Next Article in Journal
Comparative Study of CFD and LedaFlow Models for Riser-Induced Slug Flow
Next Article in Special Issue
Evaporated MAPbI3 Perovskite Planar Solar Cells with Different Annealing Temperature
Previous Article in Journal
Effects of Injector Spray Angle on Performance of an Opposed-Piston Free-Piston Engine
Previous Article in Special Issue
Geometry Optimization of Top Metallic Contacts in a Solar Cell Using the Constructal Design Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Histidine Decorated Nanoparticles of CdS for Highly Efficient H2 Production via Water Splitting

1
Graduate School of Science and Engineering, Yamagata University, 4-3-16 Jonan, Yonezawa 992-8510, Japan
2
Faculty of Science, Yamagata University, 1-4-12 Kojiragawa-machi, Yamagata 992-8560, Japan
*
Author to whom correspondence should be addressed.
Energies 2020, 13(14), 3738; https://doi.org/10.3390/en13143738
Submission received: 30 June 2020 / Revised: 14 July 2020 / Accepted: 18 July 2020 / Published: 20 July 2020
(This article belongs to the Special Issue Nanotechnology for Solar Energy Conversion)

Abstract

:
Pure cadmium sulfide and histidine decorated cadmium sulfide nanocomposites are prepared by the hydrothermal or solvothermal method. Scanning electron microscopy (SEM) analysis shows that the particle sizes of pure cadmium sulfide (pu/CdS) and histidine decorated cadmium sulfide prepared by the hydrothermal method (hi/CdS) range from 0.75 to 3.0 μm. However, when a solvothermal method is used, the particle size of histidine decorated cadmium sulfide (so/CdS) ranges from 50 to 300 nm. X-ray diffraction (XRD) patterns show that all samples (pu/CdS, hi/CdS and so/CdS) have a hexagonal wurtzite crystal structure but so/CdS has a poor crystallinity compared to the others. The as-prepared samples are applied to photocatalytic hydrogen production via water splitting and the results show that the highest H2 evolution rate for pu/CdS and hi/CdS are 1250 and 1950 μmol·g−1·h−1, respectively. On the other hand, the so/CdS sample has a rate of 6020 μmol·g−1·h−1, which is about five times higher than that of the pu/CdS sample. The increased specific surface area of so/CdS nanoparticles and effective charge separation by histidine molecules are attributed to the improved H2 evolution.

1. Introduction

Hydrogen production from photocatalytic water splitting by using semiconductor photocatalysts has become one of the most promising ways of clean energy generation. For an efficient system of water splitting a semiconductor, photocatalyst requires to have a narrow bandgap (<3.0 eV) and a relatively high conduction band potential to allow the absorption of visible light and the reduction of water, respectively [1,2]. Among various semiconductor photocatalysts, a narrow bandgap of 2.4 eV makes cadmium sulfide (CdS) a visible-light-sensitive photocatalyst. Moreover, the flat band potential of CdS (−0.9 V vs. NHE) is more negative than the reduction potential of H+/H2 (0 V vs. NHE) [3]. These two criteria enable CdS an excellent candidate to be studied for photocatalytic water splitting. However, low energy conversion efficiency and severe photo corrosion under light irradiation inhibit its practical application in water splitting as well as other photocatalytic systems [4,5,6].
To improve its energy conversion efficiency, many researchers have reported the fabrication of nanoparticles with different shapes and sizes [7,8,9,10]. Miniaturization of the CdS particles into nanoscales resulted in a high specific surface area and an improved light-harvesting ability, which could enhance their photocatalytic activities. On the other hand, a stable H2 evolution can be obtained over CdS photocatalyst in the presence of S2− and SO32− as election donors, which come from sulfur-based sacrificial reagents. However, the stability is still insufficient and for an enhanced stability some researchers have reported the fabrication of nanocomposites in which a material or molecules is introduced to CdS nanoparticles. The addition of such material also can improve the specific surface area, light absorption, and charge separation in CdS leading to improved photocatalytic activities. For example, Zhang et al. [11] reported ferrous oxalate dehydrate (FOD)/CdS composites, which shows 5.9 times higher photocatalytic activity compared to pure CdS. In the case of pure CdS, the activity was decreased by 36.3% after the fourth cycle of hydrogen production, whereas for FOD/CdS only 5.2% hydrogen production was decreased after the fourth cycle. Lu et al. have reported PANI/CdS nanocomposite supported on SiO2 hydrogen for the enhanced and stable photocatalytic activity of CdS. Their results showed that the recovery rate of the composite catalyst reached 90%, and hydrogen production efficiency remained unchanged after five cycles, indicating excellent stability of CdS. They have shown that the conductive polymer PANI played the role of charge separation and effectively inhibited the photocorrosion of CdS [12]. Recently, Wang et al. [13] have shown that the addition of histidine to CdS can improve its photocatalytic activity and simultaneously reduce its photo corrosion. After consecutive stable H2 production for four cycles, the amount of H2 production was decreased by 12%. Although the detailed mechanism of improved activity of CdS is not clear yet, it is assumed that the lone pair electrons in histidine molecule can interact with the photogenerated holes in CdS, which leads to enhanced charge separation as well as reduced photo corrosion. However, the performance of the photocatalyst in still low for its practical application in H2 production. One of the reasons for low performance might be its large particle size (2~5 μm) and small surface area (3.26 m2/g). To improve the hydrogen production ability, we attempted to increase the surface area of histidine/CdS composite by miniaturization of the particles. Here, we report the successful synthesis of histidine decorated CdS nanoparticles by a solvothermal method with about three times higher hydrogen production compared to the report by Wang et al. [13]. It is worth to note that in most cases, toxic chemicals are used to synthesize the nanoparticles of CdS [2,7,14,15,16,17,18]. But in this report, we used ethylene glycol (EG) as a reaction solvent, which is widely used, cheap and non-toxic to the environment [19,20]. Moreover, the present method is easy and a low-temperature process and thus friendly to the environment.

