Subscriber access provided by UNIVERSITY OF ADELAIDE LIBRARIES
Article
Two-Dimensional C/TiO2 Heterogeneous Hybrid for Noble-Metal-Free Hydrogen Evolution Song Ling Wang, Jing Li, Shijie Wang, Ji'en Wu, Ten It Wong, Maw Lin Foo, Wei Chen, Kai Wu, and Guo Qin Xu ACS Catal., Just Accepted Manuscript • DOI: 10.1021/acscatal.7b02331 • Publication Date (Web): 28 Aug 2017 Downloaded from http://pubs.acs.org on August 28, 2017
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
ACS Catalysis is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 27
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Catalysis
Two-Dimensional C/TiO2 Heterogeneous Hybrid for Noble-Metal-Free Hydrogen Evolution Song Ling Wang,† Jing Li,† Shijie Wang,ǁ Ji’en Wu,† Ten It Wong,ǁ Maw Lin Foo,† Wei Chen,† Kai Wu,# Guo Qin Xu,†, *
†
Department of Chemistry, National University of Singapore, 3 Science Drive 3,
Singapore 117543, Singapore ǁ
Institute of Materials Research and Engineering, A*STAR, 3 Research Link, 117602,
Singapore #
College of Chemistry and Molecular Engineering, Peking University, Beijing 100871, P.
R. China
1 ACS Paragon Plus Environment
ACS Catalysis
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 2 of 27
ABSTRACT Developing catalysts to improve excitonic charge-carrier transfer and separation properties is critical for solar energy conversion through photochemical catalysis. Layer staking of two-dimensional materials has opened up opportunities to engineer heteromaterials for strong interlayer excitonic transition. However, scalable fabrication of heteromaterials with seamless and clean interfaces remains challenging. Here, we report an in-situ growth strategy for synthesizing a two-dimensional C/TiO2 heterogeneous hybrid. Layered structure of TiO2 and chemically bonded Ti-C between graphitic carbon and TiO2 generate synergetic effects promoting interfacial charge transfer and separation, leading to more electrons participating in photo-reduction for hydrogen evolution. The Ti-C bond as reactive sites, like platinum behavior, makes it an interesting potential substitue for noble metals in hydrogen evolution. In the absence of noble metals, the C/TiO2 hybrid exhibits a significant enhancement of hydrogen evolution from water splitting using solar light, around 3.046 mmol h-1 g-1. The facile and scalable fabrication of two-dimensional heterogeneous hybrid with enhanced interfacial charge transfer and separation provides perspectives for the creation of two-dimensional heteromaterials in optoelectronics and solar light harvesting applications.
KEYWORDS: hydrogen evolution, photocatalysis, water splitting, TiO2, graphitic carbon, heterogeneous hybrid
2 ACS Paragon Plus Environment
Page 3 of 27
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Catalysis
INTRODUCTION Hydrogen evolution from photocatalytic water splitting provides an exciting platform for transformation and storage of solar energy to meet environmental and energy demand.1-3 The ability to control charge-transfer events of catalysts is related to driving the proton reduction to generate hydrogen.4 To achieve high activities of hydrogen evolution, the excited charge carriers (electrons and holes) must migrate from interior to the surface of catalysts with a low recombination possibility.5 In principle, shrinking catalyst size in geometric dimension can promote charge transfer to surfaces by shortening migration routes, hence improving photocatalytic efficiency. Beyond traditional semiconductor catalysts, ultrathin sheets are highly desirable for improving charge transfer due to their fast carrier mobility,6 such as two-dimensional (2-D) transition-metal dichalcogenides (TMDs).7, 8 In general, transition metal oxides materials (TMOs) are photochemically stable. Engineering bulk TMOs into ultrathin layers is a novel strategy to promote excited charge transfer and separation.6,
9, 10
For example, long-distance diffusion of excited
