Next Article in Journal
Impact of the Non-Uniform Catalyst Particle Size on Product Selectivities in Consecutive Reactions
Next Article in Special Issue
Single-Atom Catalysts: Preparation and Applications in Environmental Catalysis
Previous Article in Journal
CYP153A71 from Alcanivorax dieselolei: Oxidation beyond Monoterminal Hydroxylation of n-Alkanes
Previous Article in Special Issue
Nitrogen-Doped Pitch-Based Activated Carbon Fibers with Multi-Dimensional Metal Nanoparticle Distribution for the Effective Removal of NO
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Durian-like TiO2@CdS Core-Shell Structure and Study on H2 Generation Properties

1
The School of Material Science & Chemical Engineering, Harbin University of Science & Technology, Harbin 150040, China
2
Heilongjiang Provincial Key Laboratory of CO2 Resource Utilization and Energy Catalytic Materials, Harbin 150040, China
3
Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, China
4
School of Materials Science and Engineering, Nanjing Institute of Technology, Nanjing 211167, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(10), 1211; https://doi.org/10.3390/catal12101211
Submission received: 25 August 2022 / Revised: 29 September 2022 / Accepted: 30 September 2022 / Published: 11 October 2022
(This article belongs to the Special Issue Exclusive Papers in Environmentally Friendly Catalysis in China)

Abstract

:
Novel durian-like TiO2@CdS core-shell particles were synthesized through a solvothermal method in ethylenediamine solution and the obtained nanocomposites were characterized by scanning electron microscopy (SEM), powder X-ray diffraction (XRD), and transmission electron microscopic (TEM) techniques. It can be seen from the characterization that the synthesized core-shell structured particles show uniform size. The possible formation mechanism of TiO2@CdS core-shell particles is also presented schematically. CdS grows on the TiO2 surface in the form of nanorods, turning the TiO2@CdS composite particles into durian-like structures. The durian-like TiO2@CdS core-shell particles prepared in the experiment can overcome the disadvantages of TiO2 and CdS, respectively. They not only produce a higher yield of H2 than pure TiO2; the durian-like TiO2@CdS nanostructures formed at 180 °C for 16 h produced 2.5 times as much H2 as did TiO2, also showing enhanced stability as compared with pure CdS.

1. Introduction

In recent years, photocatalysis via titanium dioxide (TiO2) has been widely investigated as a promising method for environmental and energy applications [1,2,3], and functional properties have been extensively used in various fields including photocatalytic degradation of pollutants [4,5,6], photocatalytic CO2 reduction into energy fuels, water splitting, supercapacitors and lithium-ion batteries [7,8,9,10]. However, due to its broad band gap of 3.2 eV, TiO2 is only sensitive to the light of wavelengths below 380 nm belonging to the UV range, which covers only 5% of the entire solar spectrum [11]. This draw back dramatically limits its conversion efficiency in solar applications. Recent research has focused on improving the photocatalytic efficiency of TiO2 through a variety of material engineering approaches, such as improving charge electroactive surface area [12], utilization of co-catalyst immobilization, controlled faceting, and enhancing the visible-light activity of TiO2 through doping [13,14,15,16,17]. Irfan et al., successfully utilized CoSe as a cocatalyst and ZnSe as a visible-light active semiconductor to significantly enhance the photocatalytic hydrogen production on ZnSe/CdS under visible light irradiation for inexpensive conversion of solar energy to H2 [18]. Pan et al., have successfully developed a hybrid photocatalytic system consisting of CdS NSs as PS and free-standing Fe2P NPs as cocatalysts for photocatalytic hydrogen production from water under visible light [19]. According to previous reports, the charge can be separated efficiently at the heterointerface, thus suitable design of the heterointerface between the other semiconductor catalysts and TiO2 will be an effective strategy for enhancing the performance of photocatalysis [20,21].
Cadmium sulfide (CdS) which is an important II–VI semiconductor (Eĝ = 2.42 eV (515 nm) at room temperature) with many excellent physical and chemical properties [22], has been used as a light-harvesting sensitizer to improve the photoelectric properties of TiO2 because of its appropriate direct energy band gap. Recently, TiO2/CdS composite materials have attracted great interest. They could compensate the disadvantages of the individual components based on some synergistic effects, such as efficient charge separation and migration, expanded visible light response, and improved photostability [23,24,25]. Cao et al., prepared TiO2/CdS core-shell nanorod arrays with visible light activity by a two-step method. Thish increased the probability of electron–hole separation and extended the range of the TiO2 photoresponse from ultraviolet to visible regions [26].
Many fundamental properties of semiconductor materials are dependent on the size and shape of their particles. Controlling the relevant size and shape would provide opportunities for tailoring properties of materials and offer possibilities for observing attractive and useful physical phenomena [27]. Therefore, materials engineering in nanometer-scale has attracted much attention in optics, electronics, magnetics, catalysis, and ceramics [28,29,30,31,32]. In this paper, TiO2@CdS core-shell particles, with CdS nanorods as shell layers and TiO2 nanoparticles as the corresponding core units, were successfully fabricated, in order to overcome the disadvantages of both TiO2 and CdS, thus gaining better optical and electric performances. To the best of our knowledge, this is the first report on TiO2@CdS core-shell particles with a durian-like surface structure.