2. Materials and Methods

Cadmium nitrate tetrahydrate (Cd(NO3)2·4H2O, thiourea (H2NCSNH2), histidine (C6H9N3O2), Ethylene glycol ((HOCH2CH2OH) were purchased from FUJIFILM Wako Pure Chemical Co. Ltd. All chemicals were of analytical grade and used as received without further purification. For all experimental works, deionized and doubly distilled water was used.
Cadmium sulfide (CdS) and L-histidine decorated CdS are prepared by a low-temperature hydrothermal or solvothermal method, as shown in Figure 1a. In a typical experiment, Cd(NO3)2·4H2O (0.005 mmol), histidine (0.0015 mmol) and thiourea (0.015 mmol) were dissolved in water or/and ethylene glycol (EG). The solution is stirred for another 60 min and then transferred into a Teflon-lined autoclave followed by heating at 180 °C for 4 h. Finally, the product was washed with plenty of water and ethanol and dried in a furnace at 80 °C for 8 h.
CdS samples with histidine (+H) and without histidine (−H) were synthesized using different amounts of EG (0~100%) to water as a reaction solvent, and the samples were denoted as EG0+H, EG25+H, EG50+H, EG75+H, EG100+H, EG100-H. Platinum (Pt, 0.3~4.8 wt%) was deposited on CdS nanoparticles by the conventional photoreduction method. Briefly, the as-prepared CdS sample and required amount of H2PtCl6 was dissolved in an aqueous NaOH solution. The solution was irradiated by light for 30 min to 300 min under the Ar atmosphere. After irradiation, the filtered Pt-loaded sample was washed with distilled water and ethanol and dried in a desiccator.
For hydrogen production measurements, 30 mg of CdS photocatalyst was dispersed in an aqueous solution (Figure 1b) of Na2SO3 (0.1 M) and Na2S (0.1 M) as a sacrificial reagent. The catalyst was maintained in suspension by a magnetic stirrer, and the top of the flask was sealed with a silicone rubber septum. To remove oxygen, the suspended solution was bubbled with Ar gas (about 5 mL/min) for 1 h with stirring. The light irradiation was carried out under an Ar atmosphere by using a xenon lamp (Ushio, 300 W) equipped with an optical filter (λ > 390 nm). The evolved gas was sampled through the silicone rubber septum by using a locking-type syringe at a constant time interval, and the gas was quantitatively analyzed by a gas chromatography (GG-8A, Shimadzu corp., detector: thermal conductivity detector (TCD), column packing: molecular sieve 5 Å, carrier gas: Ar) at 25 °C.
The X-ray diffraction (XRD) patterns were recorded using an X-ray diffractometer (Rigaku Ultima IV) with a Cu X-ray source (Wavelength λ: Kα = 1.544426 Å). The morphology of the catalyst was examined using field emission scanning electron microscopy (FE-SEM; JSM-7600FA, JEOL Ltd., Akishima, Japan) with an accelerating voltage of 15 kV. For the particle size distribution, sizes of about 150~300 particles in several SEM images for each sample were calculated. The specific surface areas of the photocatalysts were determined by applying the Brunauer-Emmett-Teller (BET) method to the N2 adsorption/desorption isotherms measured by a BELsorp-max (BEL Japan Inc., Osaka, Japan). The pore size and pore volume the samples were determined from the nitrogen adsorption-desorption isotherm by the Barrett-Joyner-Halenda (BJH) method. UV-visible diffuse reflectance spectra and photoluminescence spectra were recorded using a spectrophotometer (UV-2700, Shimadzu Corp., Kyoto, Japan) and photoluminescence spectroscopy (RF-5300PC, Shimadzu Corp., Kyoto, Japan), respectively.

