View Article Online View Journal

ChemComm Accepted Manuscript

This article can be cited before page numbers have been issued, to do this please use: S. Cao, Y. Chen, C. wang, X. lv and W. F. Fu, Chem. Commun., 2015, DOI: 10.1039/C5CC01799H.

This is an Accepted Manuscript, which has been through the Royal Society of Chemistry peer review process and has been accepted for publication. Accepted Manuscripts are published online shortly after acceptance, before technical editing, formatting and proof reading. Using this free service, authors can make their results available to the community, in citable form, before we publish the edited article. We will replace this Accepted Manuscript with the edited and formatted Advance Article as soon as it is available. You can find more information about Accepted Manuscripts in the Information for Authors. Please note that technical editing may introduce minor changes to the text and/or graphics, which may alter content. The journal’s standard Terms & Conditions and the Ethical guidelines still apply. In no event shall the Royal Society of Chemistry be held responsible for any errors or omissions in this Accepted Manuscript or any consequences arising from the use of any information it contains.

www.rsc.org/chemcomm

Page 1 of 4

ChemComm

Journal Name

View Article Online

DOI: 10.1039/C5CC01799H

RSCPublishing

Cite this: DOI: 10.1039/x0xx00000x

Received 00th January 2012, Accepted 00th January 2012

Spectacular photocatalytic hydrogen evolution using metal-phosphide/CdS hybrid catalysts under sunlight irradiation Shuang Cao,a Yong Chen,*a Chuan-Jun Wang,a Xiao-Jun Lv,a and Wen-Fu Fu*a,b

DOI: 10.1039/x0xx00000x www.rsc.org/

A highly efficient and robust heterogeneous photocatalytic hydrogen evolution system was established for the first time by using CoP/CdS hybrid catalyst in water under solar irradiation. The H2-production rate can reach up to 254,000 μmol·h -1·g -1 during 4.5 h of sunlight irradiation, which is one of the highest values ever reported on CdS photocatalytic systems in literature. Photocatalytic hydrogen production has attracted increasing attention due to the depleting of natural fossil fuels and the emerging of energy crisis.[1-3] In the past few decades, various semiconductors have been developed and widely used as photocatalysts for hydrogen evolution.[4] In recent years, many researchers have focused their interest on the development of visible-light-responsive photocatalysts with the aim of efficiently harnessing the abundant visible light (around 420-780 nm) which accounts for about 43% of the solar spectrum.[5,6] CdS has long been recognized as one of the most attractive visible-light-active photocatalysts due to its suitable band edge position.[7,8] However, the photocatalytic efficiency of CdS alone is very low and the practice of introducing a cocatalyst can dramatically enhance the activity of photocatalysts.[9,10] In general, platinum group based metals are considered to be the best cocatalysts for photocatalytic hydrogen evolution, but they suffer from high cost which considerably limits their practical value. [11,12] Therefore, it is still urgently required to find novel highly efficient and noble-metal-free cocatalysts. To this end, state-of-the-art research has been directed to the development of efficient cocatalysts based on earth-abundant metals (Mo, Ni, Co and so on). Li and co-workers have discovered that MoS2 can perform as an efficient cocatalyst when combined with CdS. The system displayed a high hydrogen evolution rate of 5,400 μmol·h -1·g-1 at optimized experimental conditions.[13] NiS has also been reported as a robust cocatalyst for photocatalytic water splitting with an enhanced hydrogen evolution rate of 7,267 μmol·h-1·g-1 in the corresponding NiS/CdS system.[14] Nevertheless, despite all these inspiring results, the development of earth-abundant element based artificial photocatalytic systems with much improved efficiency and outstanding stability to replace platinum remains a long-standing scientific challenge. Moreover, a prospectively real-world applicable photocatalytic system should be not only operative in pure water but