2. Results and Discussion

2.1. SEM and TEM Characterizations

SEM and TEM techniques were employed to characterize the morphology of the TiO2@CdS particles obtained with different reaction times. As shown in Figure 1a, the commercially available TiO2 nanoparticles with diameters of about 100 nm have a smooth surface. After adding a CdCl2 solution, SC(NH2)2, H2NCH2CH2NH2 and conducting continuous heating for 8 h, we find the surface of TiO2 is no longer smooth due to the formation of a TiO2@CdS composite (Figure 1b). With an extended heating time (16 h), we observe the formation of durian-like TiO2@CdS particles (Figure 1c) with various nanorods on the TiO2 surface, with a diameter of 50 nm. We further used TEM to characterize the TiO2@CdS particles. As shown in Figure 1d, the clear contrast between the core and the shell indeed proves that the CdS nanorods coat the surface of the TiO2 nanoparticles. Furthermore, as shown in Figure 2 and Table 1, the ratio of S:Cd is ca. 1:1. The results of EDX basically indicate that the surface rods are CdS.
Based on the above observations, we can conclude that the reaction time played a crucial role in controlling the durian-like morphology of the TiO2@CdS. With the help of ethylenediamine, the CdS grows into a rod-like structure on the surface of the TiO2 particles forming durian-like TiO2@CdS particles (Figure 1c), as further confirmed by TEM (Figure 1d).

2.2. XRD Characterization

The powder XRD patterns of the TiO2, TiO2@CdS and CdS samples prepared with different reaction times are shown in Figure 3. A typical diffraction peak located at 2θ = 25.3° is attributed to the (101) crystal planes of the anatase-phase TiO2 (JCPDS no. 21-1272). The peaks at 26.5° and 28.2°observed in the XRD of TiO2@CdS and CdS particles are attributed to (002) and (101) crystal planes of hexagonal CdS. If the reaction time lasts longer, the (002) diffraction peak of the TiO2@CdS samples becomes both stronger and narrower, as shown in Figure 3 (TiO2@CdS, 16 h). The unusual (002) diffraction peak indicates that the sample (TiO2@CdS, 16 h) is preferentially orientated along the c axis. The rod shape is confirmed by SEM and TEM photographs. In addition, the typical diffraction peak located at 2θ = 25.3° of anatase-phase TiO2 is weak which can be explained by the thick coating of CdS on the surface of TiO2. This result is also verified by the TEM photograph.

