TiO2 Octahedron for Hydrogen

Feb 14, 2019 - †Key Laboratory of Optoelectronic Materials Chemistry and Physics, Fujian Institute of Research on the Structure of Matter, and §Sta...
0 downloads 0 Views 4MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2019, 4, 3392−3397

http://pubs.acs.org/journal/acsodf

Highly Active Photocatalyst of Cu2O/TiO2 Octahedron for Hydrogen Generation Guojing Li,†,‡ Jiquan Huang,*,† Jian Chen,† Zhonghua Deng,† Qiufeng Huang,† Zhuguang Liu,† Wang Guo,*,† and Rong Cao*,†,§ †

Key Laboratory of Optoelectronic Materials Chemistry and Physics, Fujian Institute of Research on the Structure of Matter, and State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou, Fujian 350002, People’s Republic of China ‡ University of the Chinese Academy of Sciences, Beijing 100039, People’s Republic of China Downloaded via 91.216.3.129 on February 16, 2019 at 10:04:07 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

§

S Supporting Information *

ABSTRACT: Heterojunction catalysts are attracting attention in the field of photocatalytic hydrogen generation for their effective light utilization and charge separation personalities. In this work, we report a simple and low-cost two-step solvothermal method for synthesizing Cu2O/TiO2 heterojunction catalysts with an octahedral morphology and a mean particle size of about 30 nm. It is found that the introduction of Cu2O astonishingly enhances the photocatalytic performance of TiO2. Under the condition of methanol acting as a sacrificial agent, the heterojunction with 0.19% Cu species shows an optimal hydrogen generation rate of 24.83 mmol g−1 h−1, which is nearly 3 orders of magnitude higher than that of the pristine TiO2 catalyst.

1. INTRODUCTION Energy and environment are important to the progress and future of humanity. The continuous increase of world population together with the constant expansion of modern industry speeds up the existing energy consumption. Also, traditional energy utilization has little consideration on environment conservation and energy sustainability, which bring about imperious energy crisis and serious environmental issues.1−3 Development of renewable clean energy is put forward for solving these problems at this crucial moment. Hydrogen, as a renewable and clean energy with unique advantages of high energy density and outstanding combustion efficiency, has been an important renewable clean energy and even been considered as the most hopeful energy in the 21st century.4−7 Up to date, H2 can be produced by steam reforming of natural gas, electrolysis of water, photoelectrochemical or photocatalytic water splitting, and biological/microbial approaches. Among these methods, photocatalytic water splitting into H2 and O2 using solar energy and semiconductor photocatalysts has been considered as the most promising strategy to address the environment conservation and energy sustainability simultaneously.8−11 The overall water-splitting reaction requires a photocatalyst owning a more minus conduction band than H+/H2 reduction potential and a more positive valence band than O2/H2O oxidation potential. Representative photocatalysts are TiO2, ZrO2, KTaO3, SrTiO3, WO3, ZnS, CdS, MoS2, ZnO, SnO2, In2O3, C3N4, SrTiO3, BiVO4, Bi2MoO6, K4Nb6O17, and so on.10,12−14 Unfortunately, most of these materials are not ideal © 2019 American Chemical Society

candidates for water splitting because of their wide band gap, rapid recombination of photogenerated electron/hole pairs, or inferior photocorrosion resistance. For example, TiO2, as one of the most prominent photocatalysts for its high chemical stability and high photocatalytic activity under UV light, is also subject to its poor absorption of visible light and noticeable recombination of photogenerated carriers.3,15−17 Various strategies have been proposed to solve these problems, for instance, by doping, metal loading, and/or constructing heterojunction.18−21 Constructing heterojunctions by introduction of co-catalysts has been proved to be one of the most promising ways to improve the solar-to-chemical energy conversion (and chemical stability) of TiO2 and many other wide band gap (or unstable) semiconductors.22−27 A good cocatalyst for constructing heterojunction should meet the following basic conditions: (i) small band gap (for the harvest of visible light); (ii) electron injection should be fast and efficient; and (iii) nontoxic, lowcost, and chemically stable. Besides, the energy position of the conduction band of TiO2 is very close to the H+/H2 reduction potential.28−31 Therefore, the cocatalyst is generally required to possess a more negative conduction band than that of TiO2 for constructing a TiO2based heterojunction. Cu2O is a semiconductor that meets all these requirements, and therefore the Cu2O/TiO2 heterojunction has attracted much attention.32−37 The Cu species Received: December 4, 2018 Accepted: January 23, 2019 Published: February 14, 2019 3392