This journal is © The Royal Society of Chemistry 2012

also capable of evolving hydrogen in the presence of atmospheric oxygen, which has been surprisingly much less explored. In the context, the discovery of robust cocatalysts which can more effectively increase the catalytic rate of photocatalysts like CdS may hold the key to solve the problems. Transition-metal phosphides (TMPs) which possess similar characteristics of zero-valent metals have been intensively studied as catalysts for the hydrodesulfurization (HDS) reaction and have exhibited high activity in electrocatalytic water reduction.[15-19] However, their use in photocatalytic hydrogen evolution systems remains almost unexplored. Recently, our group discovered that M2P (M = Ni, Co) nanoparticles could be used as cocatalysts together with CdS nanorods as photo-absorber for photocatalytic hydrogen evolution, and the H2-production rate of 34,800 μmol·h-1·g-1 and 19,373 μmol·h-1·g-1 was achieved, respectively.[20,21] In the previous system, the surfactant poly(vinylpyrrolidone) (PVP) was introduced as stabilizing agent to control the size and morphology of Ni2P and Co2P nanoparticles as well as to increase their water solubility. However, the presence of a protective shell suppresses to some extent electron transfer from photocatalyst to cocatalyst, which is unfavourable for hydrogen production taking place on active sites at catalyst surfaces. We envision that the hydrogen production performance can be dramatically improved by construction of a heterogeneous system in which the surface-clean CdS and TMPs nanoparticles can be in close contact. Herein, we report on a highly robust and inexpensive photocatalytic system with TMPs nanoparticles (CoP, Ni2P and Cu3P) as cocatalysts, CdS as photocatalyst and L(+)-lactic acid as electron donor (Scheme 1). The assembled surface-free TMPs/CdS systems all exhibited outstanding hydrogen evolution performance. Most extraordinarily, the H2 production rate of 251,500 μmol·g-1·h-1 (quantum efficiency of 25.1 %, irradiation λ > 420 nm) for the CoP/CdS system in water represents one of the highest values found for CdS-based photocatalysts up to now. Moreover, the CoP/CdS system displayed continuous and substantial bubbling of H2 gas in the presence of atmospheric oxygen at a rate of 202,800 μmol·h-1·g-1. The system exhibited excellent stability which could maintain its high photocatalytic activity even after 100 h of light irradiation. Very recently, we noted that Sun et al. published an advance article on the use of CoP nanowires for photochemical hydrogen evolution

J. Name., 2012, 00, 1-3 | 1

ChemComm Accepted Manuscript

Published on 16 April 2015. Downloaded by West Virginia University Libraries on 17/04/2015 14:42:41.

COMMUNICATION

ChemComm

Published on 16 April 2015. Downloaded by West Virginia University Libraries on 17/04/2015 14:42:41.

with a fluorescein-based dye as photosensitizer.[22] However, the system showed low photocatalytic activity and poor stability. The hydrogen production rate considerably slowed down after 30 min of visible light irradiation.

Page 2 of 4

Journal Name microscope (SEM) image (Fig. 1b) reveals that the CoP nanoparticles are polydispersed and the corresponding energy dispersive X-ray spectroscopy (EDX) analysis confirmed that the atomic ratio of Co : P is close to 1:1 (the Si, C and O signals come from the Si substrate). As shown in the TEM (Fig. 1c) and high revolution TEM (HRTEM) (Fig. 1d) images, the polydispersed CoP nanoparticles are less than 10 nm in diameter and the lattice fringes with a d-spacing of 0.19 nm is assigned to the strong reflection of (211) plane. The corresponding selective area electron diffraction (SAED) pattern (Fig. 1c, inset) shows five bright rings, which consist of discrete spots and can also be indexed to the (011), (111), (112), (211), and (301) planes of orthorhombic CoP.

Scheme 1 A schematic representation of the TMPs/CdS hybrid photocatalysts for visible-light-driven H2-production.

Fig. 2 (a) TEM, (b) HRTEM, (c) STEM and (d) EDX elemental mapping images of the hybrid CoP/CdS(5) sample.

Fig. 1 (a) XRD pattern and (b) EDX spectrum corresponding to the selected SEM image (inset). (c) TEM image, SADE pattern (inset) and (d) HRTEM image of the obtained CoP nanoparticles.

The hexagonal CdS was prepared according to a previously reported method; Cd(OAc)2 and Na2S were selected as Cd and S precursors, respectively.[23] The product was characterized by X-ray diffraction (XRD) and transmission electron microscope (TEM) (Fig. S1 and S2). The CoP nanoparticles were obtained by a thermal reaction of NaH2PO2 and Co(OH)2 under Ar atmosphere at 300 o [24] C. To prepare Ni2P and Cu3P nanoparticles, Ni(OH)2 and Cu(OH)2 were used instead of Co(OH)2.[25] All the experimental details are provided in the Supporting Information. The TMPs/CdS hybrid photocatalysts were prepared by grounding corresponding TMPs nanoparticles and CdS. The weight ratios of TMPs to CdS were varied from 0.01, 0.03, 0.05, 0.08 to 0.15. The resulting samples were labelled as TMPs/CdS(x), where x = 1, 3, 5, 8, and 15, respectively. Fig. 1a shows the XRD pattern of the obtained orthorhombic CoP nanoparticles (JCPDS 29-0497) and no other impurities such as metallic Co and Co(OH)2 are detected in the sample. Five main diffraction peaks at 2θ = 31.7, 36.5, 46.3, 48.4 and 56.6 o can be clearly observed, which correspond to the (011), (111), (112), (211) and (301) planes of CoP, respectively.[26] The scanning electron