2.3. The Formation Mechanism of TiO2@CdS Core-Shell Particles

The possible formation mechanism of TiO2@CdS core-shell particles is depicted schematically in Figure 4. The first step is to link the Cd2+ to the TiO2 surface through inorganic grafting which is achieved by impregnating TiO2 nanoparticles in a CdCl2 solution. In the second step, the addition of SC(NH2)2 promotes generation of CdS on the surface of the TiO2 nanoparticle, which is used as the seed. In the third step, CdS grows in the form of nanorods on the surface of TiO2 with the help of ethylenediamine, which turns the composite particle into a durian-like structure.

2.4. Photocatalytic Reaction for H2 Generation

In order to verify the photocatalyst activity of the TiO2@CdS particles, the photocatalytic hydrogen evolution ability under light irradiation was studied, and the results are shown in Table 2. We find that the efficiency of H2 generation in the produced hybrid materials increases in the order TiO2 < TiO2@CdS (8 h) < durian-like TiO2@CdS (16 h) with a maximum H2 generation rate of 1065 μmol·h−1·g−1 for the durian-like TiO2@CdS core/shell particles (16 h). This is two times larger than the activity of bare TiO2 particles. The yield of H2 of durian-like TiO2@CdS (16 h) is higher than that of TiO2@CdS (8 h). The results indicate that the durian-like TiO2@CdS particles have better photocatalytic activity than TiO2 particles since the CdS nanorod can expand the visible light response and improve charge separation and migration. The yield of H2 of pure CdS is higher than that of durian-like TiO2@CdS (16 h), but pure CdS is less stable, as suggested by a color change from yellow to black. The color of the hybrid materials TiO2@CdS is not changed during the photocatalytic reaction.
Additionally, the reproducibility experiments have been performed. The H2 generation rate of durian-like TiO2@CdS (16 h) decreased to 800 μmol·h−1·g−1 after four cycle reaction (see the Figure 5). In contrast, the H2 generation rate of pure CdS dropped to 400 μmol·h−1·g−1 in the second cycle reaction. Meanwhile, the SEM experiments of both CdS and durian-like TiO2@CdS after photocatalytic reaction have also been performed (Figure 6). The results indicated that the particle size of durian-like TiO2@CdS slightly changed as compared with the pure CdS. In addition, no obvious change was observed before and after catalysis in the XRD of the TiO2@CdS (Figure 6e). Thus, the reproducibility of H2 generation rate of durian-like TiO2@CdS is basically good. Compared with the state-of-the-art catalysts, the activity and stability of TiO2@CdS is comparable.
From the above results, we can observe that the activity of H2 generation over TiO2@CdS was higher than that of the TiO2, and the stability of TiO2@CdS was better than that of CdS. Thus, we can assume that the charged species can be separated over the composites, and photocorrosion is severely inhibited. Hence, we can assume that a typical Z-scheme mechanism was occurring over the TiO2@CdS. Under simulated sunlight illumination, both TiO2 and CdS can be excited to produce photoinduced electrons and holes. The photoinduced electrons in CdS tend to keep in the CB of CdS, and the photoinduced holes in TiO2 remain in the VB of TiO2. Meanwhile, the photoinduced electrons in the CB of TiO2 combine with the photoinduced holes in VB of CdS (Figure 6f). Thus, the holes over the CdS can be reduced, and the photooxidation reaction may occur on the surface of TiO2 rather than the CdS. Therefore, the carriers can be separated and photocorrosion can be inhibited. In addition, the photo-current responses of CdS, TiO2@CdS and TiO2 were tested to gain a deeper understanding of the efficacy of photoexcited charge separation. As shown in Figure 6g, we can observe that the photocurrent intensity of TiO2@CdS (16) was higher than that of TiO2, indicating more effective charge separation in TiO2@CdS (16). Although the CdS exhibited the highest photocurrent, the H2 generation rate of pure CdS dropped seriously.

3. Materials and Methods

3.1. Subsection

All chemicals were commercially purchased and used without purification. The source of reagents that we used: TiO2 (Tianjin No.3 Chemical Reagent Factory, Tianjin, China), CdCl2 (Tianjin GuangFu Science and Technology Development Co., Ltd., Tianjin, China), thiourea (Tianjin Zhiyuan Chemical Reagent Co., Ltd., Tianjin, China), ethylenediamine (Tianjin Zhiyuan Chemical Reagent Co., Ltd., Tianjin, China) and Ethanol (Tianjin TianLi Chemical Co., Ltd., Tianjin, China).