DOI: 10.1021/acsomega.8b03404 ACS Omega 2019, 4, 3392−3397

ACS Omega

Article

usually appeared in three forms: metal Cu, oxidation state CuI and CuII, which could exhibit concurrently in the composite and difficult to distinguish. Especially, the quantities were small in the composite. So far, little routes are available for fabricating such Cu2O/TiO2 catalysts that can both load with rare content of Cu and get pure Cu species of Cu2O.38−40 Herein, we successfully prepared the Cu2O/TiO2 catalyst with an octahedral structure and a nanoscaled size. The Cu species existing in the catalyst turned out to be Cu2O, and its content is no more than 1%. Furthermore, the composite catalyst exhibits superior photocatalytic performance when applied to generate hydrogen from a methanol solution. Along with the bonus of a high surface to volume, the Cu2O/TiO2 catalyst provides additional benefits of reuse to reduce the use cost sharply in the industrial-scale production.

with spacings of 4.70, 3.51, 3.48, and 2.40 Å were observed, which correspond to (004), (101), (−101), and (103) planes of the anatase phase, respectively. Besides, as shown in Figure 1a,b, the nanoparticles were highly dispersed, which promised a high specific surface area. According to the nitrogen adsorption−desorption isotherm (Figure S1), the Brunauer− Emmett−Teller (BET) surface area of sample T-3 was as high as 101 m2/g. 2.2. Crystalline Structure. Figure 2 shows the X-ray diffraction (XRD) patterns of the Cu2O/TiO2 photocatalysts.

2. RESULTS AND DISCUSSION 2.1. Surface Morphology. Appropriate solvothermal reaction can convert protonated titanate and cupric acetate into titanium oxide and cuprous oxide nanoparticles, respectively. It is found that the morphology of the products was independent on the Cu/Ti ratio in this study. Figure 1

Figure 2. XRD patterns of TiO2 catalysts loaded with different contents of Cu2O. The standard diffraction pattern of anatase TiO2 (JCPDS card no. 21-1272) is provided at the bottom.

For all the samples, only anatase TiO2 (JCPDS card no. 211272) was observed. It is found that the peak position of the anatase phase remained unchanged with the increasing copper content, which excluded the possibility of copper doping in the TiO2 lattice (or the doping concentration was very low). The difficult substitution of Ti by Cu can be attributed to their large mismatching of the valence state and ionic radius (0.605 Å for Ti4+, 0.73 Å for Cu2+, and 0.77 Å for Cu+). Furthermore, diffraction peaks of Cu2O and any other copper species were not observed in all the Cu2O/TiO2 samples, which may be due to the low content and the poor crystallinity of copper species because the inductively coupled plasma optical emission spectrometry (ICP−OES) (Table 1) and scanning TEM− Table 1. Cu2O/TiO2 Catalysts Loaded with Different Contents of Cu Figure 1. (a) SEM image and (b) TEM image, (c) HRTEM image, and (d) corresponding FFT diffraction pattern of the selected area in (c) of catalyst T-3.

Cu/Ti (at. %) sample designed T-0 T-1 T-2 T-3 T-4 T-5

shows the scanning electron microscopy (SEM) and transmission electron microscopy (TEM) images of the representative sample T-3. The product consisted mainly of octahedron nanoparticles with a mean size of about 30 nm. On the basis of the knowledge of crystallography, these octahedron nanoparticles can be referred to anatase TiO2, and the eight exposed facets can be determined to be {101}. According to the Wulff construction, the octahedron structure enclosed by {101} facets is the thermodynamic equilibrium shape of anatase crystals.41 The high-resolution TEM (HRTEM) image of the octahedron nanoparticles is displayed in Figure 1c and the corresponding fast Fourier transform (FFT) diffraction pattern is presented in Figure 1d. At least three sets of lattice fringes

0 0.02 0.05 0.2 0.5 1

detected (ICP−OES)

H2 generation rate (mmol g−1 h−1)

