 |
NA |
NA |
51 |
N-TiO2 |
Xe lamp |
Methanol |
NA |
2980 |
2.98 |
NA |
NA |
52 |
Pt/N-TiO2 |
Solar simulator |
Methanol |
50 |
205.8 |
0.2 |
NA |
NA |
53 |
N-TiO2 |
Xe lamp |
NA |
NA |
NA |
0.00003 |
NA |
NA |
54 |
Pt/N-TiO2 |
Xe-arc lamp |
Methanol |
NA |
1036.36 |
0.57 |
NA |
NA |
55 |
C,N-TiO2 |
Xe lamp |
Methanol |
NA |
105.55 |
0.081 |
0.89 |
NA |
56 |
W/N-TiO2 |
Vis |
Ethanol |
NA |
43.88 |
0.0176 |
NA |
NA |
57 |
C-TiO2/rGO |
Xe lamp |
Methanol |
135 |
495 |
1.5 |
NA |
20 |
This study |
C-TiO2 |
Xe lamp |
TEOA |
135 |
16.25 |
0.049 |
NA |
NA |
This study |
C-TiO2/rGO |
Xe lamp |
TEOA |
135 |
21.67 |
0.066 |
NA |
NA |
This study |
In this study, C-TiO2 was synthesized and further anchored to rGO to create hybrid nanocomposites (C-TiO2/rGO) for H2 production under visible light irradiation. TiO2 was used as a base photocatalytic material mainly because of its excellent photoreactivity, chemical stability (no dissolution or ion release), facile synthesis, low cost, and relative nontoxicity. Material properties such as morphology, chemical and colloidal stability, crystallinity, surface compositions and band structure were systematically assessed. We compared the photocatalytic H2 production rates for C-TiO2 and C-TiO2/rGO and further evaluated the longevity of C-TiO2/rGO. The role of rGO in improving photocatalytic activity and mechanisms of photocatalytic reactions were analyzed to provide insight into the design of novel visible light-responsive photocatalysts.
2. Experimental
2.1. Synthesis of C-TiO2/rGO
C-TiO2 was synthesized via a reported one-pot solvothermal process with minor modifications to create nano-sized C-TiO2.20 Briefly, 2 ml of titanium isopropoxide was mixed with 60 ml of anhydrous acetone. The mixture was stirred at ambient conditions for 30 min and then transferred to a 125 ml Teflon-lined stainless-steel autoclave (Parr Instrument Co.) for heating at 200 °C for 12 h. After synthesis, the white precipitate was separated by centrifugation (2000 × g for 10 min) and washed several times with deionized (DI) water. The precipitate then went through 2 h calcination in a furnace at 200 °C subsequently. Compared to other carbon doping methods, such as exposing TiO2 to CO2 or air at high temperatures,58–63 the one-pot hydrothermal method we employed relied on the aldol condensation reaction between titanium alkoxide and acetone,20 which was reported to lead to high surface area, tunable pore and grain sizes, and high crystallinity elsewhere.64
GO was synthesized by the simplified Hummer's method,65 which is described in details in ESI.† C-TiO2/rGO was obtained also via hydrothermal synthesis with a rGO loading ratio of 2%, which was optimized in previous studies.25,41 Briefly, 10 mg GO was dispersed in a solution of DI water (20 mL) and ethanol (30 mL) by 100 W ultrasonic treatment for 1 h. Then, 0.5 g C-TiO2 was added to the obtained GO suspension and stirred for another 2 h to get a homogeneous suspension. The suspension was then placed in a 150 mL Teflon sealed autoclave and heated at 120 °C for 3 h to simultaneously achieve the reduction of GO and the deposition of C-TiO2 on the rGO sheet. Finally, the resulting nanocomposite was purified and separated by filtration (0.2-micron Nylon Millipore filter), followed by rigorous DI water rinsing and air drying at room temperature. rGO was obtained by the same procedure without adding C-TiO2.
2.2. Characterization
Morphology and size distribution were determined by a Hitachi H-7500 transmission electron microscope (TEM). Hydrodynamic particle size distribution (PSD) and zeta potential were measured by dynamic light scattering (DLS) on a Zetasizer nano ZS instrument (Malvern Instruments, UK). X-ray diffraction (XRD) was recorded for the crystallography using a Philips PW3040 X-ray diffractometer. Surface compositions and crystalline phases were assessed by Fourier transform infrared (FTIR) and Raman Spectrometers. FTIR was performed on a Nicolet Thermo Electron FTIR spectrometer combined with a MIRacle attenuated total reflectance (ATR) platform assembly and a Ge plate, while Raman was carried out with a Thermo Scientific DXR Raman microscope using an argon ion laser excitation (λ = 514.5 nm) at powers of 2–10 mW. The thermal behavior of our composite catalysts with rGO was analyzed by a temperature programmed desorption (TPD) technique on a Micromeritics® AutoChem II 2920 system with a mass spectrometer (SRS QMS200). The UV-vis absorption spectra were obtained using a Thermo scientific Evolution 201 PC spectrophotometer.