3.2. Preparation of TiO2@CdS Core-Shell Particles

These particles were synthesized by a solvothermal method [33,34]. The commercial anatase-phase TiO2 nanoparticles (1 g, 0.013 mol) were added to dilute nitric acid (10 mL, 1 mol·L−1) with ultrasound for 4 h, and were collected and washed with distilled water to neutral. The final products were dried at 80 °C for 4 h in a vacuum box. The obtained TiO2 (0.1 g) nanoparticles and 5 mL of an aqueous solution of CdCl2 (1 mol·L−1) were added to a beaker with a capacity of 100 mL and were stirred magnetically for 4 h at room temperature. 5 mL of a thiourea solution (1.5 mol·L−1) and the mixture described above were added to a teflon-lined stainless steel autoclave with a 25 mL capacity. The autoclave was filled with 10 mL ethylenediamine solution up to about 80% of the total volume. The solution was stirred and treated with ultra-sound. The autoclave was maintained at 180 °C for 8–16 h and then allowed to cool to room temperature. Yellow precipitates were collected and washed with ethanol and distilled water to remove residues of organic material. The final products were dried at 80 °C for 4 h in a vacuum box. The pure CdS was prepared by the same method without TiO2.

3.3. Characterization of TiO2@CdS Core-Shell Particles

The morphologies of the TiO2 nanoparticles and durian-like TiO2@CdS core-shell particles were characterized by scanning electron microscopy (SEM, SU8020) and transmission electron microscopy (TEM, JEM-2100). To facilitate the investigation of the detailed surface morphology of TiO2@CdS, no conductive metal layer such as Au, Pd/Au, Cr, or carbon layer was coated on the sample surface. The samples were further analyzed with a Bruker D8 Avance X-ray diffractometer (XRD) using Ni-filtered Cu Kα radiation at 40 kV and 40 mA in the 2θ range of 20–80°, with a scan rate of 0.02° per second.

3.4. Photocatalytic Experiment

Photocatalytic reactions of hydrogen production by water splitting were conducted in a gas-closed system with a side irradiation Pyrex cell. An aluminum alloy shell was employed outside the Pyrex cell to reflect and gather the visible light originating from Xe lamp. A water cycling system was used to maintain the reaction temperature.
Photocatalyst powder (10 mg) was dispersed in a solution (2 mL) containing 0.35 mol·L−1 Na2S and 0.25 mol·L−1Na2SO3. After being evacuated and flushed by N2 gas for over 10 min, the photocatalysts were irradiated by visible light (λ ≥ 300 nm) from a 300 W Xe lamp for 4 h. The amount of H2 gas was determined using a gas chromatograph (Bruker GC-4890, using column type TDX01).

4. Conclusions

In summary, novel durian-like TiO2@CdS core-shell particles were synthesized through a solvothermal method in ethylenediamine solution. As demonstrated, CdS coated the surface of the TiO2 material. The reaction time plays a crucial role in controlling the nucleation and growth of crystallites; with a prolonged reaction time, CdS gradually forms a rod shape, which effectively enhances visible light absorption and possesses a high electron affinity. The durian-shaped TiO2@CdS composite not only produced a higher yield of H2 than pure TiO2, but also possessed an enhanced stability as compared with the pure CdS. Therefore, the durian-like TiO2@CdS core-shell particles might be a good catalyst for application in photocatalysis.