2 | J. Name., 2012, 00, 1-3

Furthermore, TEM, HRTEM, scanning TEM (STEM), UV-vis diffuse reflectance spectra (DRS) and XRD experiments were conducted to characterize the prepared CoP/CdS hybrid material. As depicted in the TEM image (Fig. 2a), CoP nanoparticles (the white circles) are dispersed among the CdS photocatalyst. The HRTEM image further shows that part CdS and CoP nanoparticles are in close contact with each other (Fig. 2b). The lattice spacings of ca. 0.33 nm and 0.19 nm which belong to the diffraction planes of CdS and CoP nanoparticles, respectively, can be clearly observed in the HRTEM image. In addition, STEM image and the corresponding EDX elemental mapping of CoP/CdS(5) hybrid as shown in Fig. 2cd clearly indicate the existence of Cd, S, Co and P elements in the sample and the nonuniform distribution of both Co and P elements. Collectively, it is conclusive from these experimental results that the CoP/CdS hybrid structure was successfully constructed. The interface created between the closely-connected CoP and CdS components would highly favor the vectorial transfer of charge carriers within the hybrids and thus improve the rate of water reduction reaction.[27] The incorporation of CoP into CdS leads to an obvious increase in the visible absorption (Fig. S3). Compared to pure CdS samples, a progressive increase of broad background absorption in the visible light region can be observed with increasing loading of CoP in the CoP/CdS hybrid materials. This is helpful for improving the light absorbance and thus the photocatalytic performance. Further scrutiny

This journal is © The Royal Society of Chemistry 2012

ChemComm Accepted Manuscript

COMMUNICATION

View Article Online

DOI: 10.1039/C5CC01799H

ChemComm

Published on 16 April 2015. Downloaded by West Virginia University Libraries on 17/04/2015 14:42:41.

Journal Name of these spectra shows no significant change in the absorption edge of the CoP/CdS samples, implying that CoP is incorporated onto the surface of CdS rather than into the lattice. In addition, as depicted in Fig. S4, no characteristic XRD diffraction peaks of CoP are observed for each sample from CoP/CdS(1) to CoP/CdS(15), which is probably due to the low loading content and weak crystallization of CoP nanoparticles. Similar phenomenon has also been observed for other photocatalysts.[27,28] The characterized samples CoP/CdS(x) loaded with various amounts of CoP cocatalyst were then subjected to photocatalytic hydrogen evolution experiments with L(+)-lactic acid as electron donor in pure water. The activity of the composite photocatalysts showed strong dependence on the amount of CoP incorporated into the system. Fig. 3a displays a comparison of the photocatalytic H2evolution activity using various hybrid samples from CoP/CdS(1) to CoP/CdS(15) at different irradiation timescales. It is found that with increasing CoP concentration from 1 wt% to 5 wt%, progressive enhancement of hydrogen evolution efficiency can be observed. This is due to the fact that upon increasing the loading amount of CoP cocatalyst, more active sites can be provided for reduction reactions. However, further increasing the amount of CoP from 5 wt% to 15 wt% results in a drop of hydrogen evolution activity, due in part to that excess CoP nanoparticles covered on the surface of CdS may shield the incident light.[29] Therefore, these experimental results reveal an optimized loading value of 5 wt% of CoP cocatalyst for the most efficient hydrogen evolution in the present photocatalytic system. The control experiments show that negligible amounts of H 2 evolution can be detected when any of the components of CoP, CdS, L(+)-lactic acid and light is absent (Fig. S5). The most efficient CoP/CdS(5) sample is therefore selected as the subject for further investigations in subsequent experiments. Surprisingly, an exceedingly high photocatalytic hydrogen evolution rate of up to 251,500 μmol·h-1·g-1 can be achieved for 1 mg CoP/CdS(5) sample during 10 h of visible light irradiation (Movie S1). To the best of our knowledge, this value is one of the highest among the family of CdS-based photocatalysts hitherto reported (Table S1).