0 0.0396 0.0722 0.1928 0.3752 0.9763

0.03 5.09 9.93 24.83 20.89 10.09

energy dispersed X-ray spectroscopy (STEM−EDS) (Figures S2 and S3) characterization confirm the presence of Cu. After calcining in argon atmosphere at 500 °C for 2 h, characteristic diffraction peaks of Cu2O were detected in T-6 with a copper content of about 4.4%, as shown in Figure S4. The emergence of Cu2O diffraction peaks after calcination in an inert gas 3393

DOI: 10.1021/acsomega.8b03404 ACS Omega 2019, 4, 3392−3397

ACS Omega

Article

demonstrated the existence of copper species, and copper existed in the form of Cu2O. The structural property of the samples was further examined by Raman spectra. Anatase TiO2 is tetragonal and belongs to the space group (D19 4h = I41/amd), according to the factor group analysis, the fifteen optical modes have the irreducible representation Γopt = A1g (R) + 2B1g (R) + 3Eg (R) + B2u (ia) + A2u (IR) + 2Eu (IR).42 As shown in Figure 3, the Raman

Figure 3. Raman spectra for the Cu2O/TiO2 catalysts.

peaks appeared at 144, 196, 396, 517, and 638 cm−1 can be definitely assigned to Eg(1), Eg(2), B1g(1), A1g + B1g(2), and Eg(3) modes of the anatase phase, respectively. Shift of the Raman peak was not detected, which further demonstrates that there was no doped copper in the TiO2 lattice. For further understanding the chemical environment of Ti and Cu, X-ray photoelectron spectrum (XPS) analysis was carried out. Figure 4 shows the Ti 2p and Cu 2p spectra collected for the Cu2O/TiO2 photocatalysts. The corresponding peaks of the Ti 2p spin orbital appeared at 458.6 and 464.3 eV for Ti 2p3/2 and Ti 2p1/2, respectively, suggesting the presence of Ti4+ in all Cu2O/TiO2 photocatalysts. The highly symmetrical Ti 2p3/2 peak implied the absence of Ti3+, while its constant binding energy indicated the unchanged chemical environment. Figure 4b shows the Cu 2p spectra. Binding energies of Cu 2p3/2 and Cu 2p1/2 were found to be located at 932.7 and 952.5 eV, respectively, which can be assigned to Cu+. For Cu species, the shake-up satellite peak located at about 940−945 eV is characteristic for Cu2+ because of its d9 configuration in the ground state, while the d shell of Cu+ is completely filled (d10), and thus there was no satellite peak in the Cu 2p spectra for Cu2O.32,43 Therefore, the absence of the shake-up satellite peak excluded the existence of Cu2+. On the basis of the analysis of the XPS and XRD results, we can conclude that Cu species exist in the form of Cu2O and there was no doped Cu atom in the TiO2 crystal lattice. UV−vis reflectance spectra were measured and converted from reflectance to absorption by the Kubelka−Munk method.44,45 As shown in Figure 5a, additional visible-light absorption was induced and enhanced with increasing Cu2O content, which can be attributed to the absorption of Cu2O whose absorption band edge is 450−515 nm.46,47 The enhanced absorption of visible light may be beneficial to improve the utilization efficiency of sunlight during the photocatalytic water-splitting process and thereby improve the photocatalytic performance. According to the band

Figure 4. XPS spectra for the Cu2O/TiO2 catalysts. (a) Ti 2p and (b) Cu 2p.

structure of the Cu2O/TiO2 heterojunction (Figure 5b), under the irradiation of visible light, photogenerated electrons transfer from the conduction band of Cu2O to that of TiO2 and then further transfer to the surface-active sites for H+/H2 reduction reaction. Obviously, this heterojunction structure can boost visible-light activity and reduce the recombination rate of photogenerated carriers. The photocatalytic activity of the samples was investigated in a methanol solution under Xe lamp irradiation (Figure 6) and visible light irradiation (Figure S5). Figure 6a shows the time-dependent hydrogen generation. For all the samples, the amount of produced H2 increased linearly with irradiation time. Figures 6b and S5b show the dependence of photocatalytic activity on the Cu2O content. It is found that the increase in copper loading (Cu/Ti ratio) results in increasing the hydrogen production up to 0.19% (i.e., sample T-3), beyond which a negative effect was dominated. The optimal photocatalytic hydrogen generation rate under the Xe lamp irradiation over sample T-3 was 24.83 mmol g−1 h−1, which was about 830 times higher than that of pristine TiO2 (0.03 mmol g−1 h−1). Similarly, the photocatalytic hydrogen generation activity under visible light (λ > 420 nm) irradiation over sample T-3 was also optimized by enhancing the hydrogen generation rate 17 times compared to pristine TiO2. A contrast was made to reveal the excellent performance for the Cu2O/TiO2 octahedron compared with other works as shown in Table S1. The improved reactivity for the Cu2Oloaded TiO2 photocatalysts can be attributed to both the enhancement of visible light absorption and the formation of the Cu2O/TiO2 heterostructure. As described above, the 3394