Author Contributions

Conceptualization, Z.C. and D.L.; methodology, Z.C. and X.W.; software, C.C.; investigation, D.L. and Z.Z.; data curation, X.W.; writing—original draft preparation, Z.C.; writing—review and editing, D.L., Z.C., X.W., Z.Z., C.C. and M.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

The authors thank the Harbin University of Science & Technology and Chinese Academy of Sciences for experimental and technical support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ge, M.; Cao, C.; Huang, J.; Li, S.; Chen, Z.; Zhang, K.-Q.; Al-Deyab, S.S.; Lai, Y. A review of one-dimensional TiO2 nanostructured materials for environmental and energy applications. J. Mater. Chem. A 2016, 4, 6772–6801. [Google Scholar] [CrossRef]
  2. Pelaez, M.; Nolan, N.T.; Pillai, S.C.; Seery, M.K.; Falaras, P.; Kontos, A.G.; Dunlop, P.S.M.; Hamilton, J.W.J.; Byrne, J.A.; O’Shea, K. A review on the visible light active titanium dioxide photocatalysts for environmental applications. Appl. Catal. B Environ. 2012, 125, 331–349. [Google Scholar] [CrossRef] [Green Version]
  3. Rempel, A.A.; Valeeva, A.A.; Vokhmintsev, A.S.; Weinstein, I.A. Titanium dioxide nanotubes: Synthesis, structure, properties and applications. Russ. Chem. Rev. 2021, 90, 1397–1414. [Google Scholar] [CrossRef]
  4. Falletta, E.; Bianchi, C.L.; Morazzoni, F.; Polissi, A.; Di Vincenzo, F.; Bellobono, I.R. Tungsten Trioxide and Its TiO2 Mixed Composites for the Photocatalytic Degradation of NOX and Bacteria (Escherichia coli) Inactivation. Catalysts 2022, 12, 822. [Google Scholar] [CrossRef]
  5. Bergamonti, L.; Graiff, C.; Bergonzi, C.; Potenza, M.; Reverberi, C.; Ossiprandi, M.C.; Lottici, P.P.; Bettini, R.; Elviri, L. Photodegradation of Pharmaceutical Pollutants: New Photocatalytic Systems Based on 3D Printed Scaffold-Supported Ag/TiO2 Nanocomposite. Catalysts 2022, 12, 580. [Google Scholar] [CrossRef]
  6. Yu, J.; Caravaca, A.; Guillard, C.; Vernoux, P.; Zhou, L.; Wang, L.; Lei, J.; Zhang, J.; Liu, Y. Carbon Nitride Quantum Dots Modified TiO2 Inverse Opal Photonic Crystal for Solving Indoor VOCs Pollution. Catalysts 2021, 11, 464. [Google Scholar] [CrossRef]
  7. Xiao, T.; Chen, Y.; Liang, Y. Ni(II) Tetra(4-carboxylphenyl)porphyrin-Sensitized TiO2 Nanotube Array Composite for Efficient Photocatalytic Reduction of CO2. J. Phys. Chem. C 2022, 126, 9742–9752. [Google Scholar] [CrossRef]
  8. Liu, J.; Liu, B.; Ren, Y.; Yuan, Y.; Zhao, H.; Yang, H.; Liu, S. Hydrogenated nanotubes/nanowires assembled from TiO2 nanoflakes with exposed {111} facets: Excellent photo-catalytic CO2 reduction activity and charge separation mechanism between (111) and (111) polar surfaces. J. Mater. Chem. A 2019, 7, 14761–14775. [Google Scholar] [CrossRef]
  9. Appadurai, T.; Subramaniyam, C.M.; Kuppusamy, R.; Karazhanov, S.; Subramanian, B. Electrochemical Performance of Nitrogen-Doped TiO2 Nanotubes as Electrode Material for Supercapacitor and Li-Ion Battery. Molecules 2019, 24, 2952. [Google Scholar] [CrossRef] [Green Version]
