Next Article in Journal / Special Issue
Facile Construction of Bi2Sn2O7/g-C3N4 Heterojunction with Enhanced Photocatalytic Degradation of Norfloxacin
Previous Article in Journal
N-O Ligand Supported Stannylenes: Preparation, Crystal, and Molecular Structures
Previous Article in Special Issue
Rearrangement of Diferrocenyl 3,4-Thiophene Dicarboxylate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of CuO Loading on the Photocatalytic Activity of SrTiO3 for Hydrogen Evolution

1
Faculty of Chemistry, Thai Nguyen University of Education, No 20 Luong Ngoc Quyen Street, Thai Nguyen City 24000, Vietnam
2
Institute of Applied Technology and Sustainable Development, Nguyen Tat Thanh University, Ho Chi Minh City 700000, Vietnam
*
Authors to whom correspondence should be addressed.
Inorganics 2022, 10(9), 130; https://doi.org/10.3390/inorganics10090130
Submission received: 19 July 2022 / Revised: 22 August 2022 / Accepted: 29 August 2022 / Published: 31 August 2022
(This article belongs to the Special Issue Inorganics: 10th Anniversary)

Abstract

:
A CuO-loaded SrTiO3 catalyst showed highly photocatalytic activity for H2 evolution. This catalyst was prepared by an impregnation method and characterized by XRD, TEM, BET, XPS, Uv-vis DRS and PL techniques. Under optimum conditions, the best rate of H2 evolution of the CuO-loaded SrTiO3 catalyst is 5811 µmol h−1g−1, whereas it is a mere 34 µmol h−1g−1 for the pure SrTiO3. High efficiency, low cost and good stability are some of the merits that underline the promising potential of CuO-loaded SrTiO3 in the photocatalytic hydrogen.

1. Introduction

It is well known that photocatalytic hydrogen evolution from water splitting has received much attention in recent years for its potential application in providing hydrogen as a clean and renewable energy resource, even on a large scale [1,2,3,4,5,6]. Therefore, to find an efficient photocatalyst for hydrogen evolution is a vital topic.
Strontium titanate (SrTiO3) has typical perovskite structure with the advantages of low cost and excellent chemical stability. It has been widely used as a photocatalyst [7,8,9]. Although perovskite-type SrTiO3 appears to be a promising candidate for photocatalytic hydrogen evolution from water splitting since its suitable band structure for facilitating hydrogen and oxygen formation [10], the photocatalytic efficiency of pure SrTiO3 for hydrogen evolution is still very low, mainly due to the fast recombination of the photo-generated electrons and holes. It is necessary to modify the SrTiO3 particle to obtain an active photocatalyst for water splitting.
Up to now, many outstanding methods have been developed to improve the photocatalytic efficiency of SrTiO3 semiconductor for hydrogen evolution, such as doped with metals or non-metals [11,12,13,14,15,16,17,18], or coupled with other semiconductors [19,20,21,22,23,24,25,26,27], have been investigated to solve the above mentioned issues.
It is known that CuO can act as an efficient cocatalyst of TiO2 for its photocatalytic H2 production. The band gap energy (Eg) of SrTiO3 and TiO2 is similar (3.2 eV), and it is higher than that of CuO (1.7 eV). Therefore, it can be expected that the CuO may be an efficient cocatalyst for SrTiO3, which is meaningful for us to prepare a highly efficient, cheap and stable SrTiO3-based photocatalyst. In fact, several investigators have been focused on the visible photocatalysts based on SrTiO3 and CuO. Choudhary et al. have prepared CuO/SrTiO3 bilayered thin films by sol–gel spin-coating technique, which were used for a water splitting reaction. According to the literature, the bilayered system offered enhanced photoconversion efficiency, attributed to improved conductivity, which ameliorate separation of the photo-generated carriers at the CuO/STO interface and higher value of flatband potential [28]. Recently, Ahmadi et al. [29] synthesized CuO/SrTiO3 nanoparticles using a combination of impregnation and precipitation-deposition method, and the photocatalytic activity of CuO/SrTiO3 nanoparticles was evaluated by degradation of RhB under UV light irradiation. According to the mechanism in the literature, the CuO could help to separate the photo-generated electron-hole efficiently. Sepideh et al. [30] synthesized CuO/SrTiO3 composites. The prepared materials were studied as photocatalysts for the hydrogen evolution from aqueous methanol solution at an ambient temperature under UV light irradiation. In comparison to unmodified SrTiO3, the highest increase in photocatalytic activity was more than threefold, i.e., from 39 to 130 µmol h−1. According to the literature, the improvement of the photocatalytic activity of SrTiO3 for hydrogen evolution from aqueous methanol solution is attributed to the combination with CuO as a narrow band gap semiconductor to obtain CuO/SrTiO3 composites.
In the present work, we will explore the effect of CuO as a cocatalyst on the photocatalytic activity of the SrTiO3 photocatalyst for hydrogen evolution from aqueous solution containing various electron donors. The prepared CuO-SrTiO3 photocatalysts were characterized by XRD, TEM, EDX and XPS analysis. The DSR spectra were used to explore whether or not the effect of CuO on the band gap energy of the SrTiO3 semiconductor and the PL emission spectra were used to reveal the efficiency of trapping, transfer and separation of charge carriers, and to investigate their lifetime in the semiconductors. Moreover, the effects of some factors, such as type and concentration of electron donors, reaction temperature and photocatalyst concentration on the hydrogen evolution rate, will be systematically investigated in detailed. In addition, for comparison, the photocatalytic hydrogen evolution over Pt-SrTiO3 was also carried out under the same conditions with the case of CuO-SrTiO3 samples.