Fig. 3 (a) Photocatalytic activities of CoP/CdS hybrid photocatalysts with different CoP contents. (b) Time courses of H2-production under LED (Ar), LED (Air) and sunlight (Ar, Beijing, Dec 6 2014, r.t.) irradiation for CoP/CdS(5). (c) Photocatalytic durability test for CoP/CdS(5). (d) Activity comparisons between different TMPs/CdS and Pt/CdS photocatalysts containing 5 wt% cocatalysts. All the experiments are conducted in 15 mL aqueous solution containing 3 mL L(+)-lactic acid and 1 mg hybrid photocatalyst.

This journal is © The Royal Society of Chemistry 2012

View Article Online

DOI: 10.1039/C5CC01799H

COMMUNICATION To promisingly access the potential in practical applications, we also performed photocatalytic experiments with CoP/CdS(5) sample in air-saturated conditions (Fig. 3b). To our delightful surprise, the photocatalytic system without degassing can also produce hydrogen efficiently with a rate as high as 202,800 μmol·h-1·g-1 during 10 h of LED light irradiation. In comparison with its photocatalytic hydrogen evolution in degassed solution, the system maintained most of its extraordinary hydrogen evolution activity in the presence of atmospheric oxygen. It may be attributed to the facile electron transfer at the interface created between the intimately contacted CoP and CdS materials. In additional, the produced vast bubbling of hydrogen gas during photocatalytic reactions creates a constantly degassing effect, which is therefore beneficial for preventing oxygen from reacting with the photogenerated electrons to form superoxide radicals.[30] Very interestingly, when the photocatalytic system was subjected to sunlight irradiation equipped with a window glass filter, an excellent H2-evolution rate of 254,000 μmol·h-1·g-1 was obtained after 4.5 h of irradiation (10:00 am - 14:30 pm, Fig. S6 and Movie S2). The highly efficient H2-production rates exhibited by the CoP/CdS samples under both sunlight and aerobic condition strongly suggest the potentially practical value of the photocatalytic system. The photocatalytic durability was then tested for the system containing 1 mg CoP/CdS(5) hybrid photocatalyst. As depicted in Fig. 3c, the photocatalytic hydrogen evolution activity only exhibits a slight decrease after 50 h of irradiation, which is due to the photocorrosion of CdS and consumption of L(+)-lactic acid.[31] When 1 mg CdS and 1 mL L(+)-lactic acid were added to the system respectively, the photocatalytic activity was restored rapidly with 16,303 μmol H2 obtained after 100 h of irradiation. Collectively, these results indicate that the CoP cocatalyst is durable under the photocatalytic reaction conditions (Fig. S7). To provide insight into the efficient photocatalytic hydrogen evolution activity of the CoP/CdS system, X-ray photoelectron spectroscopy (XPS) experiments were conducted to examine the charge distribution characterictics of Co and P in the CoP NPs.[26,3233] Fig. S8 shows the XPS spectra in the Co(2p3/2) and P(2p) regions for the prepared CoP nanoparticles with all the peaks corrected in reference to the C signal (285.0 eV) (Fig. S9). The peak located at 779.1 eV originates from the Co(2p3/2) species in the CoP, which is positively shifted in comparison to that of metallic Co (778.1−778.2 eV). Meanwhile, the peak at 129.9 eV which belongs to the P(2p) species in CoP is negatively shifted compared to that of elemental P (130.2 eV). The shifts of the binding energies (BE) of Co(2p 3/2) and P(2p) regions indicate that Co carries a partial positive charge (δ+) while P has a partial negative charge (δ). The binding energy at 130.4 eV is ascribed to oxidized P species due to the superficial oxidation of CoP upon exposing to air. Based on the XPS results and those reported for metal phosphides, the Co-P bond is of covalent nature with charge transfer characteristics. This finding is very similar to that for [NiFe] hydrogenase where pendant base (δ−) attached to metal centers (δ+) serves as active sites to accept protons and electrons for hydrogen production.[25,34] The charge transfer from Co(δ+) metal centre to the pendant base P(δ−) would facilitate the formation of Co-hydride as in molecular metal complexes while at the same time create high electron density spots in the active sites of P(δ−) for accelerated accepting of protons. The ensemble effect would highly facilitate the fast hydrogen release from both the Co and P active sites. Therefore, the CoP nanoparticles combine the merits of both metallic nanoparticles and metal complexes, and simultaneously avoid suffering from the decomposition of molecular catalysts, which enables a highly efficient and robust photocatalytic H2-production. Photocatalytic hydrogen evolution activities of TMPs (Ni2P, Cu3P)/CdS(5) composites (characterization in Fig. S10-S15) were