  10. Li, H.; Wang, S.; Wang, M.; Gao, Y.; Tang, J.; Zhao, S.; Chi, H.; Zhang, P.; Qu, J.; Fan, F.; et al. Enhancement of Plasmon-Induced Photoelectrocatalytic Water Oxidation over Au/TiO2 with Lithium Intercalation. Angew. Chem. Int. Ed. 2022, 61, e202204272. [Google Scholar]
  11. Nam, Y.; Lim, J.H.; Ko, K.C.; Lee, J.Y. Photocatalytic activity of TiO2 nanoparticles: A theoretical aspect. J. Mater. Chem. A 2019, 7, 13833–13859. [Google Scholar] [CrossRef]
  12. Al Qarni, F.; Alomair, N.A.; Mohamed, H.H. Environment-Friendly Nanoporous Titanium Dioxide with Enhanced Photocatalytic Activity. Catalysts 2019, 9, 799. [Google Scholar] [CrossRef] [Green Version]
  13. Yang, B.; Ma, Z.; Wang, Q.; Yang, J. Synthesis and Photoelectrocatalytic Applications of TiO2/ZnO/Diatomite Composites. Catalysts 2022, 12, 268. [Google Scholar] [CrossRef]
  14. Cai, M.; Cao, S.; Zhuo, Z.; Wang, X.; Shi, K.; Cheng, Q.; Xue, Z.; Du, X.; Shen, C.; Liu, X.; et al. Fabrication of Ni2P Cocatalyzed CdS Nanorods with a Well-Defined Heterointerface for Enhanced Photocatalytic H2 Evolution. Catalysts 2022, 12, 417. [Google Scholar] [CrossRef]
  15. Gao, X.-F.; Sun, W.-T.; Hu, Z.-D.; Ai, G.; Zhang, Y.-L.; Feng, S.; Li, F.; Peng, L.-M. An Efficient Method To Form Heterojunction CdS/TiO2 Photoelectrodes Using Highly Ordered TiO2 Nanotube Array Films. J. Phys. Chem. C 2009, 113, 20481–20485. [Google Scholar] [CrossRef]
  16. Kang, X.; Chaperman, L.; Galeckas, A.; Ammar, S.; Mammeri, F.; Norby, T.; Chatzitakis, A. Water Vapor Photoelectrolysis in a Solid-State Photoelectrochemical Cell with TiO2 Nanotubes Loaded with CdS and CdSe Nanoparticles. ACS Appl. Mater. Int. 2021, 13, 46875–46885. [Google Scholar] [CrossRef]
  17. Pal, S.; Taurino, A.; Catalano, M.; Licciulli, A. Block Copolymer and Cellulose Templated Mesoporous TiO2-SiO2 Nanocomposite as Superior Photocatalyst. Catalysts 2022, 12, 770. [Google Scholar] [CrossRef]
  18. Irfan, R.M.; Tahir, M.H.; Maqsood, M.; Lin, Y.; Bashir, T.; Iqbal, S.; Zhao, J.; Gao, L.; Haroon, M. CoSe as non-noble-metal cocatalyst integrated with heterojunction photosensitizer for inexpensive H2 production under visible light. J. Catal. 2020, 390, 196–205. [Google Scholar] [CrossRef]
  19. Pan, Z.; Wang, R.; Li, J.; Iqbal, S.; Liu, W.; Zhou, K. Fe2P nanoparticles as highly efficient freestanding co-catalyst for photocatalytic hydrogen evolution. Int. J. Hydrogen Energy 2018, 43, 5337–5345. [Google Scholar] [CrossRef]
  20. Kim, H.-I.; Kim, J.; Kim, W.; Choi, W. Enhanced Photocatalytic and Photoelectrochemical Activity in the Ternary Hybrid of CdS/TiO2/WO3 through the Cascadal Electron Transfer. J. Phys. Chem. C 2011, 115, 9797–9805. [Google Scholar] [CrossRef]
  21. Lv, Z.; Wang, Y.; Liu, Y.; Wang, J.; Qin, G.; Guo, Z.; Zhang, C. NiB as a Substitute for the Pt Cocatalyst in CdS with Enhanced Visible-Light Photocatalytic H2 Production. J. Phys. Chem. C 2022, 126, 9041–9050. [Google Scholar] [CrossRef]