2. Results

2.1. Characterization of Photocatalyst

The XRD patterns of the pure SrTiO3 and the 1.5 wt.% CuO-loaded SrTiO3 are shown in Figure 1.
The result in Figure 1 shows that the X-ray diffraction pattern of the 1.5 wt.% CuO-SrTiO3 sample almost coincides with that of the pure SrTiO3. Diffraction peaks at 32.42°, 39.98°, 46.48°, 57.79°, 67.80° and 77.17° in all of XRD patterns are corresponded to the (110), (111), (200), (211), (220) and (310) planes of cubic SrTiO3 (JCPDS card No. 35-0734) [31], respectively. This result indicates that the obtained SrTiO3 possesses pure cubic phase. The phase of CuO cocatalyst is not observed in the XRD pattern of the 1.5 wt.% CuO-SrTiO3. It is possibly due to the concentration of the CuO cocatalyst in the photocatalyst is lower than that afforded by XRD detection limits or high dispersion of CuO on the SrTiO3 surface. The experimental result below also confirms this deduction.
As shown in TEM images of Figure 2, the introduction of CuO does not obviously change the morphology of SrTiO3 nanoparticles. The mean diameter of the CuO-loaded SrTiO3 is approximately 35 nm, and the surface of the CuO-loaded SrTiO3 nanoparticles is fairly smooth. It is indicated that the CuO particles are homogenously distributed through the surface of SrTiO3 supports. It is consistent with the result of XRD analysis.
The evidence for the composition of 1.5% CuO-SrTiO3 was obtained by EDX analysis as shown in Figure 2c. The result of EDX shows that the product is composed of the elements Sr, Ti, O and Cu. Further, the XPS analysis of 1.5% CuO-SrTiO3 sample was performed, and the survey spectrum and high-resolution scans are shown in Figure 3.
From the XPS survey spectrum shown in Figure 3a, Cu, Sr, Ti and O photoelectron lines from 1.5% CuO-SrTiO3 sample are detected along with C peak. It is consistent with EDX results. It should be noted that the carbon peak (C 1 s) in the survey spectra is attributed to the residual carbon from the sample and trace hydrocarbon from XPS instrument itself [32]. In Figure 3b, spin orbital splitting photoelectrons of Cu 2p3/2 and Cu 2p1/2 are located at 933.6 eV and 952.6 eV, respectively, which correspond to CuO [33,34]. A shake-up line among the binding energies of Cu 2p3/2 and Cu 2p1/2 further confirms this oxidation state, which is in a good agreement with the results of the literature reported. As shown in Figure 3c, the respective binding energies of Ti 2p3/2 and Ti 2p1/2 are located at 458.2 and 464.0 eV. The two bands are assigned to typical Ti4+ [35]. In Figure 3d, the binding energies of Sr d5/2 and Sr d3/2 at 133.5 and 134.4 eV, respectively, are assigned to typical Sr2+ [31]. The result of XPS analysis demonstrates that the Cu element in SrTiO3 is in the form of CuO.

2.2. Effect of Loading Amount of CuO on Hydrogen Evolution Activity over SrTiO3

In order to evaluate the influence of loading CuO on the photocatalytic activity for hydrogen evolution over SrTiO3 photocatalysts, the photocatalytic activity for hydrogen evolution is conducted over pure SrTiO3, the CuO-SrTiO3 photocatalysts and Pt-SrTiO3 photocatalysts under the same conditions, respectively. The amount of hydrogen evolution for the first 2 h is used as a comparative indicator of the hydrogen evolution activity. The experimental results are shown in Figure 4 and Table 1.
It is clear that the amount of hydrogen evolution over pure SrTiO3 is only 1.45 µmol. While the photocatalytic activity of the SrTiO3 photocatalyst for hydrogen evolution can be greatly enhanced by loading CuO cocatalyst. The maximum amounts of hydrogen evolution from 60 vol.% methanol aqueous solution over CuO-SrTiO3 (1.5 wt.% loaded) and Pt-SrTiO3 (2.0 wt.% loaded) are 259 µmol and 403 µmol, respectively. Although the photocatalytic activity of the CuO-SrTiO3 is lower than that of Pt-SrTiO3, its rate of H2 evolution is still more than 60% of that of SrTiO3 nanoparticles with optimum loading amount of Pt.
It can be seen that the CuO cocatalyst play critical roles in improving both the activity and stability of SrTiO3 photocatalysts in photocatalytic processes. Firstly, CuO can be contribute to the electron–hole separation at the CuO/SrTiO3 interface because, in the case of the SrTiO3 photocatalyst loaded with CuO cocatalyst, the photogenerated electrons can migrate to the surface of cocatalyst and the photogenerated holes are trapped in the SrTiO3 surface. Thus, the recombination rate of e/h+ pairs is decreased, and the rate of hydrogen evolution is increased. Secondly, CuO cocatalyst could decrease the activation energy or overpotential for redox reactions such as H2 and O2 evolution. Thirdly, CuO cocatalyst can inhibit photo-corrosion and improve the stability of semiconductor photocatalysts. In photocatalytic reactions, there are a number of visible-light-responsive semiconductors that are likely oxidized by photogenerated holes and causing their self-decomposition. The cocatalysts can extract the photogenerated holes and enhance the robustness of semiconductors [36,37]. These results imply that CuO is an efficient alternative to Pt for SrTiO3.
In Figure 4a, it shows that the amount of H2 evolution gradually increases from 166 µmol to 259 µmol with increasing the CuO content from 0.5 wt.% to 1.5 wt.%. However, followed with further increasing the CuO content from 1.5 to 2.5 wt.%, the amount of H2 evolution drops rapidly from 259 µmol to 200 µmol. From these results, it can be seen that an excess amount of CuO (beyond the optimum loading) increases the probability of a recombination reaction, leading to a decrease in the photocatalytic hydrogen evolution activity. Therefore, the optimum loading amount of CuO cocatalyst in SrTiO3 photocatalyst is to be 1.5 wt.% where the CuO-SrTiO3 photocatalyst shows the highest photocatalytic activity.

2.3. Effects of Type and Concentration of Electron Donors on Hydrogen Evolution over CuO-SrTiO3

It is known that electron donors are usually added into the reaction solution in order to improve the photocatalytic activity of photocatalyst by reducing the recombination of e/h+ pairs. 60 mg of 1.5% CuO-SrTiO3 photocatalyst is added into 60 mL of a solution containing various electron donors such as methanol (MeOH), glycerol (C3H5(OH)3), ethylenediaminetetraacetic acid (EDTA) or triethanolamine (TEOA) to investigate the effects of electron donors on hydrogen evolution over CuO-SrTiO3.
As shown in Figure 5, the photocatalytic hydrogen evolution activity over the 1.5% CuO-SrTiO3 photocatalyst by adding different electron donors is in the following order: methanol > EDTA > glycerol > TEOA. Very clearly, compared to the other electron donors studied, methanol is the most effective electron donor for the CuO-SrTiO3 photocatalyst. This is possibly due to its stronger ability of donating electrons to scavenge the valence band holes, which can effectively prevent photo-generated charge recombination. In the case of methanol as an electron donor, the amount of hydrogen evolution increases with enhancing concentration of methanol up to 60 vol.%. Beyond 60 vol.%, the hydrogen evolution rate increases little. An explanation of this result is that the reactions at the interface dominate the whole process of hydrogen evolution. Therefore, the rate of hydrogen evolution is limited by the amount of reactive species produced by methanol from the solution to the surface of the photocatalyst. When the concentration of methanol is lower than 60 vol.%, the adsorption of methanol on the surface of photocatalyst cannot saturate. Thus, the amount of reactive species does not yet reach maximum. When the concentration of methanol is beyond 60 vol.%, more methanol molecules are adsorbed on the surface of the photocatalyst. However, the relative amount of reactive species on the surface of the photocatalyst could not increase because the intensity of light and the amount of photocatalyst remain constant. Therefore, amount of hydrogen evolution enhances little followed with the increased amount of methanol. The most suitable concentration of methanol for hydrogen production over the 1.5% CuO-SrTiO3 photocatalyst is 60 vol.%. Based on the above results, 60 vol.% of methanol aqueous solution is used in following investigations.

2.4. Effect of Reaction Temperature on Hydrogen Evolution over CuO-SrTiO3

In order to determine the other suitable conditions for hydrogen production from the photocatalytic water splitting using methanol as the electron donor, the effect of reaction temperature on hydrogen evolution over CuO-SrTiO3 was studied and the results are shown in Figure 6.
It can be seen that the reaction temperature is a significant factor for the photocatalytic hydrogen evolution over the CuO-SrTiO3 photocatalyst from the methanol aqueous solution. The results in Figure 6 show that the photocatalytic activity for hydrogen evolution is increased when increasing the reaction temperature. The amount of hydrogen evolution can reach 444 µmol at 45 °C, which is 1.75 times higher than that at 25 °C. According to the reports of Puangpetcha and Korzhak [31,38], the effect of temperature on the photocatalytic hydrogen evolution reaction can be related to the thermal activation energy and desorption of the oxidation products from the sacrificial reagent (methanol).

2.5. Effect of Amount of Photocatalyst on Hydrogen Evolution

Figure 7 shows the effect of the amount of the 1.5% CuO-SrTiO3 photocatalyst on the photocatalytic hydrogen evolution activity.
The result in Figure 7 shows that the amount of hydrogen evolution increases from 187 µmol to 742 µmol when the addition amount of the 1.5% CuO-SrTiO3 is enhanced from 20 mg to 120 mg. In contrast, the amount of hydrogen evolution decreases from 742 µmol to 570 µmol with further increasing the amounts of the 1.5% CuO-SrTiO3 from 120 mg to 180 mg. Therefore, the maximum amount of hydrogen evolution is obtained when the addition amount of 1.5% CuO-SrTiO3 photocatalyst is 120 mg per 60 mL (2 g/L) in the present system. It could be explained that at a low concentration of photocatalyst, the photocatalytic reaction is mainly governed by active sites which are available for absorption of light and adsorption of reactant. The active sites are increased with increment of the concentration of the photocatalyst. However, when the concentration of the photocatalyst is above the optimum level, the reaction system becomes turbid, and UV light is greatly scattered by the suspended photocatalyst. Therefore, the transmission of UV light in suspension is greatly inhibited, and results in a sharp decrement in photocatalytic activity for hydrogen evolution.