electrons and holes in K4Nb6O17 is greatly reduced due to the ultrathin layered structure.11 Recently interesting advances were made in creating heterostructures of 2-D materials as basic building blocks. Such heterostructures could be formed via van der Waals stacking, where multiple layers are vertically stacked with each other.12, 13 More importantly, the stacked heterostructures can accelerate charge transfer and facilitate electron-hole separation.10, 12, 14, 15 Titania (TiO2) as one of TMOs catalysts had been extensively investigated for its applications in photocatalytic water splitting because of its favorable band-edge
3 ACS Paragon Plus Environment
ACS Catalysis
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
positions, stability, and low cost.4,
16, 17
Page 4 of 27
Thus, developing graphitic carbon-TiO2
heterogeneous hybrid could further improve the efficiency of charge-carrier transfer and separation, hence enhancing H2 evolution from water splitting. Although mechanical transfer techniques had been applied for fabricating heterostructures by stacking different 2-D materials with van der Waals, some limitations should be considered, including poor adhesion, the lack of precise control of stacking orientation, possible contamination of interfaces between layers, and challenging mass fabrication.18-20 Direct and scalable growth of heterogeneous hybrid, particularly has yet to be achieved under mild conditions. Here we report a graphitic carbons-layered TiO2 (GC-LT) heterogeneous hybrid synthesized with an in-situ method. Furthermore, such graphitic carbon can effectively control the formation of two-dimensional TiO2 layers. The photo-induced charge transfer and separation of GC-LT hybrid can be greatly accelerated as a result of the layered structure of TiO2 and interfacial interaction between graphitic carbon and TiO2 layers. Most importantly, despite the absence of noble metals (Pt, Au, Ag, etc.), the GC-LT hybrid catalyst gives rise to a significantly enhanced efficiency of H2 evolution from water splitting and photocurrent response with solar light. EXPERIMENTAL SECTION Characterizations: Field-emission scanning electron microscopic (FE-SEM) images were obtained on JEOL JSM-6701F. Transmission electron microscopy (TEM) and highresolution TEM (HR-TEM) studies were carried out using JEOL 2010 scanning TEM. Xray Diffraction (XRD) patterns were recorded on Bruker D8 High-Resolution XRD featuring a four-bounce Ge(022) incident
beam monochromator and
highly
4 ACS Paragon Plus Environment
Page 5 of 27
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Catalysis
monochromatic X-rays with Cu radiation (Cu Kα= 0.15406 nm). X-ray photoelectron spectroscopy (XPS) measurement was conducted using an ESCALAB Mk2 system (Vacuum Generators) with Mg-Kα. All the binding energy positions were referenced to the C 1s peak of the surface adventitious carbon (at 284.6 eV). Thermogravimetric analysis (TGA) was made on TA 2960 (DTA-TGA). The solid carbon and liquid H nuclear magnetic resonance spectra, 13C-NMR and 1H-NMR, were collected using Bruker DRX400 with CPMAS and Bruker AV500 with autotune 5 mm BBO probes, respectively. Electron spin resonance (ESR) experiments were performed on JEOL FA200 (X-band) spectrometer at room temperature, centered at 331.756 mT with a sweep width of 50 mT. The elemental analysis (EA) for the accurate quantitation of carbon (C), hydrogen (H), nitrogen (N), and Sulphur (S) contents was carried out by the Elementar vario MICRO cube. The metal was determined by the inductively coupled plasma-optical emission spectrometer (ICP) by using the Perkin Elmer Optima 5300 DV. Synthesis of graphitic carbon-layered TiO2 (GC-LT) hybrid: We first prepared the transparent solution by mixing 1 mL of titanium (IV) isopropoxide (99.9%) and oleylamine (5 mL) at room temperature. 5 mL of 1, 2-ethanedithiol (99.5%) was added into the above mixture under stirring. The color of the solution was immediately changed from transparent to red. After stirring for around one hour, this precursor mixture was transferred to a Teflon autoclave with a capacity of 50 mL for solvothermal reaction at 200 oC for 24 hour. Upon completion, the autoclave was cooled down to room temperature and the black precipitate (GC-LT) was collected after washing with ethanol/ hexane, freeze-drying for 2-3 days, and followed by drying in tubular furnace with air flow, at 200 oC, for 24 hours.
5 ACS Paragon Plus Environment
ACS Catalysis
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 6 of 27
Preparation of rGO, rGO-P25, and rGO-TiO2 composites: The rGO was synthesized by hydrothermal reduction of graphene oxide (GO).2 The rGO-P25 and rGO-TiO2 composites were fabricated by integrating 93.2 wt% TiO2 nanoparticles with 5.86 wt% rGO. (1) For synthesis of rGO-P25 nanoparticles composite.21 The 20.0 mg of rGO was suspended in a mixed solution including water (30 ml) and ethanol (15.0 ml). When the solution became homogeneous after ultrasonicating for 5 hours, the commercial TiO2 (P25) powder (321.3 mg) was added and vigorously magnetically stirred for 2 hours. Then, the homogeneous suspension was transferred to the 50 ml of Teflon-sealed autoclave and maintained at 120 oC for 24 h. Finally, the rGO-P25 composite consisting of around 93.2 wt% TiO2 and 5.86 wt% rGO was achieved by this hydrothermal treatment, after washing with water, centrifuging, and drying at 100 oC in oven for overnight. (2) To synthesize rGO-TiO2 nanoparticles composite,22 the TiO2 colloidal suspension was first obtained by the hydrolysis of titanium isopropoxide (0.68 ml) in ethanol (20 ml) under magnetic stirring. After continuous stirring at room temperature for 12 hours, the 22.8 mg of rGO was added into the suspension. The uniform distribution of TiO2 nanoparticles and rGO in the suspension was ensured by 30-min ultrasonication. The resulting rGO-TiO2 powder with around 93.2 wt% TiO2 nanoparticles and 5.86 wt% graphene was collected after washing with water, centrifuging, and drying in oven. Photocurrent responses measurement: A three-electrode system was used to measure the photocurrent. The working electrode is the GC-LT sample coated on ITO (1 × 0.5 cm). Platinum gauze and Ag/AgCl serve as the counter and reference electrodes, respectively. The NaOH aqueous solution (1M) was employed as the electrolyte in a quartz reaction cell. A 300 W xenon lamp provides the simulated sunlight (AM 1.5). The
6 ACS Paragon Plus Environment
Page 7 of 27
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Catalysis
chronoamperometric measurement of GC-LT hybrid was obtained at 0.0 V while the simulator was switched on and off every 10 seconds manually using a shutter to record the change of photocurrent. The P25, rGO-P25, and rGO-TiO2 composites were tested with the same experimental procedure for comparison. Photocatalytic H2 generation: The photocatalytic H2 evolution through water splitting was carried out in a sealed gas circulation system at a pressure of 101 kPa and room temperature (298 K). The quartz reaction cell (150 mL) and pump were connected into this circulation system. The 5 mg of GC-LT hybrid was suspended in 50 mL aqueous methanol solution (40 mL H2O, 10 mL methanol) in quartz reaction cell. Methanol serves as the sacrificial reagent. 300 W Xe lamp was employed as light source, which is the AM 1.5 simulated solar power system. The amount of H2 generated was determined by gas chromatography with the flow of argon as carrier gas. The photocatalytic H2 evolution activities of rGO, P25, rGO-P25, rGO-TiO2 composites were performed with the same procedure for comparison. The main oxidation product of sacrificial methanol reagent is CO2. The hydrocarbons (CH4, C2H4, and C2H6) and CO were also detected, possibly attributable to the CO2 reduction. The apparent quantum efficiency (QE) of hydrogen production was tested by using the same experimental setup for the hydrogen production above. To be more accurate measurement of QE, 50 mg of catalyst powder was used and the irradiation light (350 nm) was obtained by fixing a band-pass filter (350±10 nm) to the 500-W Xe lamp. The light intensity was 6.5 mW cm-2, and the irradiation area was 15.2 cm2. The QE was calculated based on the following equation: =
Number of reacted electrons × 100% Number of incident photons
=
Nubmer of evolved H2 molecules × 2 × 100% !"# $% &'(&)"'* +ℎ$*$'-
7 ACS Paragon Plus Environment
ACS Catalysis
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 8 of 27
RESULTS AND DISCUSSION We developed a lamellar organic micelles strategy for controlling the fabrication of the two-dimensional graphitic carbon-layered TiO2 hybrid (GC-LT). The oleylamine (OA) molecules can electrostatically coordinate to Ti4+ and self-assemble to form lamellar micelles of OA-Ti complex.23-25 This is due to the binding of OA’s weakly cationic amphiphile and lone pairs of nitrogen centers with Ti4+. 1, 2-ethanedithiol (ED) as linkers can also coordinate with Ti4+ to form an organotitanium complex precursor, hence producing isopropanol reacting with residual ED molecule to generate H2O at 200 oC (see Figure S1). For this process, we can see the color change from transparent to red (Figures 1a and b). Meanwhile, the hydrolysis of Ti4+ can occur on the micelles and consequently form two-dimensional TiO2 layers. In this process, the graphitic carbon was synergistically in-situ formed and incorporated to TiO2 layers with black color (Figure 1c). Thus, the two-dimensional GC-LT heterogeneous hybrid does not roll into nanotubes or rods. Further interpretation on the formation of GC-LT hybrid is described with typical characterizations (Figures S1-4) in the Supporting Information. In Figures 2a, S5, and S6, TEM images of the GC-LT hybrid show ultrathin sheet morphology. The XRD patterns of GC-LT hybrid (Figure 2b) present the diffraction peaks of graphitic carbon and TiO2 phases. The TiO2 largely resembles to the anatase phase (JCPDS No. 21-1272). Furthermore, the AFM image and corresponding height profile (Figure 2c) indicate the GC-LT hybrid sheets with a thickness of around ~3.0 nm. To gain more insight into the structure of the hybrid, we carried out HR-TEM studies. The HR-TEM images clearly display crystal structure in Figure 2d. Based on the inserted
8 ACS Paragon Plus Environment
Page 9 of 27
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Catalysis
fast Fourier transform (FFT) patterns, we found that the hexagonal and tetragonal patterns are assigned to the graphitic carbon and TiO2, respectively. Interestingly, the GC-LT heterogeneous hybrid was also found to have layered structures, as indicated in magnified TEM images (Figures 3a, S5, and S6). By examining the edge configuration (e.g., A and B in Figure 3a), we found different spacing or thickness for the sheets. In Figures 3b and c, the spacing distances of ~ 0.7 and ~ 0.76 nm are attributed to the TiO2 monolayer based on literature.7,
26
Due to the existence of
chemical bonds (e.g., Ti-C, Ti-O-C bonds), the spacing of graphitic carbon mononlyer is close to that of graphene oxide (~ 1.0 nm).27, 28 Consequently, the two thicknesses of ~1.0 and ~2.0 nm are assigned to the single- and double- layers of graphitic carbon, respectively. The presence of graphitic carbon and two-dimensional TiO2 layers in GC-LT heterogeneous hybrid material was also evidenced by Raman shifts, FTIR, GC-MS, solid 13
C-NMR, and XPS measurements. Apart from the signals of TiO2 layers in Raman shift
spectrum (Figure 4a) such as Eg (1), B1g (1), A1g+B1g (1), and Eg (2), two Raman shift peaks can be related to the D (~1383 cm-1) and G (~1570 cm-1) bands of graphitic sheets, which are ascribed to the symmetric breathing, A1g of sp3 carbon and in-plane E2g mode vibration of sp2 carbon, respectively.29, 230 FTIR spectrum (Figure 4b) displays that the C=C vibration from aromatic zooms of the graphitic carbon can be directly identified at 1623 cm-1.31-33 More importantly, the graphitic sp2 carbon (at 122.5 ppm) was also found in solid
13
C-NMR spectrum (Figure 4c), further indicating the existence of graphitic
carbon.33, 34 By comparing with reduced graphene oxide (rGO) (120.8 ppm), the slightly left-shifted and broadened resonance peak of GC-LT hybrid suggests the existence of
9 ACS Paragon Plus Environment
ACS Catalysis
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 10 of 27
graphitic carbon flake domains or non-uniformity plausibly due to impure honeycomb lattice incorporating five-, six-, seven-membered carbon rings.35 Therefore, these results together with the two-dimensional layers morphology strongly suggest the in-situ formation of graphitic carbon-layered TiO2 heterogeneous hybrid. The TGA, ICP, and EA characterizations provide the information on the relative amount of graphitic carbon in GC-LT hybrid. As shown in TGA curve (Figure 4d), the weight loss is ~ 6 wt% at 450-700 °C, which is assigned to the removal of carbon skeletons, forming CO or CO2 by oxidation.36 The ICP and EA analysis (Table S1) shows Ti (40.28 wt%), C (5.86 wt%), and H (0.94 wt%) in the GC-LT hybrid. Thus, the GC-LT hybrid contains 93.2 wt% of TiO2 and 5.86 wt% of graphitic carbon. To study the interaction between graphitic carbon and TiO2 layers, XPS spectra was measured over the GC-LT heterogeneous hybrid and reduced graphene oxide (rGO). The rGO was used for comparison and fabricated by reducing graphene oxide based on previous literature.21 The core-level XPS C 1s spectra of both GC-LT hybrid and rGO (Figure 4e and S7) display a main peak with a binding energy of 284.6 eV which is attributable to the non-oxygenated sp2 carbon (C=C) ring in graphitic carbon.32,
37, 38
Furthermore, the peak centered at the binding energy of 285.9 eV is probably ascribed to the C-O oxygenated carbon band.38,
39
Two shoulder peaks positioned at 282 eV and
283.6 eV are possibly related to the C-Ti and Ti-O-C bonds, respectively.40, 41 In addition, the chemical bond (C-Ti) was further examined by the high-resolution of XPS Ti 2p spectrum of GC-LT heterogeneous hybrid. As shown in Figure 4f, the binding energies located at 458.5 (Ti 2p3/2) and 464.6 (Ti 2p1/2) eV are assigned to the lattice Ti-O bond in TiO2 layers.42 Apart from these bands, another set with binding energies at 455.5 (Ti 2p3/2)
10 ACS Paragon Plus Environment
Page 11 of 27
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
ACS Catalysis
and 461.6 (Ti 2p1/2) eV is attributable from the C-Ti bond formed at the interface between graphitic carbon and TiO2.38 Such Ti-C chemical bond was formed during the in-situ synthesis of graphitic carbon-TiO2 heterogeneous hybrid. Whereas the Ti-C bond was not detectable for rGO-P25 composite obtained by mechanical mixing rGO and P25 particles (Figure S8). More importantly, the presence of Ti-C chemical bond can significantly enhance charge transfer and separation in TiO2 layers. Thus, graphitic carbon may act as noble metals (e.g., Pt, Au, etc.) for hydrogen evolution from water splitting. We further studied the optical absorption spectra and valence band position of the GCLT in order to identify whether the band structure was changed because of the presence of graphitic carbon. As shown in Figure 5a, the GC-LT and rGO powders with black color inserted display the similar UV-vis diffuse reflectance (DRS) spectra, significantly different from that of white P25. Furthermore, GC-LT and P25 materials exhibited the same valence band maximum (VBM) positions verified by XPS valance band measurement (Figure 5b), indicating that the layered TiO2 was not lattice-doped by graphitic carbon. As such, C in Ti-C may be attributable to C at the surface linking with other graphitic carbon atoms atoms at the edges or defective sites. This is consistent with the XPS result (Figure 4f) that reduced Ti3+ or Ti2+ species were not observed in GC-LT sample. As a result, graphitic carbon did not lead to narrowing band gap of GC-LT sample. We performed noble-metal-free hydrogen evolution from water splitting over GC-LT hybrid under solar light irradiation. The rGO, rGO-P25, and rGO-TiO2 materials were utilized for comparison. Figure S9 presents XRD patterns of rGO-P25, rGO-TiO2 composites, which are similar to that of GC-LT hybrid. Interestingly, the GC-LT catalyst
11 ACS Paragon Plus Environment
ACS Catalysis
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 12 of 27
gave a much greater photocurrent response than P25, rGO-P25, and rGO-TiO2 materials, as shown in Figure 5c. The H2 evolution efficiency (Figure 5d) for GC-LT hybrid catalyst can achieve a value of around 3046 µmol h-1g-1, which is much higher than those of rGOTiO2 (687.6 µmol h-1g-1), rGO-P25 (355.3 µmol h-1g-1), and P25 (83 µmol h-1 g-1). Pure rGO seems to be inactive for H2 generation, which is consistent with the report.43 We further evaluated the quantum efficiency (QE) of the GC-LT hybrid under light irradiation 350 nm. Figure 5e shows the comparison of QE values over grapene/TiO2based composites: 0.36% (P25) < 1.52% (rGO/P25)21