3. Results and Discussion

Figure 2 shows the SEM images of CdS samples prepared by water or/and ethylene glycol as a reaction solvent. It is seen that the samples prepared by using 100% water solvent (EG0+H and EG0-H) have a particle size range of 0.75~3.0 μm (Figure 2a,b). With increasing the amount of EG to water, particle sizes gradually decrease. The particle size distribution of EG100+H sample is shown in Figure 2h and the average particle size was calculated to be 105 nm. It should be noted that the particle size distributions of (EG100+H) and (EG100-H) samples were almost similar (see Figure 2g,h). This result indicates that histidine does not have a considerable effect on the particle size of CdS.
However, the reason for the miniaturization of particle size might be explained in terms of the dielectric constant of the solvent. The dielectric constant of water is 80.1, which is much higher than that of EG (38.66) [21]. Since Cd2+ and S2− ions are hardly saturated in a high dielectric constant solution, the monomer concentration of CdS is lower in a water solvent. In this case, the growth of CdS crystals is controlled by the thermodynamic equilibrium [15] and due to the minimization of surface tension, larger particles are formed. With an increase in the amount of EG to water, dielectric constants of the solutions decrease. The growth of CdS particle in the low dielectric constant solvent is controlled by the kinetics [8] because the solubility of Cd2− and S2− in high dielectric constant solution becomes limited and solvent can be saturated more rapidly. The high specific surface tension could slow the formation of CdS crystal leading small particles. Similar results were observed in other reports, too [22].
XRD analysis was used to investigate the crystal phases of the samples. The XRD patterns, shown in Figure 3, can be readily indexed as the (100), (110), (103) and (122) plane of the hexagonal phase of the wurtzite CdS structure (space group: P63mc) with lattice parameters α = 4.141 Å and c = 6.720 Å, which match well with the standard values (JCPDS No: 41-1049). However, the CdS samples prepared by water only (EG0-H and EG0+H), a small reflection (marked by ∇) corresponding to the (200) plane of cubic phase (JCPDS 80-0019) appears at 2θ = 30.58°. This result indicates that these samples have mixed hexagonal and cubic crystals. Interestingly, even with the addition of only 25 vol% of EG to water solvent, the peak at 2θ = 30.58° is disappeared. Murugan et al. have shown that EG with lower dielectric constant can minimize the free energy difference between the hexagonal and the cubic phases, and thus favor the formation of pure hexagonal phase CdS [16].
However, with increasing the amount of EG, the peak intensities of the sample decrease, which implies that the crystallinity of the nanoparticles decreases gradually. Li et al. [23] reported that a decrease in particle size results in peak broadening and poor crystallization of the particle. Debye-Scherrer formula was used to calculate the crystal sizes of the samples by using the full width at half maximum of 2θ = 43.69° (110) and the results are shown in Table 1. It is seen that with increasing EG content to water, crystal sizes decrease. Since the crystal sizes of CdS samples are smaller than their particle sizes, it can be concluded that a particle consists of several CdS crystals.
UV-visible absorption spectra for CdS samples are shown in Figure 4. The absorption intensity EG0-H and EG0+H shows a rapid increase at ca. 550 nm. Applying the equation for bandgap EG = 1240/λ, the bandgap for both samples is found to be 2.25 eV. However, it is seen that with increasing the amount of EG as a reaction solvent, the absorption edges show blueshifts. The calculate bandgaps and the corresponding wavelengths (used to calculate the bandgaps) of the samples are tabulated in Table 2. For the sample, EG100+H, the calculated bandgap is 2.56 eV. The widening of the bandgap could be attributed to quantum confine effect due to the miniaturization of the particles. This quantum confines effect is common in semiconductor nanoparticles and consistent with the other reports [17,24].
To give further insight into the surface properties (surface area, pore size, pore volume) of the as-prepared photocatalysts, N2 adsorption-desorption isotherms, and pore-size distribution were investigated. Figure 5a shows the N2 adsorption-desorption isotherms of selected as-prepared CdS samples. According to Brunauer-Deming-Deming-Teller classification of N2 adsorption-desorption isotherms, the majority of physisorption isotherms can be grouped into six types [25]. EG0+H sample has an isotherm of type II, indicating the presence of micropores. On the other hand, EG100-H and EG100+H samples have isotherms of type IV and the shape of the hysteresis loops are of type H3, which indicate the presence of mesopores with a low contribution of micropores [2].
The pore size distribution of the corresponding samples, obtained by the BJH method, are shown in Figure 5b. Broad pore size distributions were found for both EG100-H and EG100+H samples. The pore volume for the EG100-H and EG100+H was found to be 0.074 and 0.11 cm3g−1, respectively. The BET surface areas of the EG0+H, EG100-H, EG100+H samples are calculated to be 3.7, 16.65, and 23.6 m2/g, respectively. The surface area for EG0+H is the same as reported by Wang [13]. However, both samples prepared by the solvothermal method have a larger surface area compared to that prepared by the hydrothermal method. The increased surface area is attributed to decreased particle sizes and the presence of mesopores in the samples. Generally, the higher specific surface area provides a higher number of active sites for photocatalytic reactions and thus a crucial factor in the enhancement of catalytic performance of the material [13,23,26].