J. Name., 2012, 00, 1-3 | 3

ChemComm Accepted Manuscript

Page 3 of 4

ChemComm

also measured for comparison and the results, together with that of Pt/CdS(5) are shown in Fig. 3d. For the Ni2P/CdS(5) system, an outstanding H2-production rate of 143,600 μmol·h-1·g-1 is achieved. Besides, Cu3P/CdS(5) hybrid also displays an impressive water splitting activity, with a H2-evolution rate of 77,600 μmol·h-1·g-1. All of the three metal-phosphide/CdS hybrids show a higher activity than the Pt/CdS catalyst under the same experimental conditions (Fig. S16-S18) and the Pt-based photocatalytic systems reported in literature.[23,35-37] Therefore, these prepared metal phosphides based on non-precious metal elements hold great promise to serve as a much cheaper alternative catalyst and may acceptably replace the noble metal Pt nanoparticles for practical photocatalytic applications. In summary, a simple, robust and noble-metal-free heterogeneous photocatalytic system was successfully established. Excellent photocatalytic activity was observed with an especially high hydrogen evolution rate of 251,500 μmol·h-1·g-1 for CoP/CdS(5) sample during 10 h of visible light LED irradiation, and the same system also achieved splendid performance under solar irradiation with a rate of 254,000 μmol·h-1·g-1 for the initial 4.5 h. Under aerobic conditions, a high rate of 202,800 μmol·h-1·g-1 was obtained. In addition, these cocatalysts display outstanding photochemical stability which still show high H2-production rate even after 100 h of irradiation. Furthermore, the simple strategy for constructing the hybrid photocatalyst is expected to provide some insight into developing highly efficient, durable, and earth-abundant element based artificial H2-evolution photocatalysts applicable in water under visible light irradiation, which may hold great promise for practical applications. Efforts to unveil the electron transfer process and photocatalytic mechanism of this TMPs/CdS system by using of ultrafast spectrometry and theoretical calculations are currently in progress.

Acknowledgements This work was supported by the National Key Basic Research Program of China (973 Program 2013CB834804 and 2013CB632403) and by the Ministry of Science and Technology of China (2012DFH40090). We thank the Natural Science Foundation of China (21273257, 21371175), Beijing Natural Science Foundation (2142033), and CAS-Croucher Funding Scheme for Joint Laboratories. Y.C. acknowledges the support of the Chinese Academy of Sciences (100 Talents Program and the Key Research Programme KGZD-EW-T05).

Notes and references a

Key Laboratory of Photochemical Conversion and Optoelectronic Materials and HKU-CAS Joint Laboratory on New Materials, Technical Institute of Physics and Chemistry, Chinese Academy of Sciences, Beijing 100190, P.R. China E-mail: [email protected]; [email protected] b College of Chemistry and Chemical Engineering, Yunnan Normal University, Kunming 650092, P.R. China † Electronic Supplementary Information (ESI) available. See DOI: 10.1039/b000000x/ 1. 2. 3. 4. 5.

A. J. Esswein and D. G. Nocera, Chem. Rev., 2007, 107, 40224047. X. Chen, S. Shen, L. Guo and S. Mao, Chem. Rev., 2010, 110, 65036570. W. T. Eckenhoff and R. Eisenberg, Dalton Trans., 2012, 41, 1300413021. A. Fujishima and M. Honda, Nature, 1972, 238, 3738. Y. Li, L. Fang, R. Jin, Y. Yang, X. Fang, Y. Xing and S. Song,

4 | J. Name., 2012, 00, 1-3

Page 4 of 4

Journal Name 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.

Nanoscale, 2015, 7, 758764. R. Li, H. Han, F. Zhang, D. Wang and C. Li, Energy Environ. Sci., 2014, 7, 13691376. 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1546. S. Cao, C. J. Wang, X. J. Lv, Y. Chen and W. F. Fu, Appl. Catal. B: Environ., 2015, 162, 381391. J. Yang, D. Wang, H. Han and C. Li, Acc. Chem. Res., 2013, 46, 19001909. X. Zong, J. Han, G. Ma, H. Yan, G. Wu and C. Li, J. Phys. Chem. C, 2011, 115, 1220212208. K. Wu, H. Zhu, Z. Liu, W. Rodrí guez-Córdoba and T. Lian, J. Am. Chem. Soc., 2012, 134, 1033710340. J. Zhang, Z. Zhu, Y. Tang, K. Müllen and X. Feng, Adv. Mater., 2013, 26, 734738. X. Zong, H. Yan, G. Wu, G. Ma, F. Wen, L. Wang and C. Li, J. Am. Chem. Soc., 2008, 130, 71767177. W. Zhang, Y. Wang, Z. Zhang, Z. Zhong and R. Xu, Chem. Commun., 2010, 46, 76317633. S. T. Oyama, T. Gott, H. Zhao and Y. K. Lee, Catal. Taday, 2009, 143, 94107. S. J. Sawhill, D. C. Phillips and M. E. Bussell, J. Catal., 2003, 215, 208219. E. J. Popczun, J. R. McKone, C. G. Read, A. J. Biacchi, A. M. Wiltrout, N. S. Lewis and R. E. Schaak, J. Am. Chem. Soc., 2013, 135, 92679270. E. J. Popczun, C. G. Read, C. W. Roske, N. S. Lewis and R. E. Schaak, Angew. Chem. Int. Ed., 2014, 53, 54275430. Z. Xing, Q. Liu, A. M. Asiri and X. Sun, Adv. Mater., 2014, 26, 57025707. S. Cao, Y. Chen, C. J. Wang, P. He and W. F. Fu, Chem. Commun., 2014, 50, 1042710429. S. Cao, Y. Chen, C. C. Hou, X. J. Lv and W. F. Fu, J. Mater. Chem. A, 2015, 3, 60966101. J. Tian, N. Cheng, Q. Liu, W. Xing and X. Sun, Angew. Chem. Int. Ed., DOI: 10.1002/anie.201501237. H. Yan, J. Yang, G. Ma, G. Wu, X. Zong, Z. Lei, J. Shi and C. Li, J. Catal., 2009, 266, 165168. Y. Ni, J. Li, L. Jin, J. Xia, J. Hong and K. Liao, New J. Chem., 2009, 33, 20552059. L. Song, S. Zhang and Q. Wei, Catal. Commun., 2011, 12, 11571160. Q. Liu, J. Tian, W. Cui, P. Jiang, N. Cheng, A. M. Asiri and X. Sun, Angew. Chem. Int. Ed., 2014, 53, 67106714. J. Zhang, L. Qi, J. Ran, J. Yu and S. Z. Qiao, Adv. Energy Mater., DOI: 10.1002/aenm.201301925. J. Yu and J. Ran, Energy Environ. Sci., 2011, 4, 13641371. J. Zhang, S. Z. Qiao, L. Qi and J. Yu, Phys. Chem. Chem. Phys., 2013, 15, 1208812094. M. Lu, Z. Pei, S. Weng, W. Feng, Z. Fang, Z. Zheng, M. Huang and P. Liu, Phys. Chem. Chem. Phys., 2014, 16, 2128021288. Y. Tang, X. Hu and C. Liu, Phys. Chem. Chem. Phys., 2014, 16, 2532125329. Z. Huang, Z. Chen, Z. Chen, C. Lv, M. G. Humphrey and C. Zhang, Nano Energy, 2014, 9, 373382. J. Tian, Q. Liu, A. M. Asiri and X. Sun, J. Am. Chem. Soc., 2014, 136, 75877590. P. Liu and J. A. Rodriguez, J. Am. Chem. Soc., 2005, 127, 1487114878. L. Amirav and A. P. Alivisatos, J. Phys. Chem. Lett., 2010, 1, 10511054. S. R. Lingampalli, U. K. Gautam and C. N. R. Rao, Energy Environ. Sci., 2013, 6, 35893594. Y. Chen, L. Wang, G. Lu, X. Yao and L. Guo, J. Mater. Chem., 2011, 21, 51345141.

This journal is © The Royal Society of Chemistry 2012

ChemComm Accepted Manuscript

Published on 16 April 2015. Downloaded by West Virginia University Libraries on 17/04/2015 14:42:41.

COMMUNICATION

View Article Online

DOI: 10.1039/C5CC01799H

CdS hybrid catalysts under sunlight irradiation.

A highly efficient and robust heterogeneous photocatalytic hydrogen evolution system was established for the first time by using the CoP/CdS hybrid ca...
1MB Sizes 0 Downloads 9 Views