2.3. Photocatalytic H2 production
Photocatalytic reactions were carried out in a 250 mL Pyrex flask reactor, which was sealed with a silicone rubber septum. The visible light irradiation between 400 nm and 690 nm was provided by a 300 W Xe lamp (PerkinElmer, PE300BF). The light intensity at the reaction suspension was maintained at approximately 135 mW cm−2 (the exposure area was about 7 cm2), measured by a spectroradiometer with a waterproof probe (Spectral Evolution, SR-1100). In the photocatalytic H2 reactions, the catalyst suspension was made by dispersing 0.066 g of the catalyst (C-TiO2 or C-TiO2/rGO) in 200 mL of DI water with methanol as electron donor at an initial concentration of 25% v/v or 197.95 g L−1 (pH = 5.2). The suspension was mixed continuously using a magnetic stirrer. Before exposure to irradiation, the system was purged with N2 gas for at least 30 min to remove dissolved oxygen. All measurements of produced H2 concentrations at different irradiation times were performed three times to confirm the significance of the presented data and report the mean values with standard deviation as error bars. The H2 concentration in gas phase of the overhead space was determined with gas chromatography (Agilent GC-5890) using HP-MS5 column, TCD, and N2 as the carrier phase. Quantum yield was calculated following the method reported elsewhere.66–68
In the assessment of longevity, the same photocatalytic reactions were conducted for multiple cycles (a cycle is defined as the period, in which photocatalytic H2 production started and terminated due to pseudo equilibrium). After each cycle, we used N2 to purge the suspension at least 30 min to remove the aqueous- and air-phase H2. Then, the illumination was turned on to resume photocatalytic H2 production.
3. Results and discussion
3.1. Morphology and particle size distribution
Fig. 1a–c shows the TEM images of rGO, C-TiO2, and C-TiO2/rGO. GO and rGO have crumpled layered structures with irregular shapes or sizes (also see Fig. S1† for more pictures of these materials). The energy-dispersive X-ray spectroscopy (EDS) result in Fig. S2† indicated that the mass percentage of carbon dopant in C-TiO2 was approximately 2.32%. Fig. 1b shows C-TiO2 nanocrystals aggregated into big clusters. Fig. 1c shows C-TiO2 nanocrystals were anchored onto the carbon sheet of rGO. Individual particles of C-TiO2 had 9.1 ± 3.3 nm in diameter determined from TEM images with the size distribution shown in Fig. 1d, which is consistent with the literature.20 Furthermore, the Scherrer's equation was used to compute the crystallite size of C-TiO2 based on the XRD pattern indexed at (101), (004), (200), and (105). The calculated crystallite sizes ranged from 0.9 nm to 2 nm, indicating C-TiO2 was likely polycrystalline.
 |
| Fig. 1 (a–c) TEM images of GO, C-TiO2 and C-TiO2/rGO. (d) PSD of C-TiO2 determined by ImageJ on TEM images (n = 511). | |
3.2. Hydrodynamic diameters and zeta potentials
The hydrodynamic size distributions and zeta potentials influence colloidal interactions and stability of dispersed photocatalysts. Fig. 2a shows the PSD of the as-synthesized C-TiO2, C-TiO2/rGO, and rGO in DI water dispersion, which all exhibited bimodal size distribution. Hydrodynamic sizes of C-TiO2 or C-TiO2/rGO appear to be greater than those determined by TEM probably because of catalyst aggregation and tendency of DLS to measure large colloids.69 GO had a mean hydrodynamic diameter of about 44 nm, whereas rGO had two peaks at 1480 nm and 4800 nm, which was also reported else.70
 |
| Fig. 2 (a) PSD diagram of GO, rGO, C-TiO2 and C-TiO2/rGO in DI water. (f) Zeta potential of C-TiO2, C-TiO2/rGO, GO and rGO as a function of pH in DI water. | |
Fig. 2b shows zeta potentials of GO/rGO, C-TiO2 and C-TiO2/rGO measured in solutions of different pH (1.5–11.06). Due to the ionization of the multiple surface oxygenated functional groups, GO and rGO exhibited negative charges (e.g., −40 to −50 mV at neutral pH), which is consistent with other studies.71 C-TiO2 and C-TiO2/rGO were positively charged at pH lower than 4 and 5 respectively. As discussed later, carbon dopants likely substituted oxygen atoms in TiO2 (Fig. 3a) and trap excessive electrons,72 which led to a more negative charged surface for C-TiO2. By contrast, P25 TiO2 is usually more positively charged at pH 4–5.73 At higher pHs, they both shifted to negative charges, which is in agreement with other reports.74 The isoelectric point of C-TiO2 and C-TiO2/rGO was 4.3 and 5.2, respectively. The difference was clearly caused by rGO decoration. Moreover, zeta potential of C-TiO2 and C-TiO2/rGO in 25% methanol solution (pH = 5.3) used in our photocatalytic experiments was −0.5 ± 0.6 mV and 14.0 ± 0.6 mV, respectively. A greater positive charge for C-TiO2/rGO was observed although C-TiO2 or rGO both had less positive or even negative charges at the pH range of 4–5. The possible reason for this unexpected positive charge is that after conjugation with C-TiO2, rGO sheets were probably functionalized by some Ti cations as illustrated in Fig. 3b.75