The photocatalytic performance of the samples was evaluated in terms of H2 production under visible light irradiation. The irradiation time vs. amounts of H2 production curves for various samples (with 0.3 wt% Pt deposition) are shown in Figure 6a. It is seen that hydrogen production from all samples increases linearly with irradiation time. Moreover, the amounts of H2 increase with increasing the EG content during CdS sample synthesis. After two hours of light irradiation, the amount of produced gas from EG0+H and EG100+H samples were 400 μmol·g−1 and 1200 μmol·g−1, respectively. This means that H2 production is increased by three times when 100% EG is used for histidine/CdS composite preparation. This increment is attributed to the miniaturization of particles and the increased specific surface area of the samples.
The time of Pt photo-deposition also found to affect the amount of H2 evolution [3]. Therefore, to optimize the Pt photo deposition, the irradiation time was changed from 30 min to 300 min (keeping the Pt content fixed to 0.3 wt%), and the results are shown in Figure 6b. With increasing the Pt contents to CdS, H2 production increases continuously. However, the increment after 180 min was negligible, and therefore, we decided to keep the irradiation time for Pt photo-deposition to be fixed at 180 min.
To observe the effect of Pt content on H2 evolution, the amounts of Pt-loading on EG100+H sample were changed from 0.3 wt% to 5 wt%. The results in Figure 7 show that with increasing the Pt contents, H2 production also increases. With increasing the Pt content, effective charge separation, as well as the number of active sites for the reactions ise increased. As a result, H2 evolution increases. The highest rate of the H2 production observed for the sample with 4.2 wt% of Pt. Above this content of Pt, H2 production was not increased considerably. It is known that the excess amount of Pt deposition inhibits the light absorption by the photocatalyst [7]. This trend of Pt co-catalyst is consistent with the other literature [3,18,27,28], and similar trends are also observed in other co-catalysts [23,28,29] in the photocatalytic systems.
Finally, we made a comparison of H2 evolution from 4.2 wt% of Pt deposited EG100+H sample with EG0-H, EG0+H, and EG100-H samples and the results are shown in Figure 8. It is seen that the CdS sample with histidine (EG0+H) shows a higher (about 1.56 times) amount of H2 production compared to the CdS sample without histidine (EG0-H). This result indicates that histidine has a synergic effect of improving the photocatalytic activity of CdS, which is discussed later in this manuscript. However, even without histidine, the CdS sample prepared by using 100 % EG (EG100-H) has about two times higher H2 production than the EG0+H sample.
The improvement of H2 production from EG prepared samples could be explained by the following points—(1) the increased specific surface area by the miniaturization of the particles resulted in an increased number of active sites for reactions; (2) Widening of bandgap might lead shifting of CdS conduction band to more negative in the redox potential. Shifting of band edge by bandgap widening could increase the rate of electron transfer from the photocatalyst to the solution, resulting in the improved photocatalytic performance [30]. Similar trends are also observed in other photocatalytic systems [2,31,32]; (3) It is reported that in the case of small or nanoporous particles, the distance that photogenerated electrons and holes have to migrate to reaction sites on the surface is shortened, which benefits the suppression of the electron/hole recombination in bulk [1,23].
The sample EG100+H shows a further improvement of H2 production, which implies that the synergic effect of histidine is not affected by EG solvent used in our method. The sample with histidine (EG100+H) shows more than three times higher H2 evolution compared to that of EG0+H. We compared this hydrogen production rate (6020 μmol·h−1·g−1) with the other 18 reports [2,7,8,9,10,13,14,33] that used only a single semiconductor photocatalyst (CdS) system. It was found that our material shows the third-highest hydrogen evolution rate. By improving the crystallinity of EG100+H sample or combining with a second semiconductor material (e.g., TiO2), we are expecting a further improvement of hydrogen evolution, and the detail will be published in our next report.
However, the reason of improved H2 production by histidine is not clear yet but Wang et al. proposed that the cause of improved H2 production could be the improved charge separation by the interaction of photogenerated hole in CdS and the lone pair electrons of amidogen and hydroxyl (-NH, -NH2 and -OH) groups in histidine, which is introduced in CdS [13]. This fact can be confirmed by the photoluminescence spectra for CdS nanoparticles with (E100+H) and without histidine (EG100-H). As shown in Figure 9, both samples show an emission peak at 470 nm, which is attributed to the band-edge emission of CdS nanocrystals [34] and the broad peak between 510–560 is commonly attributed to the recombination of the charge carriers in the surface states [35]. The lower emission intensity for EG100+H sample implies the better charge separation of photogenerated electrons and holes pairs of the semiconductor photocatalyst.