  22. Jie, J.S.; Zhang, W.J.; Jiang, Y.; Meng, X.M.; Li, Y.Q.; Lee, S.T. Photoconductive Characteristics of Single-Crystal CdS Nanoribbons. Nano Lett. 2006, 6, 1887–1892. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, J.; Cai, P.; Lin, J. Modulation of the Band Bending of CdS by Fluorination to Facilitate Photoinduced Electron Transfer for Efficient H2 Evolution over Pt/CdS. J. Phys. Chem. C 2022, 126, 7896–7902. [Google Scholar] [CrossRef]
  24. Du, Y.-e.; Niu, X.; He, X.; Hou, K.; Liu, H.; Zhang, C. Synthesis and Photocatalytic Activity of TiO2/CdS Nanocomposites with Co-Exposed Anatase Highly Reactive Facets. Molecules 2021, 26, 6031. [Google Scholar] [CrossRef] [PubMed]
  25. Guo, N.; Wang, C.; Bao, W.; Zeng, Y.; Yu, H. Hollow TiO2@CdS nanosphere: Interface construction for spatial charge separation and higher charge transfer efficiency. J. Environ. Chem. Eng. 2021, 9, 106211. [Google Scholar] [CrossRef]
  26. Cao, C.; Hu, C.; Shen, W.; Wang, S.; Tian, Y.; Wang, X. Synthesis and characterization of TiO2/CdS core–shell nanorod arrays and their photoelectrochemical property. J. Alloys Compd. 2012, 523, 139–145. [Google Scholar] [CrossRef]
  27. Kim, M.R.; Kang, Y.-M.; Jang, D.-J. Synthesis and Characterization of Highly Luminescent CdS@ZnS Core−Shell Nanorods. J. Phys. Chem. C 2007, 111, 18507–18511. [Google Scholar] [CrossRef]
  28. Chen, S.; Peng, Y.; Li, C.; Hou, Z. The co-decorated TiO2 nanorod array photoanodes by CdS/CdSe to promote photoelectrochemical water splitting. Int. J. Hydrogen Energy 2021, 46, 32055–32068. [Google Scholar] [CrossRef]
  29. Dang, W.; Xu, K.; Zhang, L.; Qian, Y. Fabrication of multilayer 1D TiO2/CdS/ZnS with high photoelectrochemical performance and enhanced stability. J. Alloys Compd. 2021, 886, 161329. [Google Scholar] [CrossRef]
  30. Revathi, M.; Jeyakumari, A.P. Fabrication of TiO2/CdS heterostructure photoanodes and optimization of light scattering to improve the photovoltaic performance of dye-sensitized solar cells (DSSCs). J. Mater. Sci. Mater. Electron. 2021, 32, 11921–11930. [Google Scholar] [CrossRef]
  31. Devaraji, P.; Gao, R.; Xiong, L.; Jia, X.; Huang, L.; Chen, W.; Liu, S.; Mao, L. Usage of natural leaf as a bio-template to inorganic leaf: Leaf structure black TiO2/CdS heterostructure for efficient photocatalytic hydrogen evolution. Int. J. Hydrogen Energy 2021, 46, 14369–14383. [Google Scholar] [CrossRef]
  32. Lu, Y.; Yi, G.; Zhou, H.; Liu, J.; Zhang, H.; Lu, L.; An, L. Arrays of needle-like TiO2/CdS nanorod heterostructure photoelectrodes with enhanced photoelectrochemical properties fabricate by pulsed laser deposition. Vacuum 2021, 184, 109985. [Google Scholar] [CrossRef]
  33. Yang, J.; Zeng, J.-H.; Yu, S.-H.; Yang, L.; Zhou, G.-e.; Qian, Y.-t. Formation Process of CdS Nanorods via Solvothermal Route. Chem. Mater. 2000, 12, 3259–3263. [Google Scholar] [CrossRef]
  34. Ye, C.; Meng, G.; Wang, Y.; Jiang, Z.; Zhang, L. On the Growth of CdS Nanowires by the Evaporation of CdS Nanopowders. J. Phys. Chem. B 2002, 106, 10338–10341. [Google Scholar] [CrossRef]
Figure 1. SEM and TEM image. (a) SEM image of TiO2; (b) SEM image of TiO2@CdS formed at 180 °C for 8 h; (c) SEM image of durian-like TiO2@CdS nanostructure formed at 180 °C for 16 h; (d) TEM image of a durian-like TiO2@CdS core-shell nanostructure formed at 180 °C for 16 h.