 |
| Fig. 3 Surface functional group variations on (a) TiO2 and C-TiO2 and (b) GO, rGO, and rGO conjugated with C-TiO2. | |
3.3. Crystallinity
XRD patterns in Fig. 4 show that anatase TiO2 was the main polymorph present in the samples of C-TiO2 and C-TiO2/rGO with indexed peaks at (101), (004), (200), (105), (211), (116), (220) and (215) with increasing diffraction angles. These diffraction angles also indicate a body centered tetragonal crystalline structure of TiO2.76 The spectral shift of C-TiO2 compared to commercial P25 TiO2 (Product #: 637254, Aldrich, USA) implies that oxygen atoms in the TiO2 could be substituted by carbon atoms.20,76 No apparent rGO peak in the C-TiO2/rGO sample suggested that C-TiO2 particles may have largely deposited on the surface of rGO and suppressed the XRD signal from the stacking of rGO layers.77
 |
| Fig. 4 XRD pattern for GO, TiO2, C-TiO2 and C-TiO2/rGO. | |
3.4. Surface compositions
Fig. 5 shows the FTIR spectra of GO, rGO, C-TiO2 and C-TiO2/rGO. The unique absorption peaks of GO included 1027 cm−1 for C–O stretching,78 1217 cm−1 for phenolic C–OH stretching,79 1609 cm−1 for the hydroxyl groups of molecular water,80 and 1711 cm−1 for C
O stretching.81 The intensities of absorption bands of oxygen-containing functional groups such as C–O (1052 cm−1) on rGO were dramatically reduced compared with GO. The C–OH at 1217 cm−1 and 1609 cm−1 for the hydroxyl groups of molecular water were still found on C-TiO2/rGO, implying that GO was only partially reduced to rGO by the solvothermal treatment. The strong adsorption at 600 cm−1 found on C-TiO2 and C-TiO2/rGO was due to the Ti–O–Ti bond.80
 |
| Fig. 5 FTIR spectra for C-TiO2, C-TiO2/rGO, GO and rGO. | |
3.5. Raman spectra
Fig. 6 shows the Raman spectra for C-TiO2 and C-TiO2/rGO both displayed active modes at Eg (143 cm−1), Eg (196 cm−1), B1g (398 cm−1), A1g (518 cm−1) and Eg (637 cm−1),82,83 which are related to the tetragonal structure of anatase TiO2 with a D4h space group.84 Thus, Raman spectra also confirmed our synthesized C-TiO2 catalysts were crystallized in anatase phase of TiO2.85 Two characteristic peaks of GO or rGO located at about 1361 and 1590 cm−1 correspond to disorder carbon (D-band) and graphite carbon (G-band) were observed, respectively. C-TiO2/rGO yielded the same two bands, proving the existence of rGO in the composite. These bands correspond to the E2g phonon of sp2-bonded carbon atoms in a two-dimensional hexagonal lattice, as well as the defects and disordered carbon in the graphite layers, respectively. The D/G intensity ratio (0.98) of C-TiO2/rGO was slightly larger than that (0.91) of rGO itself, suggesting a decrease in the average size of the sp2 domains upon reduction of the exfoliated GO. Generally, a Lorentzian peak for the 2D band of the monolayer graphene sheets is observed at 2679 cm−1, whereas this peak may be broaden and shifted to higher wavenumber in case of multi-layer graphene.86 In this study, the 2D band was not observed for GO or rGO, or C-TiO2/rGO, which indicates that the tested GO and rGO may have a monolayer structure and C-TiO2/rGO had no stacking structures.87
 |
| Fig. 6 Raman spectra of GO, rGO, C-TiO2, and C-TiO2/rGO. | |
3.6. Thermal behavior
The TPD system heated up P25 TiO2, C-TiO2, C-TiO2/rGO, GO and rGO from room temperature to 800 °C at a heating rate of 10 °C min−1 under the flowing of helium. Fig. 7 shows that only GO and rGO released CO2 at around 215 °C owing to decomposition of carboxyl functional groups,88 while all other except P25 TiO2 released significant amounts of H2O at 130 °C and 250 °C due to the loss of surface moisture and chemically bounded waters. Particularly, GO released a greater amount (mM CO2 per mg GO or rGO) of CO2 than rGO (Fig. 7a), indicating that rGO had a considerably reduced content of oxygen after hydrothermal treatment.
 |
| Fig. 7 Mass loss of the relevant molecular fragments in TPD analysis. | |
3.7. Band structures
To compare the band gaps of P25 TiO2, C-TiO2, and C-TiO2/rGO, UV-vis diffuse reflectance absorption spectra were obtained for their water dispersion as shown in Fig. 8a. GO and P25 TiO2 exhibited stronger absorption in UV region than in visible light region. Conversely, the absorption bands for C-TiO2 or C-TiO2/rGO shifted to the visible light region. According to the Kubelka–Munk function, the band gap can be determined from the plot (αhν)2 versus the energy of exciting light (hν).66–68,89 Here, α is the absorption coefficient and hν is the photon energy. Based on linear extrapolation (dotted lines) in Fig. 8b, the band gaps of P25 TiO2, C-TiO2 and C-TiO2/rGO was determined to be approximately 3.0 eV, 2.5 eV, 2.2 eV, respectively, which correspond to the wavelengths of 443 nm, 539 nm, and 620 nm. Clearly, carbon doping narrowed the band gap of TiO2, which effectively enables the absorption of visible light. Computational chemistry modeling suggested carbon doping may result in the formation of oxygen vacancies, which could trap excess electrons at the empty p state of the embedded C atoms of TiO2−xCx. C-TiO2 had a slightly narrower band gap than C-TiO2/rGO due to the presence of rGO.90 The presence of rGO clearly interfered the light absorbance of C-TiO2/rGO and therefore changed the apparent band gap of the hybrid structure. The bandgap of GO or rGO can vary from 0 to 4 eV depending on oxygen content or coverage.91 That is why there are no sharp adsorption edges for a precise gap energy, because GO could have different oxidation levels.92 From the approximate linear extrapolation, the synthesized GO had 2.7–3.9 eV for direct transition as indicated by the dotted line on the red curve in Fig. 8b, which are consistent with other studies.91,92
 |
| Fig. 8 (a) UV-vis absorbance spectrums for GO, TiO2, C-TiO2, and C-TiO2/rGO. (b) The band gap values from the plots of (αhν)2 versus hν TiO2 and C-TiO2–rGO (2% loading). (c) The energy level diagram of a Schottky junction with respect to the normal hydrogen electrode (NHE) and the absolute vacuum scale (AVS) as references (pH = 0). The upward band bending in an n-type semiconductor (C-TiO2) at a heterojunction with rGO. | |
For pure anatase TiO2, the conduction band minimum (EC) and the conduction band maximum (EV) is about −0.37 V (vs. NHE) and +2.83 V (vs. NHE), respectively.93 The doping of non-metals in TiO2 lattice results in an upward shift of EV, due to the contribution of 2p or 3p orbital of doped atoms.94 The shift for C-doping originates from mixing the C2p states with the valence band of TiO2. The intimate contact between rGO and C-TiO2 should form a heterojunction or a Schottky junction due to their different work functions (Ef).95,96 As shown in Fig. 8c, C-TiO2 had upward band bending at the interface so that excited electrons flew from the conduction band of C-TiO2 to rGO and from there they spread out and react with water or H+ to form H2.97 In addition, rGO in the nanocomposites was reported to act as an organic dye-like macromolecular ‘photosensitizer’.98 Upon visible light irradiation, rGO may also produce photogenerated electrons that could be transferred to the conduction band of C-TiO2 and thus improve the photocatalytic activity.
3.8. Photocatalytic H2 production and stability
The photocatalytic H2 production was performed only with C-TiO2 and C-TiO2/rGO, whereas P25 TiO2 or P25 TiO2/rGO were not tested because they were unable to produce H2 under visible light irradiation. Fig. 9 shows the photocatalytic H2 production by C-TiO2 and C-TiO2/rGO as a function of irradiation time. The generated H2 were accumulating at stable rates during the initial 6 h. The H2 production rates significantly increased from 0.67 ± 0.12 to 1.50 ± 0.2 mmol g per catalyst per h for C-TiO2 and C-TiO2/rGO respectively, the quantum yield improved from 4% to 10%, indicating that rGO greatly improved the photocatalytic activity. rGO was reported to form Ti–O–C bonds and facilitate both electron migration and efficiency of charge separation.99 The photoexcited electrons in C-TiO2/rGO could readily transfer from the conduction band of C-TiO2 to a graphene acceptor via percolation mechanism,100 therefore promoting the H2 production. This production rate is even higher than other titanium hybrid catalysts with noble metals as co-catalyst such as TiO2/Pt or C-TiO2/Pt (0.02–7 mmol H2 per g per h).101–103 Moreover, in addition to methanol, another common organic electron donor, triethanolamine (TEOA) was also used and shown to produce H2 at a rate of approximately 0.05–0.07 mmol H2 per g per h with our C-TiO2 and C-TiO2/rGO catalysts (Fig. S3†).
 |
| Fig. 9 (a) Photocatalytic H2 production kinetics by C-TiO2 and C-TiO2/rGO. (b and c) H2 production for multiple reaction cycles by C-TiO2 and C-TiO2/rGO. | |
The stability of photocatalytic H2 production over multiple reaction cycles was examined on C-TiO2 and C-TiO2/rGO to provide new insight into the reaction design. Fig. 9b and c shows the H2 production scenarios for multiple reaction cycles. The reaction cycle ended when a pseudo equilibrium reached. More than 90% of spiked methanol (∼188.5 g L−1) still remained in the suspension. Therefore, the catalyst suspension was purged by N2 gas for 30 min to expel aqueous phase H2 or CO2, which was shown effective for rebooting the photocatalytic reaction. The photocatalytic H2 rate for C-TiO2 dramatically declined after one cycle (Fig. 9b). By contrast, C-TiO2/rGO maintained fairly stable rates of H2 production for several cycles (Fig. 9c). For instance, the H2 production rate decreased from 0.058 ± 0.02, to 0.038 ± 0.03 mmol h−1 after seven reaction cycles, highlighting the robust stability of C-TiO2/rGO hybrid composite over C-TiO2. To confirm the changes to colloidal stability after several consecutive reaction cycles, we also measured the zeta potentials and PSD in the reaction suspension, which remained almost the same as Fig. 2. This observation highlights the unparalleled stability and resistance to photocorrosion of C-TiO2 as compared to other sulfide catalysts such as CdS or ZnS.
The decline of photocatalytic H2 production rates could be explained from the reaction mechanisms or stoichiometry as listed in ESI.†23 Methanol as electron donor or sacrificial species are oxidized by the photogenerated holes or radicals such as ˙OH into formaldehyde (HCHO), formic acid (HCOOH) and CO2 ultimately. Photoexcited electrons in the conduction band reduce water or H+ in the solution to form H2. The product accumulation may slow down or even inhibit the photocatalytic reaction. According to the overall reaction
in eqn (S8),† the consumed methanol after 6 hours were determined to be 0.25 mmol and 0.65 mmol for C-TiO2 and C-TiO2/rGO, respectively, accounting for 0.004% and 0.01% of the total methanol initially spiked in the solution. Thus, the most possible reason that the H2 production reached pseudo-equilibrium was the accumulation of H2 in the overhead space of the photoreactor. That is why the photocatalytic reaction could be resumed by vacating H2 from the solution via N2 purge as shown in Fig. 9.
4. Conclusion
Tailored nanostructures offer a new way of facilitating electron–hole separation following excitation. Graphene-based nanomaterials offers additional opportunities to generate unique photocatalysts that demonstrated novel light absorption, thermodynamics, and stability. This work investigated a C-TiO2 decorated onto rGO sheet, which yield enhanced photocatalytic activity and stability. The synthesized C-TiO2 particles were in anatase phase with a band gap of 2.5 eV and thus could utilize visible light. The band gap further decreased to 2.2 eV after anchoring for C-TiO2 to rGO. The H2 production rate was also significantly increased for C-TiO2/rGO compared to C-TiO2. In addition to the effects on band structures and H2 production, rGO decoration increased the longevity of photocatalytic reactions. The results confirm the beneficial roles of graphene sheets in facilitating charge separation and increasing photochemical or colloidal stability of photocatalysts. The new findings advanced the design of sustainable and efficient photocatalytic hybrid materials for renewable energy harvesting and environmental applications.
Acknowledgements
This study was supported by the Research Startup Fund at NJIT and National Science Foundation Grant CBET-1235166. Authors thank Dr Shijian Ge for the help and advice on experimental setup and appreciate the material characterization performed by Mr. Maocong Hu and Dr Xianqin Wang in Department of Chemical Biological and Pharmaceutical Engineering at NJIT.
References
- L. Yan, J. Zhang, X. Zhou, X. Wu, J. Lan, Y. Wang, G. Liu, J. Yu and L. Zhi, Int. J. Hydrogen Energy, 2013, 38, 3554–3561 CrossRef CAS.
- W. Yan, C. L. Zheng, Y. L. Liu and L. J. Guo, Int. J. Hydrogen Energy, 2011, 36, 7405–7409 CrossRef CAS.
- H. Hagiwara, M. Nagatomo, C. Seto, S. Ida and T. Ishihara, J. Photochem. Photobiol., A, 2013, 272, 41–48 CrossRef CAS.
- L. Kuang, Y. Zhao and L. Liu, J. Environ. Monit., 2011, 13, 2496–2501 RSC.
- M. Radecka, M. Rekas, A. Trenczek-Zajac and K. Zakrzewska, J. Power Sources, 2008, 181, 46–55 CrossRef CAS.
- P. D. Tran, S. K. Batabyal, S. S. Pramana, J. Barber, L. H. Wong and S. C. Loo, Nanoscale, 2012, 4, 3875–3878 RSC.
- T. Ikeda, T. Nomoto, K. Eda, Y. Mizutani, H. Kato, A. Kudo and H. Onishi, J. Phys. Chem. C, 2008, 112, 1167–1173 CAS.
- X. Chen and C. Burda, J. Am. Chem. Soc., 2008, 130, 5018–5019 CrossRef CAS PubMed.
- Y. Gai, J. Li, S.-S. Li, J.-B. Xia and S.-H. Wei, Phys. Rev. Lett., 2009, 102, 036402 CrossRef PubMed.
- X. Chen and S. S. Mao, Chem. Rev., 2007, 107, 2891–2959 CrossRef CAS PubMed.
- D. Y. Leung, X. Fu, C. Wang, M. Ni, M. K. Leung, X. Wang and X. Fu, ChemSusChem, 2010, 3, 681–694 CrossRef CAS PubMed.
- X. Fan, X. Chen, S. Zhu, Z. Li, T. Yu, J. Ye and Z. Zou, J. Mol. Catal. A: Chem., 2008, 284, 155–160 CrossRef CAS.
- L. G. Devi, S. G. Kumar, B. N. Murthy and N. Kottam, Catal. Commun., 2009, 10, 794–798 CrossRef CAS.
- T. Ohno, T. Tsubota, Y. Nakamura and K. Sayama, Appl. Catal., A, 2005, 288, 74–79 CrossRef CAS.
- J. L. Gole and J. D. Stout, J. Phys. Chem. B, 2004, 108, 1230–1240 CrossRef CAS.
- A. P. Bhirud, S. D. Sathaye, R. P. Waichal, J. D. Ambekar, C.-J. Park and B. B. Kale, Nanoscale, 2015, 7, 5023–5034 RSC.
- R. Asahi, T. Morikawa, T. Ohwaki, K. Aoki and Y. Taga, Science, 2001, 293, 269–271 CrossRef CAS PubMed.
- X. Chen, P.-A. Glans, X. Qiu, S. Dayal, W. D. Jennings, K. E. Smith, C. Burda and J. Guo, J. Electron Spectrosc. Relat. Phenom., 2008, 162, 67–73 CrossRef CAS.
- C. Xu, Y. A. Shaban, W. B. Ingler and S. U. M. Khan, Sol. Energy Mater. Sol. Cells, 2007, 91, 938–943 CrossRef CAS.
- B. Liu, L.-M. Liu, X.-F. Lang, H.-Y. Wang, X. W. Lou and E. S. Aydil, Energy Environ. Sci., 2014, 7, 2592 CAS.
- E. Barborini, A. M. Conti, I. Kholmanov, P. Piseri, A. Podestà, P. Milani, C. Cepek, O. Sakho, R. Macovez and M. Sancrotti, Adv. Mater., 2005, 17, 1842–1846 CrossRef CAS.
- V. Etacheri, M. K. Seery, S. J. Hinder and S. C. Pillai, Inorg. Chem., 2012, 51, 7164–7173 CrossRef CAS PubMed.
- X. Chen, S. Shen, L. Guo and S. S. Mao, Chem. Rev., 2010, 110, 6503–6570 CrossRef CAS PubMed.
- A. S. Barnard and P. Zapol, Phys. Rev. B: Condens. Matter, 2004, 70, 235403 CrossRef.
- X. Zhang, Y. Sun, X. Cui and Z. Jiang, Int. J. Hydrogen Energy, 2012, 37, 811–815 CrossRef CAS.
- G. D. Moon, J. B. Joo, I. Lee and Y. Yin, Nanoscale, 2014, 6, 12002–12008 RSC.
- W.-J. Ong, M. M. Gui, S.-P. Chai and A. R. Mohamed, RSC Adv., 2013, 3, 4505–4509 RSC.
- Y. Sang, Z. Zhao, J. Tian, P. Hao, H. Jiang, H. Liu and J. P. Claverie, Small, 2014, 10, 3775–3782 CrossRef CAS PubMed.
- J. G. Radich, A. L. Krenselewski, J. Zhu and P. V. Kamat, Chem. Mater., 2014, 26, 4662–4668 CrossRef CAS.
- H.-i. Kim, S. Kim, J.-K. Kang and W. Choi, J. Catal., 2014, 309, 49–57 CrossRef CAS.
- Q. Huang, S. Tian, D. Zeng, X. Wang, W. Song, Y. Li, W. Xiao and C. Xie, ACS Catal., 2013, 3, 1477–1485 CrossRef CAS.
- X. Fu, Y. Liu, X. Cao, J. Jin, Q. Liu and J. Zhang, Appl. Catal., B, 2013, 130–131, 143–151 CrossRef CAS.
- J. G. Radich and P. V. Kamat, ACS Catal., 2012, 2, 807–816 CrossRef CAS.
- F. Chen and N. J. Tao, Acc. Chem. Res., 2009, 42, 429–438 CrossRef CAS PubMed.
- T. Peng, K. Li, P. Zeng, Q. Zhang and X. Zhang, J. Phys. Chem. C, 2012, 116, 22720–22726 CAS.
- D. Wei and Y. Liu, Adv. Mater., 2010, 22, 3225–3241 CrossRef CAS PubMed.
- Y.-L. Min, K. Zhang, Y.-C. Chen and Y.-G. Zhang, Sep. Purif. Technol., 2012, 86, 98–105 CrossRef CAS.
- A. Mathkar, D. Tozier, P. Cox, P. Ong, C. Galande, K. Balakrishnan, A. Leela Mohana Reddy and P. M. Ajayan, J. Phys. Chem. Lett., 2012, 3, 986–991 CrossRef CAS PubMed.
- X.-Y. Zhang, H.-P. Li, X.-L. Cui and Y. Lin, J. Mater. Chem., 2010, 20, 2801–2806 RSC.
- M.-Q. Yang, Y. Zhang, N. Zhang, Z.-R. Tang and Y.-J. Xu, Sci. Rep., 2013, 3, 3314 Search PubMed.
- X.-J. Lv, W.-F. Fu, H.-X. Chang, H. Zhang, J.-S. Cheng, G.-J. Zhang, Y. Song, C.-Y. Hu and J.-H. Li, J. Mater. Chem., 2012, 22, 1539 RSC.
- F. Wang, M. Zheng, C. Zhu, B. Zhang, W. Chen, L. Ma and W. Shen, Nanotechnology, 2015, 26, 345402 CrossRef PubMed.
- P. Cheng, Z. Yang, H. Wang, W. Cheng, M. Chen, W. Shangguan and G. Ding, Int. J. Hydrogen Energy, 2012, 37, 2224–2230 CrossRef CAS.
- W. Fan, Q. Lai, Q. Zhang and Y. Wang, J. Phys. Chem. C, 2011, 115, 10694–10701 CAS.
- H.-i. Kim, G.-h. Moon, D. Monllor-Satoca, Y. Park and W. Choi, J. Phys. Chem. C, 2012, 116, 1535–1543 CAS.
- S. G. Babu, R. Vinoth, D. P. Kumar, M. V. Shankar, H.-L. Chou, K. Vinodgopal and B. Neppolian, Nanoscale, 2015, 7, 7849–7857 RSC.
- Y. Li, C. Xie, S. Peng, G. Lu and S. Li, J. Mol. Catal. A: Chem., 2008, 282, 117–123 CrossRef CAS.
- C. Han, Y. Wang, Y. Lei, B. Wang, N. Wu, Q. Shi and Q. Li, Nano Res., 2015, 8, 1199–1209 CrossRef CAS.
- N. Vinothkumar and M. De, Materials for Renewable and Sustainable Energy, 2014, 3, 1–10 CrossRef.
- F. Pei, S. Xu, W. Zuo, Z. Zhang, Y. Liu and S. Cao, Int. J. Hydrogen Energy, 2014, 39, 6845–6852 CrossRef CAS.
- W. Zhao, Z. Ai, J. Dai and M. Zhang, PLoS One, 2014, 9(8), e103671, DOI:10.1371/journal.pone.0103671.
- H. Li, Y. Hao, H. Lu, L. Liang, Y. Wang, J. Qiu, X. Shi, Y. Wang and J. Yao, Appl. Surf. Sci., 2015, 344, 112–118 CrossRef CAS.
- B. Naik, S. Y. Moon, S. H. Kim and J. Y. Park, Appl. Surf. Sci., 2015, 354(Part B), 347–352 CrossRef CAS.
- K.-I. Liu, C.-Y. Su and T.-P. Perng, RSC Adv., 2015, 5, 88367–88374 RSC.
- B.-S. Huang and M.-Y. Wey, Int. J. Hydrogen Energy, 2011, 36, 9479–9486 CrossRef CAS.
- S.-H. Liu and H.-R. Syu, Int. J. Hydrogen Energy, 2013, 38, 13856–13865 CrossRef CAS.
- J. Gong, C. Yang, J. Zhang and W. Pu, Appl. Catal., B, 2014, 152–153, 73–81 CrossRef CAS.
- S. Sakthivel and H. Kisch, Angew. Chem., Int. Ed., 2003, 42, 4908–4911 CrossRef CAS PubMed.
- J. H. Park, S. Kim and A. J. Bard, Nano Lett., 2006, 6, 24–28 CrossRef CAS PubMed.
- H. Irie, Y. Watanabe and K. Hashimoto, Chem. Lett., 2003, 32, 772–773 CrossRef CAS.
- G. Wu, T. Nishikawa, B. Ohtani and A. Chen, Chem. Mater., 2007, 19, 4530–4537 CrossRef CAS.
- E. M. Rockafellow, X. Fang, B. G. Trewyn, K. Schmidt-Rohr and W. S. Jenks, Chem. Mater., 2009, 21, 1187–1197 CrossRef CAS.
- H. Wang, Z. Wu and Y. Liu, J. Phys. Chem. C, 2009, 113, 13317–13324 CAS.
- N. Pinna, G. Garnweitner, M. Antonietti and M. Niederberger, J. Am. Chem. Soc., 2005, 127, 5608–5612 CrossRef CAS PubMed.
- W. S. Hummers and R. E. Offeman, J. Am. Chem. Soc., 1958, 80, 1339 CrossRef CAS.
- G. Zhang, W. Zhang, P. Wang, D. Minakata, Y. Chen and J. Crittenden, Int. J. Hydrogen Energy, 2013, 38, 1286–1296 CrossRef CAS.
- G. Zhang, W. Zhang, D. Minakata, P. Wang, Y. Chen and J. Crittenden, Int. J. Energy Res., 2014, 38, 1513–1521 CrossRef CAS.
- G. Zhang, W. Zhang, J. Crittenden, D. Minakata, Y. Chen and P. Wang, J. Renewable Sustainable Energy, 2014, 6, 033131 CrossRef.
- W. Zhang, J. Crittenden, K. Li and Y. Chen, Environ. Sci. Technol., 2012, 46, 7054–7062 CrossRef CAS PubMed.
- S. Stankovich, D. A. Dikin, R. D. Piner, K. A. Kohlhaas, A. Kleinhammes, Y. Jia, Y. Wu, S. T. Nguyen and R. S. Ruoff, Carbon, 2007, 45, 1558–1565 CrossRef CAS.
- M. Wang, J. Oh, T. Ghosh, S. Hong, G. Nam, T. Hwang and J.-D. Nam, RSC Adv., 2014, 4, 3284–3292 RSC.
- C. Di Valentin, G. Pacchioni and A. Selloni, Chem. Mater., 2005, 17, 6656–6665 CrossRef CAS.
- Z. Magdolenova, D. Bilaničová, G. Pojana, L. M. Fjellsbø, A. Hudecova, K. Hasplova, A. Marcomini and M. Dusinska, J. Environ. Monit., 2012, 14, 455–464 RSC.
- K. Suttiponparnit, J. Jiang, M. Sahu, S. Suvachittanont, T. Charinpanitkul and P. Biswas, Nanoscale Res. Lett., 2011, 6, 27 Search PubMed.