DOI: 10.1021/acsomega.8b03404 ACS Omega 2019, 4, 3392−3397

ACS Omega

Article

Figure 6. Hydrogen yielded from a methanol solution over pristine TiO2 and Cu2O/TiO2 catalysts. Dependence of the amount of evolved hydrogen on the irradiation time (a) and the corresponding hydrogen generation rate (b).

Figure 5. UV−vis absorption spectra (a) and the energy band diagram (b) for the Cu2O/TiO2 catalysts.

catalytic hydrogen generation performance from water splitting in the presence of the methanol sacrificial agent. The nanosized Cu2O/TiO2 heterojunction catalysts have an octahedral morphology. It is found that the presence of Cu2O can enhance the photocatalytic performance of TiO2 by broadening the light absorption from the ultraviolet region to about 515 nm. The catalysts with 1.9% optimum copper content displays the highest hydrogen yield activity. The corresponding hydrogen generation rate from methanol solution is 24.83 mmol g−1 h−1, which is nearly 3 orders of magnitude better than that of the pristine TiO2. This work has demonstrated a simple and facile method to synthesize environment-friendly Cu2O/TiO2 heterojunction catalysts, which offers useful guidance for developing other transition-metal oxide composited with TiO2 photocatalysts using in energy conversion fields.

introduction of Cu2O with a narrow band gap enhances the light absorption of TiO2 (Figure 5a), which may induce visiblelight photocatalytic activity. Furthermore, TiO2 is an n-type semiconductor, while Cu2O is a p-type semiconductor whose conduction band is more negative than that of TiO2. The contact of TiO2 and Cu2O nanoparticles constructs a p−n junction. Driven by the built-in electric field, the photogenerated electrons migrate from Cu2O to TiO2, while holes migrate in the opposite direction, decreasing the bulk recombination probability of photogenerated carriers. The electrons further migrate to the TiO2 crystal surfaces and react with water to generate H2, while the holes are consumed by the sacrificial agent. Obviously, the formation of heterojunction can greatly facilitate the separation of photogenerated carriers and improve the photocatalytic performance. Besides, the catalyst T-3 was evaluated ten cycles of reuse in methanol solution for H2 production under Xe lamp irradiation (Figure S6) and the photocatalytic activity did not decrease obviously. The used catalysts were annealed and characterized by XRD analysis as shown in Figure S7, and no observable change in the structure was detected. These observations indicated the high chemical stability of the photocatalysts.

4. EXPERIMENTAL SECTION 4.1. Catalysts Preparation. A protonated titanate nanotube was used as the precursor for the synthesis of Cu2O/TiO2 heterogeneous photocatalysts. The H2Ti3O7 nanotube was prepared through a hydrothermal reaction procedure. In a typical process, 5 g of Degussa P25 was dispersed into 80 mL of 8 mol/L NaOH solution and magnetically stirred for 1 h. The mixture was transferred into a 100 mL autoclave and then heated at 180 °C for 10 h. After the hydrothermal reaction, the precipitate was collected and washed with diluted HNO3 solution (pH = 4) repeatedly until pH = 4, followed by washing with distilled water to neutral. The as-obtained

3. CONCLUSIONS In summary, we have successfully synthesized nano-sized Cu2O/TiO2 heterojunction catalysts with outstanding photo3395

DOI: 10.1021/acsomega.8b03404 ACS Omega 2019, 4, 3392−3397

ACS Omega

Article

protonated titanate product was further washed with absolute ethanol for several times and subsequently dried overnight at 60 °C in air. A series of Cu2O/TiO2 catalysts were prepared by the alcohol-thermal method, cupric acetate monohydrate was dissolved into absolute ethanol to prepare 0.01 mol/L and 0.1 mol/L Cu(Ac)2 solution, and 5 g of H2Ti3O7 and a certain amount of 0.01 or 0.1 mol/L Cu(Ac)2 solution (0−65 mL) was added into ethanol and then magnetically stirred for 1 h. The mixture was transferred into a 100 mL autoclave for alcohol-thermal reaction under 150 °C for 20 h. The resulting precipitate was washed by water and alcohol repeatedly and dried overnight at 60 °C in air. The obtained Cu2O/TiO2 catalysts with different copper contents were named as T-0, T1, T-2, T-3, T-4, and T-5, as shown in Table 1. 4.2. Catalyst Characterization. The crystal structure of the Cu2O/TiO2 photocatalyst was characterized by powder XRD on Rigaku MiniFlex 600 (40 kV, 30 mA) with a radiation source of Cu Kα (λ = 1.5418 Å). The data were collected in the range of 10°−80° at a scanning rate of 0.5°/min. The morphology of the catalyst was observed by field emission SEM using a Hitachi SU8010 with an applied voltage of 15 kV and TEM. The surface species analysis was characterized by XPS using a Thermo Fisher ESCALAB 250Xi electron spectrometer with a radiation source of monochromic Al Kα radiation and C 1s peak at 284.8 eV as internal standard. The Cu concentration was determined using ICP−OES (Ultima-2 ICP−OES analyzer, JY, France) and STEM−EDS. UV−vis absorption spectra were recorded on a Lambda 950 spectrophotometer (PerkinElmer, USA). Raman spectra were collected in the range of 100−900 cm−1 from a LabRAM HR instrument (Horiba Jobin-Yvon, France) using a 532 nm laser as the light source. The BET surface area of the catalyst was determined by N2 adsorption using an ASAP-2020 surface area analyzer. 4.3. Photocatalytic Activity Testing. In a typical procedure, 0.1 g of the Cu2O/TiO2 photocatalyst was dispersed in 100 mL 10 vol % aqueous methanol solution in a quarts reactor. The reactor was then equipped on an online photocatalytic activity evaluation system (CEL-SPH2N, CEAULight, China). With the existing of a low-temperature thermostat bath keeping the temperature of the reactor at 5 °C during the photocatalytic reaction, the aqueous methanol solution and catalyst would not be pumped as the airstream when the air in the system was removed by a pump and the yielded hydrogen was sampled during the entire photocatalytic process. The photocatalytic reaction began under the irradiation of a 300 W xenon lamp with or without a band pass filter (λ > 420 nm) accompanied with hydrogen yielded from the catalyst surface and detected by an appendant online chromatography.





photocatalytic stability, and comparison of this work with other works (PDF)

AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. Phone: +86-591-63179098. Fax: +86-591-83721039 (J.H.). *E-mail: [email protected]. Phone: +86-591-63179098. Fax: +86-591-83721039 (W.G.). *E-mail: [email protected]. Phone: +86-591-63173698. Fax: +86-591-63173698 (R.C.). ORCID

Jiquan Huang: 0000-0002-3983-3400 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the National Key R & D Program of China (2016YFB0701003), NSFC (21521061, 21331006), and the Strategic Priority Research Program of the Chinese Academy of Sciences (XDB20000000).



REFERENCES

(1) Saraswat, S. K.; Rodene, D. D.; Gupta, R. B. Recent advancements in semiconductor materials for photoelectrochemical water splitting for hydrogen production using visible light. Renewable Sustainable Energy Rev. 2018, 89, 228−248. (2) Maldonado, M. I.; López-Martín, A.; Colón, G.; Peral, J.; Martínez-Costa, J. I.; Malato, S. Solar pilot plant scale hydrogen generation by irradiation of Cu/TiO2 composites in presence of sacrificial electron donors. Appl. Catal., B 2018, 229, 15−23. (3) Xu, H.; Dai, D.; Li, S.; Ge, L.; Gao, Y. In situ synthesis of novel Cu2CO3(OH)2 decorated 2D TiO2 nanosheets with efficient photocatalytic H2 evolution activity. Dalton Trans. 2018, 47, 348− 356. (4) Dubey, P. K.; Kumar, R.; Tiwari, R. S.; Srivastava, O. N.; Pandey, A. C.; Singh, P. Surface modification of aligned TiO2 nanotubes by Cu2O nanoparticles and their enhanced photo electrochemical properties and hydrogen generation application. Int. J. Hydrogen Energy 2018, 43, 6867−6878. (5) Wang, H.; Chen, F.; Li, W.; Tian, T. Gold nanocluster-sensitized TiO2 nanotubes to enhance the photocatalytic hydrogen generation under visible light. J. Power Sources 2015, 287, 150−157. (6) Hu, Q.; Huang, J.; Li, G.; Jiang, Y.; Lan, H.; Guo, W.; Cao, Y. Origin of the improved photocatalytic activity of Cu incorporated TiO2 for hydrogen generation from water. Appl. Surf. Sci. 2016, 382, 170−177. (7) Sreethawong, T.; Suzuki, Y.; Yoshikawa, S. Photocatalytic evolution of hydrogen over mesoporous TiO2 supported NiO photocatalyst prepared by single-step sol−gel process with surfactant template. Int. J. Hydrogen Energy 2005, 30, 1053−1062. (8) Joshi, R. K.; Shukla, S.; Saxena, S.; Lee, G.-H.; Sahajwalla, V.; Alwarappan, S. Hydrogen generation via photo electrochemical water splitting using chemically exfoliated MoS2 layers. AIP Adv. 2016, 6, 015315. (9) Singh, L.; Wahid, Z. A. Methods for enhancing bio-hydrogen production from biological process: A review. J. Ind. Eng. Chem. 2015, 21, 70−80. (10) Zhang, P.; Zhang, J.; Gong, J. Tantalum-based semiconductors for solar water splitting. Chem. Soc. Rev. 2014, 43, 4395−4422. (11) Huang, J.; Li, G.; Zhou, Z.; Jiang, Y.; Hu, Q.; Xue, C.; Guo, W. Efficient photocatalytic hydrogen production over Rh and Nb codoped TiO2 nanorods. Chem. Eng. J. 2018, 337, 282−289. (12) Liu, G.; Yang, H. G.; Pan, J.; Yang, Y. Q.; Lu, G. Q.; Cheng, H.M. Titanium dioxide crystals with tailored facets. Chem. Rev. 2014, 114, 9559−9612.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.8b03404. Preparation of the Cu2O/TiO2 catalyst with a high Cu content, nitrogen adsorption−desorption isotherm, pore size distribution, STEM−EDS elemental mapping, XRD patterns, H2 yield under visible light irradiation, 3396

DOI: 10.1021/acsomega.8b03404 ACS Omega 2019, 4, 3392−3397

ACS Omega

Article

(31) Dahl, M.; Liu, Y.; Yin, Y. Composite titanium dioxide nanomaterials. Chem. Rev. 2014, 114, 9853−9889. (32) Babu, S. G.; Vinoth, R.; Kumar, D. P.; Shankar, M. V.; Chou, H.-L.; Vinodgopal, K.; Neppolian, B. Influence of electron storing, transferring and shuttling assets of reduced graphene oxide at the interfacial copper doped TiO2 p-n heterojunction for increased hydrogen production. Nanoscale 2015, 7, 7849−7857. (33) Liao, Y.; Deng, P.; Wang, X.; Zhang, D.; Li, F.; Yang, Q.; Zhang, H.; Zhong, Z. A Facile method for preparation of Cu2O-TiO2 NTA heterojunction with visible-photocatalytic activity. Nanoscale Res. Lett. 2018, 13, 221. (34) Rather, R. A.; Singh, S.; Pal, B. A Cu+1/Cu0 -TiO2 mesoporous nanocomposite exhibits improved H2 production from H2O under direct solar irradiation. J. Catal. 2017, 346, 1−9. (35) Lalitha, K.; Sadanandam, G.; Kumari, V. D.; Subrahmanyam, M.; Sreedhar, B.; Hebalkar, N. Y. Highly stabilized and finely dispersed Cu2O/TiO2: A promising visible sensitive photocatalyst for continuous production of hydrogen from glycerol:water mixtures. J. Phys. Chem. C 2010, 114, 22181−22189. (36) Aguirre, M. E.; Zhou, R.; Eugene, A. J.; Guzman, M. I.; Grela, M. A. Cu2O/TiO2 heterostructures for CO2 reduction through a direct Z-scheme: Protecting Cu2O from photocorrosion. Appl. Catal., B 2017, 217, 485−493. (37) Qiu, X.; Miyauchi, M.; Sunada, K.; Minoshima, M.; Liu, M.; Lu, Y.; Li, D.; Shimodaira, Y.; Hosogi, Y.; Kuroda, Y.; Hashimoto, K. Hybrid CuxO/TiO2 nanocomposites as risk-reduction materials in indoor environments. ACS Nano 2012, 6, 1609−1618. (38) Bharad, P. A.; Nikam, A. V.; Thomas, F.; Gopinath, C. S. CuOx-TiO2 composites: Electronically integrated nanocomposites for solar hydrogen generation. ChemistrySelect 2018, 3, 12022−12030. (39) Park, S.-M.; Razzaq, A.; Park, Y. H.; Sorcar, S.; Park, Y.; Grimes, C. A.; In, S.-I. Hybrid CuxO−TiO2 heterostructured composites for photocatalytic CO2 reduction into methane using solar irradiation: Sunlight into fuel. ACS Omega 2016, 1, 868−875. (40) Devaraji, P.; Jo, W.-K. Two-dimensional mixed phase leafTi1‑xCuxO2 sheets synthesized based on a natural leaf template for increased photocatalytic H2 evolution. ChemCatChem 2018, 10, 3813−3823. (41) Wei, Z.; Janczarek, M.; Endo, M.; Wang, K.; Balčytis, A.; Nitta, A.; Méndez-Medrano, M. G.; Colbeau-Justin, C.; Juodkazis, S.; Ohtani, B.; Kowalska, E. Noble metal-modified faceted anatase titania photocatalysts: Octahedron versus decahedron. Appl. Catal., B 2018, 237, 574−587. (42) Ohsaka, T.; Izumi, F.; Fujiki, Y. Raman spectrum of anatase, TiO2. J. Raman Spectrosc. 1978, 7, 321−324. (43) Gopiraman, M.; Ganesh Babu, S.; Khatri, Z.; Kai, W.; Kim, Y. A.; Endo, M.; Karvembu, R.; Kim, I. S. An efficient, reusable copperoxide/carbon-nanotube catalyst for N-arylation of imidazole. Carbon 2013, 62, 135−148. (44) Sasaki, Y.; Nemoto, H.; Saito, K.; Kudo, A. Solar water splitting using powdered photocatalysts driven by Z-schematic interparticle electron transfer without an electron mediator. J. Phys. Chem. C 2009, 113, 17536−17542. (45) Kawasaki, S.; Nakatsuji, K.; Yoshinobu, J.; Komori, F.; Takahashi, R.; Lippmaa, M.; Mase, K.; Kudo, A. Epitaxial Rh-doped SrTiO3 thin film photocathode for water splitting under visible light irradiation. Appl. Phys. Lett. 2012, 101, 033910. (46) Rej, S.; Wang, H.-J.; Huang, M.-X.; Hsu, S.-C.; Tan, C.-S.; Lin, F.-C.; Huang, J.-S.; Huang, M. H. Facet-dependent optical properties of Pd-Cu2O core-shell nanocubes and octahedra. Nanoscale 2015, 7, 11135−11141. (47) Kuo, C.-H.; Chen, C.-H.; Huang, M. H. Seed-mediated synthesis of monodispersed Cu2O nanocubes with five different size ranges from 40 to 420 nm. Adv. Funct. Mater. 2007, 17, 3773−3780.

(13) Kapilashrami, M.; Zhang, Y.; Liu, Y.-S.; Hagfeldt, A.; Guo, J. Probing the optical property and electronic structure of TiO2 nanomaterials for renewable energy applications. Chem. Rev. 2014, 114, 9662−9707. (14) Wang, G.; Chen, H.; Li, Y.; Kuang, A.; Yuan, H.; Wu, G. A hybrid density functional study on the visible light photocatalytic activity of (Mo,Cr)-N codoped KNbO3. Phys. Chem. Chem. Phys. 2015, 17, 28743−28753. (15) Sharma, D.; Mehta, B. R. Nanostructured TiO2 thin films sensitized by CeO2 as an inexpensive photoanode for enhanced photoactivity of water oxidation. J. Alloys Compd. 2018, 749, 329− 335. (16) Sakurai, H.; Kiuchi, M.; Heck, C.; Jin, T. Hydrogen evolution from glycerol aqueous solution under aerobic conditions over Pt/ TiO2 and Au/TiO2 granular photocatalysts. Chem. Commun. 2016, 52, 13612−13615. (17) Nguyen, C.-C.; Nguyen, D. T.; Do, T.-O. A novel route to synthesize C/Pt/TiO2 phase tunable anatase−Rutile TiO2 for efficient sunlight-driven photocatalytic applications. Appl. Catal., B 2018, 226, 46−52. (18) Low, J.; Yu, J.; Jaroniec, M.; Wageh, S.; Al-Ghamdi, A. A. Heterojunction photocatalysts. Adv Mater 2017, 29, 1601694. (19) Wu, M.-C.; Wu, P.-Y.; Lin, T.-H.; Lin, T.-F. Photocatalytic performance of Cu-doped TiO2 nanofibers treated by the hydrothermal synthesis and air-thermal treatment. Appl. Surf. Sci. 2018, 430, 390−398. (20) Montoya, A. T.; Gillan, E. G. Enhanced photocatalytic hydrogen evolution from transition-metal surface-modified TiO2. ACS Omega 2018, 3, 2947−2955. (21) Priebe, J. B.; Radnik, J.; Kreyenschulte, C.; Lennox, A. J. J.; Junge, H.; Beller, M.; Brückner, A. H2 generation with (mixed) plasmonic Cu/Au-TiO2 photocatalysts: Structure-reactivity relationships assessed by in situ spectroscopy. ChemCatChem 2017, 9, 1025− 1031. (22) Wang, Y.; Zhang, Z.; Zhang, L.; Luo, Z.; Shen, J.; Lin, H.; Long, J.; Wu, J. C. S.; Fu, X.; Wang, X.; Li, C. Visible-light driven overall conversion of CO2 and H2O to CH4 and O2 on 3D-SiC@2D-MoS2 heterostructure. J. Am. Chem. Soc. 2018, 140, 14595−14598. (23) Wang, S.; Guan, B. Y.; Wang, X.; Lou, X. W. D. Formation of hierarchical Co9S8@ZnIn2S4 heterostructured cages as an efficient photocatalyst for hydrogen evolution. J. Am. Chem. Soc. 2018, 140, 15145−15148. (24) Wang, S.; Guan, B. Y.; Lou, X. W. D. Construction of ZnIn2S4In2O3 hierarchical tubular heterostructures for efficient CO2 photoreduction. J. Am. Chem. Soc. 2018, 140, 5037−5040. (25) Sun, B.; Zhou, W.; Li, H.; Ren, L.; Qiao, P.; Li, W.; Fu, H. Synthesis of particulate hierarchical tandem heterojunctions toward optimized photocatalytic hydrogen production. Adv. Mater. 2018, 30, 1804282. (26) Wang, S.; Guan, B. Y.; Lu, Y.; Lou, X. W. D. Formation of hierarchical In2S3-CdIn2S4 heterostructured nanotubes for efficient and stable visible light CO2 reduction. J. Am. Chem. Soc. 2017, 139, 17305−17308. (27) Chao, Y.; Zhou, P.; Li, N.; Lai, J.; Yang, Y.; Zhang, Y.; Tang, Y.; Yang, W.; Du, Y.; Su, D.; Tan, Y.; Guo, S. Ultrathin visible-lightdriven Mo incorporating In2O3-ZnIn2Se4 Z-scheme nanosheet photocatalysts. Adv. Mater. 2018, 31, 1807226. (28) Zhang, C.; Zhou, Y.; Bao, J.; Sheng, X.; Fang, J.; Zhao, S.; Zhang, Y.; Chen, W. Hierarchical honeycomb Br-, N-codoped TiO2 with enhanced visible-light photocatalytic H2 production. ACS Appl. Mater. Interfaces 2018, 10, 18796−18804. (29) Matsuoka, M.; Ide, Y.; Ogawa, M. Temperature-dependent photocatalytic hydrogen evolution activity from water on a dyesensitized layered titanate. Phys. Chem. Chem. Phys. 2014, 16, 3520− 3522. (30) Ni, M.; Leung, M. K. H.; Leung, D. Y. C.; Sumathy, K. A review and recent developments in photocatalytic water-splitting using TiO2 for hydrogen production. Renewable Sustainable Energy Rev. 2007, 11, 401−425. 3397

DOI: 10.1021/acsomega.8b03404 ACS Omega 2019, 4, 3392−3397