2.6. Photocatalytic Hydrogen Evolution Activity of 1.5% CuO-SrTiO3

In order to evaluate photocatalytic hydrogen evolution activity of the 1.5% CuO-SrTiO3 photocatalyst from methanol aqueous solution, the relationship between amount of hydrogen evolved and irradiation time is investigated under optimum conditions for prolonged irradiation time, with the pure SrTiO3 as comparison. The experimental results are shown in Figure 8.
As illustrated in Figure 8, the amount of hydrogen evolution almost linearly increases followed with the prolonged irradiation time. After irradiation for 48 h, the amount of hydrogen evolution over 1.5% CuO-SrTiO3 is 13.31 mmol, whereas it is only 0.08 mmol for pure SrTiO3. The results presented above clearly indicate that loading CuO on the surface of SrTiO3 can dramatically enhance the photocatalytic activity of SrTiO3, and the CuO-SrTiO3 photocatalyst is highly stable and active for the reduction of H2O to H2 in the presence of methanol.
The specific surface areas of pure SrTiO3 and the 1.5 wt.% CuO-SrTiO3 are 16.2 and 15.5 m2g−1, respectively. Very clearly, loading CuO on SrTiO3 photocatalyst cannot change the specific surface area of SrTiO3. Thus, the enhancement of hydrogen production by loading CuO on SrTiO3 photocatalyst could not originate from the specific surface area of photocatalyst.
The UV–vis diffuse reflectance spectra of pure SrTiO3 and 1.5% CuO-SrTiO3 samples are shown in Figure 9. The absorption bands of the 1.5% CuO-SrTiO3 sample are not shifted compared to that of pure SrTiO3. Both of the samples exhibit an absorption band around 384 nm corresponding to band gap energy of 3.22 eV calculated from the formula Eg = 1240/λ [39]. Obviously, loading CuO on the surface of SrTiO3 does not change the band gap energy of the SrTiO3 semiconductor. The same result has also been reported in literature [40] for the case of the CuO-loaded SrTiO3 photocatalyst.
Moreover, the PL spectra of pure SrTiO3 and the 1.5% CuO-SrTiO3 are measured and shown in Figure 10. In general, if the amounts of the photoproduced electrons resulting from the recombination of excited electrons/holes are increased, the PL intensity of the sample increases. Consequently, its photocatalytic activity decreases. Therefore, there exists a close relationship between the PL intensity and photocatalytic activity of photocatalyst. As illustrated in Figure 10, the PL intensity of the 1.5% CuO-SrTiO3 is lower than that of pure SrTiO3. It is indicated that there exists a photo-generated electron transfer from SrTiO3 to CuO, which efficiently depresses the recombination of photoproduced electrons/holes in the photocatalyst.
Based on the above results and the literature reported [28,32], the great increment of the photocatalytic activity of 1.5% CuO-SrTiO3 ought to be ascribed to the efficient separation of electron/hole pairs. In order to better understand the effect of loading CuO on photocatalytic hydrogen evolution over the CuO-SrTiO3 photocatalyst from the methanol/water, a possible mechanism is developed. As can be seen in Figure 11, under UV light irradiation, the electrons in the valence band (VB) of the SrTiO3 are transferred to the conduction band (CB) of the SrTiO3. Due to the CB of CuO (+0.46 V) is situated below the CB of SrTiO3 (−0.29 V), the excited electrons in the CB of SrTiO3 can rapidly be transferred to the CB of CuO. The H+ transferred to the surface of CuO cocatalyst particles are reduced to H2 by the photogenerated electrons. As the electrons are captured by CuO, the recombination of electrons and holes is depressed. The holes produced on the valence band are allowed to oxidize H2O and methanol, which makes the photocatalytic reaction be able to continue.
It is noticed that the potential of the CuO is not suitable for hydrogen evolved from water splitting because the CB level of CuO semiconductors (+0.46 V) is more positive than E H 2 0 / H 2 = 0   V (vs. NHE, pH = 0) [34]. However, once the photoproduced electrons are transferred to CuO, the Fermi level of CuO is raised which results in a more negative CB potential of CuO. In other words, the accumulation of excess electrons in CuO can cause a negative shift of the CB potential of CuO in the end. Thus, the reduction of water can be carried out at the CB of the CuO as shown in Figure 12.

3. Materials and Methods

3.1. Materials

Strontium nitrate (Sr(NO3)2, Sigma-Aldrich, St. Louis, MO, USA), tetra-butyl titanate (TBOT Sigma-Aldrich), Cu(NO3)2 (Xilong Chemical Co., Ltd. Guangzhou, China). All other reagents were analytical grades and used without further purification.

3.2. Preparation of CuO-SrTiO3 Photocatalyst

The SrTiO3 nanoparticles loaded with various amounts of CuO (0.5, 1.0, 1.5, 2 and 2.5 wt.%) were prepared by impregnation method. A typical experimental procedure of 1.5 wt.% CuO-loaded SrTiO3 was described as followed: 200 mg of the SrTiO3 photocatalyst and 3.75 mL of 0.01 mol·L−1 Cu(NO3)2 solution were added into 10 mL distilled water. Afterwards, the suspension solution was stirred at room temperature for 4 h. Finally, the sample was dried at 80 °C for 12 h and calcined at 450 °C for 4 h. The other CuO-loaded SrTiO3 samples were prepared using similar procedures. The obtained products are denoted as x% CuO-SrTiO3 (x% is weight per cent of CuO).

3.3. Characterization Techniques

The structure and crystalline phase of the samples were characterized by Rigaku D/max 2550 VB/PC X-ray diffractometer (XRD) with Cu Kα radiation (λ = 0.154056 nm) at 40 kV and 40 mA (Tokyo, Japan). The microstructure and morphology of the samples were analyzed by JEOL, JEM-200CX transmission electron microscope (TEM) with 200 kV accelerating voltage (Tokyo, Japan). The X-ray photoelectron spectra (XPS) were recorded on a PHI 5000 Versaprobe spectrometer (ULVAC-PHI, Inc., Kanagawa, Japan). The energy dispersive X-ray spectrocopy (EDX) was taken with a JEOL JSM-6360LV electron microscopy (Tokyo, Japan). Fourier transform infrared spectra (FT-IR) were recorded on a Bruker Vector 22 spectrometer (resolution 4 cm−1, Yokohama, Japan), the samples being pressed into disks with KBr. The solid diffusion reflectance UV–vis spectra (DRS) were recorded on Unico UV-2102 PCS spectrometer (Tokyo, Japan). The photoluminescence spectra (PL) were recorded with a Shimadzu RF-5301PC fluorescence spectrometer (Tokyo, Japan). The N2 adsorption and desorption isotherms were measured on a Micromeritics ASAP-2020 nitrogen adsorption apparatus (Norcross, GA, USA).

3.4. Photocatalytic Water Splitting Experiments of CuO-SrTiO3 Photocatalyst

The photocatalytic reaction was carried out in a gas-closed system with a reactor made of quartz. In a typical experiment, 60 mg or 120 mg of the CuO-loaded SrTiO3 photocatalyst was added into 60 mL of the reaction solution containing various electron donors. Before photo-irradiation, the mixed solution was sonicated for 3 min, and then the suspended solution was bubbled with highly pure N2 gas (about 15 mL/min) for 30 min to remove dissolved oxygen gas in the suspension. The suspension was irradiated by a 300 W high- pressure Hg lamp with cooling water was circulated through a cylindrical quartz jacket located around the UV light source. To maintain a constant reactor temperature, the photo-catalytic reaction system was equipped with an electric fan and water with appropriate temperature was circulated around the quartz reactor. The amounts of H2 evolution were analyzed by a gas chromatograph (GC-112A, molecular sieve 5A, TCD, Chongqing Gold Equipment Co., Ltd., Chongqing, China) and N2 as a carrier gas.

4. Conclusions

Photocatalytic hydrogen evolution over CuO-loaded SrTiO3 photocatalyst from aqueous solution containing various electron donors has been studied under UV irradiation. The effects of various factors, such as loading amount of CuO, type and concentration of electron donors, reaction temperature and concentration of the photocatalyst have been systematically investigated in detail. The results showed that methanol was the most effective electron donor for photocatalytic hydrogen evolution over the CuO-SrTiO3 photocatalyst. The optimum reaction conditions for hydrogen evolution from methanol aqueous solution over CuO-SrTiO3 were found out. That is, the loading amount of CuO was 1.5 wt.%, the concentration of CuO-SrTiO3 photocatalyst was 2 g/L, the concentrating of methanol was 60 vol.%, and the reaction temperature was 45 °C. Importantly, it was found that the CuO loaded SrTiO3 nanoparticles were highly active and stable photocatalyst for the photocatalytic hydrogen production from methanol aqueous solution.

Author Contributions

Conceptualization, X.T.M. and D.N.B.; methodology, H.D.C. and T.H.L.N.; software, T.T.L.N.; formal analysis, H.D.C. and T.H.L.N.; data curation, T.T.L.N. and X.T.M.; writing—original draft preparation, D.N.B. and V.K.P.; writing—review and editing, V.K.P. and T.K.N.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported the MOET of Vietnam Project No. B2020-TNA-12.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the data are available within the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ganguly, P.; Harb, M.; Cao, Z.; Cavallo, L.; Breen, A.; Dervin, S.; Dionysiou, D.D.; Pillai, S.C. 2D Nanomaterials for Photocatalytic Hydrogen Production. ACS Energy Lett. 2019, 4, 1687–1709. [Google Scholar] [CrossRef]
  2. 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. Renew. Sustain. Energy Rev. 2007, 11, 401–425. [Google Scholar] [CrossRef]
  3. Holladay, J.D.; Hu, J.; King, D.L.; Wang, Y. An Overview of Hydrogen Production Technologies. Catal. Today 2009, 139, 244–260. [Google Scholar] [CrossRef]
  4. Navarro, R.M.; Sánchez-Sánchez, M.C.; Alvarez-Galvan, M.C.; del Valle, F.; Fierro, J.L.G. Hydrogen Production from Renewable Sources: Biomass and Photocatalytic Opportunities. Energy Environ. Sci. 2009, 2, 35–54. [Google Scholar] [CrossRef]
  5. Navarro, R.M.; Peña, M.A.; Fierro, J.L.G. Hydrogen Production Reactions from Carbon Feedstocks: Fossil Fuels and Biomass. Chem. Rev. 2007, 107, 3952–3991. [Google Scholar] [CrossRef]
  6. Staffell, I.; Scamman, D.; Velazquez Abad, A.; Balcombe, P.; Dodds, P.E.; Ekins, P.; Shah, N.; Ward, K.R. The Role of Hydrogen and Fuel Cells in the Global Energy System. Energy Environ. Sci. 2019, 12, 463–491. [Google Scholar] [CrossRef]
  7. Iwashina, K.; Kudo, A. Rh-Doped SrTiO3 Photocatalyst Electrode Showing Cathodic Photocurrent for Water Splitting under Visible-Light Irradiation. J. Am. Chem. Soc. 2011, 133, 13272–13275. [Google Scholar] [CrossRef]
  8. Rioche, C.; Kulkarni, S.; Meunier, F.C.; Breen, J.P.; Burch, R. Steam Reforming of Model Compounds and Fast Pyrolysis Bio-Oil on Supported Noble Metal Catalysts. Appl. Catal. B Environ. 2005, 61, 130–139. [Google Scholar] [CrossRef]
  9. Sangle, A.L.; Singh, S.; Jian, J.; Bajpe, S.R.; Wang, H.; Khare, N.; MacManus-Driscoll, J.L. Very High Surface Area Mesoporous Thin Films of SrTiO3 Grown by Pulsed Laser Deposition and Application to Efficient Photoelectrochemical Water Splitting. Nano Lett. 2016, 16, 7338–7345. [Google Scholar] [CrossRef]
  10. Deluga, G.A.; Salge, J.R.; Schmidt, L.D.; Verykios, X.E. Renewable Hydrogen from Ethanol by Autothermal Reforming. Science 2004, 303, 993–997. [Google Scholar] [CrossRef]
  11. Jia, Y.; Zhao, D.; Li, M.; Han, H.; Li, C. La and Cr Co-Doped SrTiO3 as an H2 Evolution Photocatalyst for Construction of a Z-Scheme Overall Water Splitting System. Chin. J. Catal. 2018, 39, 421–430. [Google Scholar] [CrossRef]
  12. Cerón, M.R.; Izquierdo, M.; Alegret, N.; Valdez, J.A.; Rodríguez-Fortea, A.; Olmstead, M.M.; Balch, A.L.; Poblet, J.M.; Echegoyen, L. Reactivity Differences of Sc 3 N@C 2n (2n = 68 and 80). Synthesis of the First Methanofullerene Derivatives of Sc 3 N@D 5h -C 80. Chem. Commun. 2016, 52, 64–67. [Google Scholar] [CrossRef]
  13. Wei, Y.; Wan, J.; Wang, J.; Zhang, X.; Yu, R.; Yang, N.; Wang, D. Hollow Multishelled Structured SrTiO3 with La/Rh Co-Doping for Enhanced Photocatalytic Water Splitting under Visible Light. Small 2021, 17, 2005345. [Google Scholar] [CrossRef] [PubMed]
  14. Yamaguchi, Y.; Terashima, C.; Sakai, H.; Fujishima, A.; Kudo, A.; Nakata, K. Photocatalytic Degradation of Gaseous Acetaldehyde over Rh-Doped SrTiO3 under Visible Light Irradiation. Chem. Lett. 2016, 45, 42–44. [Google Scholar] [CrossRef]
  15. Bi, Y.; Ehsan, M.F.; Huang, Y.; Jin, J.; He, T. Synthesis of Cr-Doped SrTiO3 Photocatalyst and Its Application in Visible-Light-Driven Transformation of CO2 into CH4. J. CO2 Util. 2015, 12, 43–48. [Google Scholar] [CrossRef]
  16. Liu, P.; Nisar, J.; Pathak, B.; Ahuja, R. Hybrid Density Functional Study on SrTiO3 for Visible Light Photocatalysis. Int. J. Hydrog. Energy 2012, 37, 11611–11617. [Google Scholar] [CrossRef]
  17. Zou, F.; Jiang, Z.; Qin, X.; Zhao, Y.; Jiang, L.; Zhi, J.; Xiao, T.; Edwards, P.P. Template-Free Synthesis of Mesoporous N-Doped SrTiO3 Perovskite with High Visible-Light-Driven Photocatalytic Activity. Chem. Commun. 2012, 48, 8514–8516. [Google Scholar] [CrossRef]
  18. Chen, W.; Liu, H.; Li, X.; Liu, S.; Gao, L.; Mao, L.; Fan, Z.; Shangguan, W.; Fang, W.; Liu, Y. Polymerizable Complex Synthesis of SrTiO3:(Cr/Ta) Photocatalysts to Improve Photocatalytic Water Splitting Activity under Visible Light. Appl. Catal. B Environ. 2016, 192, 145–151. [Google Scholar] [CrossRef]
  19. Zhou, X.; Liu, N.; Yokosawa, T.; Osvet, A.; Miehlich, M.E.; Meyer, K.; Spiecker, E.; Schmuki, P. Intrinsically Activated SrTiO3: Photocatalytic H 2 Evolution from Neutral Aqueous Methanol Solution in the Absence of Any Noble Metal Cocatalyst. ACS Appl. Mater. Interfaces 2018, 10, 29532–29542. [Google Scholar] [CrossRef]
  20. Zhou, D.; Wang, G.; Feng, Y.; Chen, W.; Chen, J.; Yu, Z.; Zhang, Y.; Wang, J.; Tang, L. CuS Co-Catalyst Modified Hydrogenated SrTiO3 Nanoparticles as an Efficient Photocatalyst for H 2 Evolution. Dalton Trans. 2021, 50, 7768–7775. [Google Scholar] [CrossRef] [PubMed]
  21. Quan, H.; Qian, K.; Xuan, Y.; Lou, L.-L.; Yu, K.; Liu, S. Superior Performance in Visible-Light-Driven Hydrogen Evolution Reaction of Three-Dimensionally Ordered Macroporous SrTiO3 Decorated with ZnxCd1−xS. Front. Chem. Sci. Eng. 2021, 15, 1561–1571. [Google Scholar] [CrossRef]
  22. Yu, K.; Zhang, C.; Chang, Y.; Feng, Y.; Yang, Z.; Yang, T.; Lou, L.-L.; Liu, S. Novel Three-Dimensionally Ordered Macroporous SrTiO3 Photocatalysts with Remarkably Enhanced Hydrogen Production Performance. Appl. Catal. B Environ. 2017, 200, 514–520. [Google Scholar] [CrossRef]
  23. Chang, Y.; Yu, K.; Zhang, C.; Yang, Z.; Feng, Y.; Hao, H.; Jiang, Y.; Lou, L.-L.; Zhou, W.; Liu, S. Ternary CdS/Au/3DOM-SrTiO3 Composites with Synergistic Enhancement for Hydrogen Production from Visible-Light Photocatalytic Water Splitting. Appl. Catal. B Environ. 2017, 215, 74–84. [Google Scholar] [CrossRef]
  24. Ganapathy, M.; Hsu, Y.; Thomas, J.; Chang, C.T.; Alagan, V. Co-Catalyst Free SrTiO3 Nano-Cube for Efficient Photocatalytic Hydrogen Production. J. Mater. Sci. 2021, 56, 18976–18988. [Google Scholar] [CrossRef]
  25. Luo, Y.; Deng, B.; Pu, Y.; Liu, A.; Wang, J.; Ma, K.; Gao, F.; Gao, B.; Zou, W.; Dong, L. Interfacial Coupling Effects in G-C3N4/SrTiO3 Nanocomposites with Enhanced H2 Evolution under Visible Light Irradiation. Appl. Catal. B Environ. 2019, 247, 1–9. [Google Scholar] [CrossRef]
  26. Lu, L.; Lv, M.; Liu, G.; Xu, X. Photocatalytic Hydrogen Production over Solid Solutions between BiFeO3 and SrTiO3. Appl. Surf. Sci. 2017, 391, 535–541. [Google Scholar] [CrossRef]
  27. Vijay, A.; Bairagi, K.; Vaidya, S. Relating the Structure, Properties, and Activities of Nanostructured SrTiO3 and SrO–(SrTiO3)n (n = 1 and 2) for Photocatalytic Hydrogen Evolution. Mater. Adv. 2022, 3, 5055–5063. [Google Scholar] [CrossRef]
  28. Choudhary, S.; Solanki, A.; Upadhyay, S.; Singh, N.; Satsangi, V.R.; Shrivastav, R.; Dass, S. Nanostructured CuO/SrTiO3 Bilayered Thin Films for Photoelectrochemical Water Splitting. J. Solid State Electrochem. 2013, 17, 2531–2538. [Google Scholar] [CrossRef]
  29. Ahmadi, M.; Seyed Dorraji, M.S.; Hajimiri, I.; Rasoulifard, M.H. The Main Role of CuO Loading against Electron-Hole Recombination of SrTiO3: Improvement and Investigation of Photocatalytic Activity, Modeling and Optimization by Response Surface Methodology. J. Photochem. Photobiol. A. 2021, 404, 112886. [Google Scholar] [CrossRef]
  30. Banakhojasteh, S.; Beckert, S.; Gläser, R. Modification of SrTiO3 as a photocatalyst for hydrogen evolution from aqueous methanol solution. J. Photochem. Photobiol. A 2018, 366, 48–54. [Google Scholar] [CrossRef]
  31. Puangpetch, T.; Sreethawong, T.; Yoshikawa, S.; Chavadej, S. Hydrogen Production from Photocatalytic Water Splitting over Mesoporous-Assembled SrTiO3 Nanocrystal-Based Photocatalysts. J. Mol. Catal. Chem. 2009, 312, 97–106. [Google Scholar] [CrossRef]
  32. Yu, J.; Hai, Y.; Jaroniec, M. Photocatalytic Hydrogen Production over CuO-Modified Titania. J. Colloid Interface Sci. 2011, 357, 223–228. [Google Scholar] [CrossRef]
  33. Song, S.; Xu, L.; He, Z.; Ying, H.; Chen, J.; Xiao, X.; Yan, B. Photocatalytic Degradation of C.I. Direct Red 23 in Aqueous Solutions under UV Irradiation Using SrTiO3/CeO2 Composite as the Catalyst. J. Hazard. Mater. 2008, 152, 1301–1308. [Google Scholar] [CrossRef]
  34. Bandara, J.; Udawatta, C.P.K.; Rajapakse, C.S.K. Highly Stable CuO Incorporated TiO2 Catalyst for Photocatalytic Hydrogen Production from H2O. Photochem. Photobiol. Sci. 2005, 4, 857. [Google Scholar] [CrossRef] [PubMed]
  35. Choi, H.; Kang, M. Hydrogen Production from Methanol/Water Decomposition in a Liquid Photosystem Using the Anatase Structure of Cu Loaded TiO2. Int. J. Hydrog. Energy 2007, 32, 3841–3848. [Google Scholar] [CrossRef]
  36. Qi, W.; Wang, C.; Yu, J.; Adimi, S.; Thomas, T.; Guo, H.; Liu, S.; Yang, M. MOF-Derived Porous Ternary Nickel Iron Nitride Nanocube as a Functional Catalyst toward Water Splitting Hydrogen Evolution for Solar to Chemical Energy Conversion. ACS Appl. Energy Mater. 2022, 5, 6155–6162. [Google Scholar] [CrossRef]
  37. Cheng, Z.; Qi, W.; Pang, C.H.; Thomas, T.; Wu, T.; Liu, S.; Yang, M. Recent Advances in Transition Metal Nitride-Based Materials for Photocatalytic Applications. Adv. Funct. Mater. 2021, 31, 2100553. [Google Scholar] [CrossRef]
  38. Korzhak, A.V.; Ermokhina, N.I.; Stroyuk, A.L.; Bukhtiyarov, V.K.; Raevskaya, A.E.; Litvin, V.I.; Kuchmiy, S.Y.; Ilyin, V.G.; Manorik, P.A. Photocatalytic Hydrogen Evolution over Mesoporous TiO2/Metal Nanocomposites. J. Photochem. Photobiol. Chem. 2008, 198, 126–134. [Google Scholar] [CrossRef]
  39. Kudo, A.; Miseki, Y. Heterogeneous Photocatalyst Materials for Water Splitting. Chem. Soc. Rev. 2009, 38, 253–278. [Google Scholar] [CrossRef]
  40. Qin, Y.; Wang, G.; Wang, Y. Study on the Photocatalytic Property of La-Doped CoO/SrTiO3 for Water Decomposition to Hydrogen. Catal. Commun. 2007, 8, 926–930. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of pure SrTiO3 (a) and 1.5 wt.% CuO-loaded SrTiO3 (b).
Figure 1. XRD patterns of pure SrTiO3 (a) and 1.5 wt.% CuO-loaded SrTiO3 (b).
Inorganics 10 00130 g001
Figure 2. TEM images of SrTiO3 (a), 1.5% CuO-SrTiO3 (b) and EDX spectra of 1.5% CuO-SrTiO3 (c).
Figure 2. TEM images of SrTiO3 (a), 1.5% CuO-SrTiO3 (b) and EDX spectra of 1.5% CuO-SrTiO3 (c).
Inorganics 10 00130 g002
Figure 3. XPS spectra of the samples: (a) XPS survey spectrum of 1.5% CuO-SrTiO3 sample; (b) high-resolution Cu 2p spectrum; (c) high-resolution Ti 2p spectrum and (d) high-resolution Sr d spectrum.
Figure 3. XPS spectra of the samples: (a) XPS survey spectrum of 1.5% CuO-SrTiO3 sample; (b) high-resolution Cu 2p spectrum; (c) high-resolution Ti 2p spectrum and (d) high-resolution Sr d spectrum.
Inorganics 10 00130 g003
Figure 4. The effects of the loading amount of CuO (a) and Pt (b) on the photocatalytic hydrogen evolution activity of SrTiO3 (catalyst: 60 mg; methanol content: 60 mL, 60 vol.%; reaction temperature: 25 °C).
Figure 4. The effects of the loading amount of CuO (a) and Pt (b) on the photocatalytic hydrogen evolution activity of SrTiO3 (catalyst: 60 mg; methanol content: 60 mL, 60 vol.%; reaction temperature: 25 °C).
Inorganics 10 00130 g004
Figure 5. The effects of type and concentration of electron donors on the photocatalytic hydrogen evolution activity over 1.5% CuO-SrTiO3: (a) methanol, (b) glycerol, (c) TEOA and (d) EDTA. The amount of catalyst: 60 mg; volume of reaction solution: 60 mL; reaction temperature: 25 °C; irradiation time: 2 h.
Figure 5. The effects of type and concentration of electron donors on the photocatalytic hydrogen evolution activity over 1.5% CuO-SrTiO3: (a) methanol, (b) glycerol, (c) TEOA and (d) EDTA. The amount of catalyst: 60 mg; volume of reaction solution: 60 mL; reaction temperature: 25 °C; irradiation time: 2 h.
Inorganics 10 00130 g005
Figure 6. The effect of reaction temperature on photocatalytic hydrogen evolution activity over 1.5% CuO-SrTiO3 photocatalyst. Catalyst: 60 mg; amount of methanol: 60 mL, 60 vol.%; irradiation time: 2 h.
Figure 6. The effect of reaction temperature on photocatalytic hydrogen evolution activity over 1.5% CuO-SrTiO3 photocatalyst. Catalyst: 60 mg; amount of methanol: 60 mL, 60 vol.%; irradiation time: 2 h.
Inorganics 10 00130 g006
Figure 7. Effect of the addition amount of the 1.5% CuO-SrTiO3 photocatalyst on the photocatalytic activity for hydrogen evolution.The amount of methanol: 60 mL, 60 vol.%; reaction temperature: 45 °C; irradiation time: 2 h.
Figure 7. Effect of the addition amount of the 1.5% CuO-SrTiO3 photocatalyst on the photocatalytic activity for hydrogen evolution.The amount of methanol: 60 mL, 60 vol.%; reaction temperature: 45 °C; irradiation time: 2 h.
Inorganics 10 00130 g007
Figure 8. Time course of hydrogen evolution over 1.5% CuO-SrTiO3 (a) and pure SrTiO3 (b). The amount of photocatalyst: 120 mg; the concentration of methanol: 60 mL, 60 vol.%; reaction temperature: 45 °C.
Figure 8. Time course of hydrogen evolution over 1.5% CuO-SrTiO3 (a) and pure SrTiO3 (b). The amount of photocatalyst: 120 mg; the concentration of methanol: 60 mL, 60 vol.%; reaction temperature: 45 °C.
Inorganics 10 00130 g008
Figure 9. Diffuse reflectance spectra of pure SrTiO3 (a) and 1.5 wt.% CuO-SrTiO3 (b).
Figure 9. Diffuse reflectance spectra of pure SrTiO3 (a) and 1.5 wt.% CuO-SrTiO3 (b).
Inorganics 10 00130 g009
Figure 10. Photoluminescence (PL) spectra of pure SrTiO3 and the 1.5% CuO-SrTiO3.
Figure 10. Photoluminescence (PL) spectra of pure SrTiO3 and the 1.5% CuO-SrTiO3.
Inorganics 10 00130 g010
Figure 11. Scheme of photocatalytic hydrogen evolved reaction over the CuO-SrTiO3 photocatalyst under UV light irradiation.
Figure 11. Scheme of photocatalytic hydrogen evolved reaction over the CuO-SrTiO3 photocatalyst under UV light irradiation.
Inorganics 10 00130 g011
Figure 12. Schematic diagram of the energy band positions of SrTiO3, CuO and the direction of electron transfer.
Figure 12. Schematic diagram of the energy band positions of SrTiO3, CuO and the direction of electron transfer.
Inorganics 10 00130 g012
Table 1. The photocatalytic hydrogen evolution over various photocatalysts (photocatalyst: 60 mg; the content of methanol: 60 mL, 60 vol.%; temperature: 25 °C; irradiation time: 2 h).
Table 1. The photocatalytic hydrogen evolution over various photocatalysts (photocatalyst: 60 mg; the content of methanol: 60 mL, 60 vol.%; temperature: 25 °C; irradiation time: 2 h).
PhotocatalystsAmount of Hydrogen Evolution (µmol)
Pure SrTiO31.45
0.5 wt.% CuO-loaded SrTiO3165
1.0 wt.% CuO-loaded SrTiO3211
1.5 wt.% CuO-loaded SrTiO3258
2.0 wt.% CuO-loaded SrTiO3222
2.5 wt.% CuO-loaded SrTiO3200
0.5 wt.% Pt-loaded SrTiO3297
1.0 wt.% Pt-loaded SrTiO3318
1.5 wt.% Pt-loaded SrTiO3361
2.0 wt.% Pt-loaded SrTiO3403
2.5 wt.% Pt-loaded SrTiO3341
3.0 wt.% Pt-loaded SrTiO3309
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mai, X.T.; Bui, D.N.; Pham, V.K.; Nguyen, T.H.L.; Nguyen, T.T.L.; Chau, H.D.; Tran, T.K.N. Effect of CuO Loading on the Photocatalytic Activity of SrTiO3 for Hydrogen Evolution. Inorganics 2022, 10, 130. https://doi.org/10.3390/inorganics10090130

AMA Style

Mai XT, Bui DN, Pham VK, Nguyen THL, Nguyen TTL, Chau HD, Tran TKN. Effect of CuO Loading on the Photocatalytic Activity of SrTiO3 for Hydrogen Evolution. Inorganics. 2022; 10(9):130. https://doi.org/10.3390/inorganics10090130

Chicago/Turabian Style

Mai, Xuan Truong, Duc Nguyen Bui, Van Khang Pham, Thi Hien Lan Nguyen, Thi To Loan Nguyen, Hung Dung Chau, and Thi Kim Ngan Tran. 2022. "Effect of CuO Loading on the Photocatalytic Activity of SrTiO3 for Hydrogen Evolution" Inorganics 10, no. 9: 130. https://doi.org/10.3390/inorganics10090130

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