- I. Chowdhury, N. D. Mansukhani, L. M. Guiney, M. C. Hersam and D. Bouchard, Environ. Sci. Technol., 2015, 49, 10886–10893 CrossRef CAS PubMed.
- R. Taziwa and E. Meyer, Adv. Nanopart., 2014, 03, 54–63 CrossRef CAS.
- X. Fan, W. Peng, Y. Li, X. Li, S. Wang, G. Zhang and F. Zhang, Adv. Mater., 2008, 20, 4490–4493 CrossRef CAS.
- K. Zhou, Y. Zhu, X. Yang, X. Jiang and C. Li, New J. Chem., 2011, 35, 353–359 RSC.
- Z. Wang, B. Huang, Y. Dai, Y. Liu, X. Zhang, X. Qin, J. Wang, Z. Zheng and H. Cheng, CrystEngComm, 2012, 14, 1687–1692 RSC.
- L. Pan, J.-J. Zou, S. Wang, X.-Y. Liu, X. Zhang and L. Wang, ACS Appl. Mater. Interfaces, 2012, 4, 1650–1655 CAS.
- J. Shen, B. Yan, M. Shi, H. Ma, N. Li and M. Ye, J. Mater. Chem., 2011, 21, 3415–3421 RSC.
- G. A. Tompsett, G. A. Bowmaker, R. P. Cooney, J. B. Metson, K. A. Rodgers and J. M. Seakins, J. Raman Spectrosc., 1995, 26, 57–62 CrossRef CAS.
- T. Ohsaka, F. Izumi and Y. Fujiki, J. Raman Spectrosc., 1978, 7, 321–324 CrossRef.
- A. Lamberti, A. Chiodoni, N. Shahzad, S. Bianco, M. Quaglio and C. F. Pirri, Sci. Rep., 2015, 5, 7808 CrossRef CAS PubMed.
- B. Wang, H. Xin, X. Li, J. Cheng, G. Yang and F. Nie, Sci. Rep., 2014, 4, 3729 Search PubMed.
- S. Thakur and N. Karak, Carbon, 2012, 50, 5331–5339 CrossRef CAS.
- G. T. S. How, A. Pandikumar, H. N. Ming and L. H. Ngee, Sci. Rep., 2014, 4, 5044 CAS.
- C. Botas, P. Álvarez, C. Blanco, R. Santamaría, M. Granda, M. D. Gutiérrez, F. Rodríguez-Reinoso and R. Menéndez, Carbon, 2013, 52, 476–485 CrossRef CAS.
- G. Zhang, W. Zhang, D. Minakata, Y. Chen, J. Crittenden and P. Wang, Int. J. Hydrogen Energy, 2013, 38, 11727–11736 CrossRef CAS.
- J. Ding, W. Yan, W. Xie, S. Sun, J. Bao and C. Gao, Nanoscale, 2014, 6, 2299–2306 RSC.
- K.-Y. Lian, Y.-F. Ji, X.-F. Li, M.-X. Jin, D.-J. Ding and Y. Luo, J. Phys. Chem. C, 2013, 117, 6049–6054 CAS.
- T.-F. Yeh, J.-M. Syu, C. Cheng, T.-H. Chang and H. Teng, Adv. Funct. Mater., 2010, 20, 2255–2262 CrossRef CAS.
- H. Yan, X. Wang, M. Yao and X. Yao, Prog. Nat. Sci., 2013, 23, 402–407 CrossRef.
- K. Palanivelu, J. S. Im and Y.-S. Lee, Carbon Letters, 2007, 8, 214–224 CrossRef.
- D. Yang, L. Zhou, L. Chen, B. Zhao, J. Zhang and C. Li, Chem. Commun., 2012, 48, 8078–8080 RSC.
- M. Reza Gholipour, C.-T. Dinh, F. Beland and T.-O. Do, Nanoscale, 2015, 7, 8187–8208 RSC.
- T. Tachikawa, S. Tojo, K. Kawai, M. Endo, M. Fujitsuka, T. Ohno, K. Nishijima, Z. Miyamoto and T. Majima, J. Phys. Chem. B, 2004, 108, 19299–19306 CrossRef CAS.
- Y. Zhang, N. Zhang, Z.-R. Tang and Y.-J. Xu, ACS Nano, 2012, 6, 9777–9789 CrossRef CAS PubMed.
- H. Zhang, X. Lv, Y. Li, Y. Wang and J. Li, ACS Nano, 2010, 4, 380–386 CrossRef CAS PubMed.
- X. Wang, L. Zhi and K. Müllen, Nano Lett., 2008, 8, 323–327 CrossRef CAS PubMed.
- J. Puskelova, L. Baia, A. Vulpoi, M. Baia, M. Antoniadou, V. Dracopoulos, E. Stathatos, K. Gabor, Z. Pap, V. Danciu and P. Lianos, Chem. Eng. J., 2014, 242, 96–101 CrossRef CAS.
- F. Sordello and C. Minero, Appl. Catal., B, 2015, 163, 452–458 CrossRef CAS.
- H. Li, X. Zhang and X. Cui, Int. J. Photoenergy, 2014, 2014, 9 Search PubMed.
Footnote |
† Electronic supplementary information (ESI) available: Table S1, Fig. S1, and synthesis of GO are included in ESI. See DOI: 10.1039/c5ra26096e |
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.