4. Conclusions

We successfully synthesized nanoparticles of CdS photocatalyst by the solvothermal method by using ethylene glycol. The as-prepared nanoparticles show increased specific surface area and bandgap widening, which leads to improved photocatalytic activity in terms of hydrogen production via water splitting. Moreover, the addition of histidine to CdS can further improve hydrogen production by effective charge separation in CdS. The maximum amount of H2 evolution is observed for the CdS nanocomposite prepared by using 100% ethylene glycol as a reaction solvent. This CdS nanocomposite shows five times higher H2 production compared to pure CdS. By improving the crystallinity by controlling the synthesis parameters of CdS nanoparticles, further improvement in H2 production can be expected. Further research is underway in our laboratory.

Author Contributions

Conceptualization, B.A.; investigation, F.T., M.I. and B.A.; resources, S.K., M.K. and F.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partially supported by JSPS KAKENHI (20K05678) and the Kurata Grants from the Hitachi Global Foundation, Japan.

Acknowledgments

This work was partly supported by Nanotechnology Platform Program (Molecule and Material Synthesis) the Ministry of Education, Culture, Sports, Science and Technology (MEXT), Japan.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kudo, A.; Miseki, Y. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef] [PubMed]
  2. Vaquero, F.; Navarro, R.M.; Fierro, J.L.G. Evolution of the nanostructure of CdS using solvothermal synthesis at different temperature and its influence on the photoactivity for hydrogen production. Int. J. Hydrog. Energy 2016, 41, 11558–11567. [Google Scholar] [CrossRef]
  3. Wang, Y.; Wang, Y.; Xu, R. Photochemical deposition of Pt on CdS for H2 evolution from water: Markedly enhance activity by controlling Pt reduction environment. J. Phys. Chem. C 2013, 117, 783–790. [Google Scholar] [CrossRef]
  4. Lingampalli, S.S.R.; Gautam, U.K.; Rao, C.N.R. Highly efficient photocatalytic hydrogen generation by solution-processed ZnO/Pt/CdS, ZnO/Pt/Cd1−xZnxS and ZnO/Pt/CdS1−xSex hybrid nanostructures. Energy Environ. Sci. 2013, 6, 2589–3594. [Google Scholar] [CrossRef]
  5. Kalisman, P.; Nakibli, Y.; Amirav, L. Perfect photo-to-hydrogen conversion efficiency. Nanol Lett. 2016, 16, 1776–1781. [Google Scholar] [CrossRef]
  6. Ma, D.; Shi, J.-W.; Zou, Y.; Fan, Z.; Ji, X.; Niu, C.; Wang, L. Rational design of CdS@ZnO core-shell structure via atomic layer deposition for drastically enhanced photocatalytic H2 evolution with excellent photostability. Nano Energy 2017, 39, 183–191. [Google Scholar] [CrossRef] [Green Version]
  7. Bao, N.; Shen, L.; Takata, T.; Domen, K. Self-templated synthesis of nanoporous CdS nanostructures for highly efficient photocatalytic hydrogen production under visible light. Chem. Matter. 2008, 20, 110–117. [Google Scholar] [CrossRef]
  8. Zhang, B.; Yao, W.; Huang, C.; Xu, Q.; Wu, Q. Shape effects of CdS photocatalysts on hydrogen production. Int. J. Hydrog. Energy 2013, 38, 7224–7231. [Google Scholar] [CrossRef]
  9. Li, C.; Yuan, J.; Han, B.; Shangguan, W. Synthesis and photochemical performance of morphology-controlled CdS photocatalysts for hydrogen evolution under visible light. Int. J. Hydrog. Energy 2011, 36, 4271–4279. [Google Scholar] [CrossRef]
  10. Muruganandham, M.; Kusumoto, Y.; Okamoto, C.; Muruganandham, A.; Mamun, M.A.; Ahmmad, B. Mineralizer-assisted shape-controlled synthesis, characterization, and photocatalytic evaluation of CdS microcrystals. J. Phys. Chem. C 2009, 113, 19506–19517. [Google Scholar] [CrossRef]
  11. Zhang, L.; Hao, X.; Jian, Q.; Jin, Z. Ferrous oxalate dehydrate over CdS as Z-scheme photo catalytic hydrogen evolution. J. Solid State Chem. 2019, 274, 286–294. [Google Scholar] [CrossRef]
  12. Lu, J.; Zhang, X.; Gao, H.; Cui, W. Three-dimensional structure of PANI/CdS NRs-SiO2 hydrogen for photocatalytic hydrogen evolution with high activity and stability. Nanomaterials 2019, 9, 427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Wang, Q.; Lian, J.; Li, J.; Wang, R.; Huan, H.; Su, B.; Lei, Z. Highly efficient photocatalytic hydrogen production of flower-like cadmium sulfide decorated by histidine. Sci. Rep. 2015, 5, 13593. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Li, Y.; Hu, Y.; Lu, G.; Li, S. Synthesis of CdS nanorods by an ethylenediamine assisted hydrothermal method for photo catalytic hydrogen evolution. J. Phys. Chem. C 2009, 113, 9352–9358. [Google Scholar] [CrossRef]
  15. Qingqing, W.; Gang, X.; Gaorong, H. Solvothermal synthesis and characterization of uniform CdS nanowires in high yield. J. Solid State Chem. 2005, 178, 2680–2685. [Google Scholar] [CrossRef]
  16. Murugan, A.V.; Sonawane, R.S.; Kale, B.B.; Apte, S.K.; Kulkarni, A.V. Microwave-solvothermal synthesis of nanocrystalline cadmium sulfide. Mater. Chem. Phys. 2001, 71, 98–102. [Google Scholar] [CrossRef]
  17. Xu, D.; Liu, Z.; Liang, J.; Qian, Y. Solvothermal synthesis of CdS nanowires in a mixed solvent of ethylenediamine and dodecanethiol. J. Phys. Chem. B 2005, 109, 14344–14349. [Google Scholar] [CrossRef] [PubMed]
  18. Vamvasakis, I.; Liu, B.; Armatas, G.S. Size effects of platinum nanoparticles in the photocatalytic hydrogen production over 3D mesoporous networks of CdS and Pt nanojunctions. Adv. Funct. Mater. 2016, 26, 8052–8071. [Google Scholar] [CrossRef]
  19. Kauthale, S.S.; Tekale, S.U.; Jadhav, K.M.; Pawar, R.P. Ethylene glycol promoted catalyst-free pseudo three-component green synthesis of bis(coumarin)s and bis(3-methyl-1-phenyl-1 H-pyrozol-5-ol)s. Mol. Divers. 2016, 20, 763–770. [Google Scholar] [CrossRef]
  20. Verma, S.; Jain, S.L. Complexation of KBr3 with poly(ethylene glycol): Efficient bromination of aromatics under solvent-less conditions. Org. Chem. Curr. Res. 2012, 1, 100105. [Google Scholar] [CrossRef] [Green Version]
  21. Akerlof, G. Dielectric constants of some organic solvent-water mixture at various temperatures. J. Am. Chem. Soc. 1932, 54, 4125–4139. [Google Scholar] [CrossRef]
  22. Saravanan, L.; Jayavel, R.; Aldeyab, S.S.; Zaidi, J.S.M.; Ariga, K.; Vinu, A. Synthesis and morphological control of europium doped cadmium sulfide nanocrystals. J. Nanosci. Nanotechnol. 2011, 11, 7783–7788. [Google Scholar] [CrossRef] [PubMed]
  23. Li, Q.; Guo, B.; Yu, J.; Ran, J.; Zhang, B.; Yan, H.; Gong, J.R. Highly efficient visible-light-driven photocatalytic hydrogen production of CdS-Cluster-decorated graphene nanosheets. J. Am. Chem. Soc. 2011, 133, 10878–10884. [Google Scholar] [CrossRef] [PubMed]
  24. Qiu, W.; Xu, M.; Yang, X.; Chen, F.; Nan, Y.; Chen, H. Novel hierarchical CdS crystals by an amino acid mediated hydrothermal process. J. Alloys Compd. 2011, 509, 8413–8420. [Google Scholar] [CrossRef]
  25. Sing, K.S.W.; Everett, D.H.; Haul, R.A.W.; Moscou, L.; Pierotti, R.A.; Rouquerol, J.; Siemieniewska, T. Reporting physisorption data for gas/solid systems with special reference to the determination of surface area and porosity. Pure Appl. Chem. 1985, 57, 603–619. [Google Scholar] [CrossRef]
  26. Yu, J.G.; Yu, H.G.; Cheng, B.; Zhou, M.H.; Zhao, X.J. Enhanced photo catalytic activity of TiO2 powder (P25) by hydrothermal treatment. J. Mol. Catal. A Chem. 2006, 253, 112. [Google Scholar] [CrossRef]
  27. Ikuma, Y.; Bessho, H. Effect of Pt concentration on the production of hydrogen by a TiO2 photocatalyst. Int. J. Hydrog. Energy 2007, 32, 2689–2692. [Google Scholar] [CrossRef]
  28. Oros-Ruiz, S.; Hernandez-Gordillo, A.; Garcia-Mendoza, C.; Rodriguez-Rodriguez, A.A.; Gomez, R. Comparative activity of CdS nanofibers superficially modified by Au, Cu, Ni nanoparticles as co-catalysts for photocatalytic hydrogen production under visible light. J. Chem. Technol. Biotechnol. 2016, 91, 2205–2210. [Google Scholar] [CrossRef]
  29. Ahmmad, B.; Kusumoto, Y.; Somekawa, S.; Ikeda, M. Carbon nanotubes synergistically enhance photocatalytic activity of TiO2. Catal. Commun. 2008, 9, 1410–1413. [Google Scholar] [CrossRef]
  30. Gerischer, H. The impact of semiconductors on the concepts of electrochemistry. Electrochim. Acta 1990, 35, 1677–1699. [Google Scholar] [CrossRef]
  31. Xu, J.; Wang, L.; Cao, X. Polymer supported graphene-CdS composite catalyst with enhanced photocatalytic hydrogen production from water splitting under visible light. Chem. Eng. J. 2016, 238, 816–825. [Google Scholar] [CrossRef]
  32. Holmes, M.A.; Townsend, T.K.; Osterloh, F. Quantum confinement controlled photocatalytic water splitting by suspended CdSe nanocrystals. Chem. Commun. 2012, 48, 371–373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Zhang, K.; Guo, L. Metal sulphide semiconductors for photo catalytic hydrogen production. Catal. Sci. Technol. 2013, 3, 1672–1690. [Google Scholar] [CrossRef]
  34. Xiao, Q.; Xiao, C.; Ouyang, L. Strong enhancement of band-edge photoluminescence in CdS nanocrystals prepared by one-step aqueous synthesis method. J. Lumin. 2008, 128, 1942–1947. [Google Scholar] [CrossRef]
  35. Ghamsari, M.S.; Ghamsari, A.H.S. CdS colloidal nanocrystals with narrow green emission. J. Nanophotonics 2016, 10, 026007. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of material synthesis and H2 production.
Figure 1. Schematic diagram of material synthesis and H2 production.
Energies 13 03738 g001
Figure 2. Scanning electron microscopy (SEM) images of various cadmium sulfide (CdS) samples ((a): EG0-H; (b): EG0+H; (c): EG25+H; (d): EG50+H; (e): EG75+H; (f): EG100+H; (g): EG100-H) and the particle size distribution of EG100+H sample (h).
Figure 2. Scanning electron microscopy (SEM) images of various cadmium sulfide (CdS) samples ((a): EG0-H; (b): EG0+H; (c): EG25+H; (d): EG50+H; (e): EG75+H; (f): EG100+H; (g): EG100-H) and the particle size distribution of EG100+H sample (h).
Energies 13 03738 g002
Figure 3. X-ray diffraction (XRD) patterns of the CdS nanocomposites.
Figure 3. X-ray diffraction (XRD) patterns of the CdS nanocomposites.
Energies 13 03738 g003
Figure 4. XRD patterns of the CdS nanocomposites.
Figure 4. XRD patterns of the CdS nanocomposites.
Energies 13 03738 g004
Figure 5. N2 adsorption-desorption isotherms (a) and pore size distribution (b) for EG0-H, EG100-H and EG100+H samples.
Figure 5. N2 adsorption-desorption isotherms (a) and pore size distribution (b) for EG0-H, EG100-H and EG100+H samples.
Energies 13 03738 g005
Figure 6. Dependence of H2 evolution on (a) the amounts of EG to water during CdS nanocomposite synthesis and (b) the light irradiation time of Pt photodepostion on CdS nanocomposite.
Figure 6. Dependence of H2 evolution on (a) the amounts of EG to water during CdS nanocomposite synthesis and (b) the light irradiation time of Pt photodepostion on CdS nanocomposite.
Energies 13 03738 g006
Figure 7. Dependence of H2 evolution on the content of Pt cocatalyst with CdS nanocomposite.
Figure 7. Dependence of H2 evolution on the content of Pt cocatalyst with CdS nanocomposite.
Energies 13 03738 g007
Figure 8. Comparison of the amounts of the evolved H2 gas from various samples after 2 h of light irradiation.
Figure 8. Comparison of the amounts of the evolved H2 gas from various samples after 2 h of light irradiation.
Energies 13 03738 g008
Figure 9. Photoluminescence spectra of EG100-H and EG100+H samples.
Figure 9. Photoluminescence spectra of EG100-H and EG100+H samples.
Energies 13 03738 g009
Table 1. Crystal sizes of the different CdS samples.
Table 1. Crystal sizes of the different CdS samples.
SampleEG0-HEG0+HEG25+HEG50+HEG75+HEG100+HEG100-H
Crystal size (nm)44.342.328.022.720.310.111.7
Table 2. Calculated bandgaps and corresponding wavelengths of various CdS samples.
Table 2. Calculated bandgaps and corresponding wavelengths of various CdS samples.
Sample EG0-H EG0+H EG25+H EG50+H EG75+H EG100+H EG100-H
Wavelength (nm)550550535520510485490
Bandgap (eV)2.252.252.322.382.432.562.53

Share and Cite

MDPI and ACS Style

Tojo, F.; Ishizaki, M.; Kubota, S.; Kurihara, M.; Hirose, F.; Ahmmad, B. Histidine Decorated Nanoparticles of CdS for Highly Efficient H2 Production via Water Splitting. Energies 2020, 13, 3738. https://doi.org/10.3390/en13143738

AMA Style

Tojo F, Ishizaki M, Kubota S, Kurihara M, Hirose F, Ahmmad B. Histidine Decorated Nanoparticles of CdS for Highly Efficient H2 Production via Water Splitting. Energies. 2020; 13(14):3738. https://doi.org/10.3390/en13143738

Chicago/Turabian Style

Tojo, Fumiya, Manabu Ishizaki, Shigeru Kubota, Masato Kurihara, Fumihiko Hirose, and Bashir Ahmmad. 2020. "Histidine Decorated Nanoparticles of CdS for Highly Efficient H2 Production via Water Splitting" Energies 13, no. 14: 3738. https://doi.org/10.3390/en13143738

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