Figure 1. SEM and TEM image. (a) SEM image of TiO2; (b) SEM image of TiO2@CdS formed at 180 °C for 8 h; (c) SEM image of durian-like TiO2@CdS nanostructure formed at 180 °C for 16 h; (d) TEM image of a durian-like TiO2@CdS core-shell nanostructure formed at 180 °C for 16 h.
Catalysts 12 01211 g001
Figure 2. The EDX data of Weight % and Atomic % of S and Cd elements in the sample.
Figure 2. The EDX data of Weight % and Atomic % of S and Cd elements in the sample.
Catalysts 12 01211 g002
Figure 3. XRD patterns of the anatase-phase TiO2 and TiO2@CdS particles.
Figure 3. XRD patterns of the anatase-phase TiO2 and TiO2@CdS particles.
Catalysts 12 01211 g003
Figure 4. Schematic process for durian-like TiO2@CdS core-shell structure.
Figure 4. Schematic process for durian-like TiO2@CdS core-shell structure.
Catalysts 12 01211 g004
Figure 5. The diagram of H2 generation rate of TiO2@CdS four cycle reaction.
Figure 5. The diagram of H2 generation rate of TiO2@CdS four cycle reaction.
Catalysts 12 01211 g005
Figure 6. (a) SEM image of pure CdS, (b) SEM image of pure CdS after reaction, (c) SEM image of TiO2@CdS, (d) SEM image of TiO2@CdS after reaction, (e) XRD patterns of TiO2@CdS before and after the reaction, (f) Illustration of the photocatalytic mechanism, (g) The photocurrent measurements over TiO2, CdS and TiO2@CdS (16 h).
Figure 6. (a) SEM image of pure CdS, (b) SEM image of pure CdS after reaction, (c) SEM image of TiO2@CdS, (d) SEM image of TiO2@CdS after reaction, (e) XRD patterns of TiO2@CdS before and after the reaction, (f) Illustration of the photocatalytic mechanism, (g) The photocurrent measurements over TiO2, CdS and TiO2@CdS (16 h).
Catalysts 12 01211 g006
Table 1. The EDX data of Weight % and Atomic % of S and K Element in the sample.
Table 1. The EDX data of Weight % and Atomic % of S and K Element in the sample.
ElementWeight %Atomic %
S K23.2551.50
Cd L76.7548.50
Totals100.00100.00
Table 2. H2 generation rate of different samples.
Table 2. H2 generation rate of different samples.
SampleH2 Generation Rate, μmol·h−1·g−1
TiO2400
TiO2@CdS (8 h)588
TiO2@CdS (16 h)1065
CdS3550
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, D.; Chen, Z.; Wang, X.; Zhong, Z.; Chen, C.; Wu, M. Synthesis of Durian-like TiO2@CdS Core-Shell Structure and Study on H2 Generation Properties. Catalysts 2022, 12, 1211. https://doi.org/10.3390/catal12101211

AMA Style

Li D, Chen Z, Wang X, Zhong Z, Chen C, Wu M. Synthesis of Durian-like TiO2@CdS Core-Shell Structure and Study on H2 Generation Properties. Catalysts. 2022; 12(10):1211. https://doi.org/10.3390/catal12101211

Chicago/Turabian Style

Li, Dongping, Zeheng Chen, Xin Wang, Zhenhong Zhong, Chunjun Chen, and Mengling Wu. 2022. "Synthesis of Durian-like TiO2@CdS Core-Shell Structure and Study on H2 Generation Properties" Catalysts 12, no. 10: 1211. https://doi.org/10.3390/catal12101211

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop