Skip to main content
ACS Omega logoLink to ACS Omega
. 2020 Dec 8;5(50):32715–32723. doi: 10.1021/acsomega.0c05106

p–n Heterojunction Photocatalyst Mn0.5Cd0.5S/CuCo2S4 for Highly Efficient Visible Light-Driven H2 Production

Mingyue Zhang , Ningjie Fang , Xincheng Song , Yinghao Chu †,‡,*, Song Shu , Yongjun Liu †,
PMCID: PMC7758946  PMID: 33376909

Abstract

graphic file with name ao0c05106_0012.jpg

It is highly important to develop efficient and cheap photocatalysts for hydrogen production. Herein, a series of p–n heterojunction Mn0.5Cd0.5S/CuCo2S4 has been successfully synthesized for the first time by the hydrothermal impregnation method. Mn0.5Cd0.5S/CuCo2S4 loading with 12 wt % CuCo2S4 shows the highest H2 evolution rate of 15.74 mmol h–1 g–1 under visible light (λ ≥ 420 nm) irradiation, which is about 3.15 and 15.28 times higher than that of bare Mn0.5Cd0.5S (4.99 mmol h–1 g–1) and CuCo2S4 (1.03 mmol h–1 g–1), respectively. In addition, it shows a relatively good stability during the five recycle tests, with about 20% loss of reaction rate compared to that of the first cycle. The superior photocatalytic performance is attributed to the effective separation and transfer of photogenerated charge carriers because of the formation of the p–n junction. The samples are systematically characterized by X-ray diffraction, ultraviolet–visible (UV–vis), diffuse reflectance spectroscopy, scanning electron microscopy, transmission electron microscopy (TEM), high-resolution TEM, X-ray photoelectron spectroscopy, photoluminescence, EIS, and so on. UV–vis and EIS show that CuCo2S4 can effectively improve the visible light response of Mn0.5Cd0.5S/CuCo2S4 and promote the electron transfer from CuCo2S4 to the conduction band of Mn0.5Cd0.5S, so as to improve the photocatalytic efficiency. This study reveals that the p–n heterojunction Mn0.5Cd0.5S/CuCo2S4 is a promising photocatalyst to explore the photocatalysts without noble metals.

1. Introduction

Because of the consumption of fossil fuels and serious environmental pollution, ecological damage and energy shortage seriously threaten the sustainable development of society.1,2 Much attention has been drawn to develop clean and green energy to address global challenges.3 Hydrogen energy is considered as an ideal energy resource because of its high conversion efficiency, environmental friendliness, and recyclability.4 Photocatalytic water splitting has been considered to be a promising technology, which can convert solar energy into chemical energy. It can also promote the energy revolution, achieve sustainable development, and has a great role in promoting the construction of a low-carbon society. So far, a large number of semiconductor materials, including oxides, sulfides, and nitrogen oxides, have been used for photocatalytic hydrogen production from water reduction.5,6 Among these, metal sulfides, such as MnxCd1–xS, have shown promising hydrogen production performances because of wide solar response and suitable band edge positions.7 However, the H2 evolution rate of MnxCd1–xS is still unsatisfied because of the rapid recombination of photogenerated electron–hole pairs.8,9 To solve these problems, several strategies, including metal doping,10 cocatalyst modification,7,11 morphological tailoring, and construction of a heterojunction,1214 have been developed to further improve the photoactivity of MnxCd1–xS. Among these approaches, cocatalyst modification and construction of a heterojunction can be considered as ideal approaches because of the accelerated charge separation. Besides, the cocatalyst can provide low activation potentials and abundant active sites.15 However, up to now, the most effective cocatalyst is still based on noble metals, such as Pt, Ag, Au, and so on. Kumar et al.16 have designed a Ag/Fe3O4-bridged SrTiO3/g-C3N4 composite to improve the H2 evolution rate. Liu et al.6 have prepared a new Mn0.5Cd0.5S/WO3/Au ternary heterostructure to boost the photocatalytic activity. However, it is difficult to achieve industrial application because of high cost.17,18 Therefore, developing an efficient, stable, and earth-abundant cocatalyst is of great significance for MnxCd1–xS solid solution systems.

Recently, transition metal sulfides, such as MoSx, CoSx, NiS, CuS, and so forth, have been widely used in photocatalytic hydrogen evolution because of their high efficiency, stability, and low cost. For example, Liu et al.19 have synthesized a novel Mn0.5Cd0.5S/RGO-MoS2 ternary heterojunction, which exhibits a higher H2 production efficiency compared to that of bare Mn0.5Cd0.5S. Wang et al.8 found that MoS2 modification via a one-pot solvothermal process could significantly improve the photocatalytic performance of Mn0.2Cd0.8S/MnS. Zhai and co-workers7 have successfully synthesized a series of transition metal sulfide XS (X = Mo, Cu, and Pd)-supported Mn0.5Cd0.5S composites. The results showed that XS can improve the H2 production activity of Mn0.5Cd0.5S.

As a bimetallic sulfide, CuCo2S4 has high conductivity because of the low electronegativity of sulfur. However, there are few reports on the application of CuCo2S4 toward photocatalytic H2 production. In this study, Mn0.5Cd0.5S and CuCo2S4 were successfully synthesized by a simple one-step hydrothermal method, separately. Moreover, Mn0.5Cd0.5S/CuCo2S4 nanocomposites were prepared by the hydrothermal impregnation method for the first time. Mn0.5Cd0.5S and CuCo2S4 can form a p–n heterojunction, which shows excellent photocatalytic activity for hydrogen evolution (15.74 mmol h–1 g–1) under visible light. This work may provide a suitable suggestion for the synthesis of p–n heterojunction composite materials without noble metals to improve H2 production activity and broaden the synthetic application of hydrogen-producing materials.

2. Results and Discussion

2.1. X-ray Diffraction Analysis

The phase structure of as-prepared MnxCd1–xS, CCS and CMS/yCCS nanocomposites was investigated by X-ray diffraction (XRD). In Figure 1a, the diffraction peaks of the MnxCd1–xS products are shifted toward higher diffraction angles from CdS (PDF#89-0440) to MnS (JCPDS#72-1534) with the increase in the x value from 0.1 to 0.5. The continuous shifts of XRD patterns show that the samples are not the mixture of CdS and MnS but MnxCd1–xS (0.1 ≤ x ≤ 0.5) solid solution. This implies that Cd2+ with a larger ionic radius (0.97 Å) than Mn2+ (0.46 Å) incorporates into the lattice of the MnS crystal, which increases the fringe lattice distance and leads to the formation of a solid solution.20 With the increase in the x value (x > 0.5) in the MnxCd1–xS, some diffraction peaks of MnS appear; the characteristic peaks at about 34.2, 49.1, and 61.2° are ascribed to the (200), (220), and (222) planes of MnS (PDF#72-1534), respectively, indicating that the MnS exists in the samples. The XRD pattern of the as-prepared CCS sample is shown in Figure 1c. The peaks at 26.6, 31.3, 37.9, 46.9, 49.9, and 54.8° are ascribed to the cubic CCS (PDF#42-1450). There are no impurity peaks, indicating the synthesis of the pure-phase CCS. In Figure 1c, pure CMS and all CMS/yCCS composites exhibit the similar diffraction peaks, meaning the good dispersion of CCS nanoparticles in the surface of CMS, and the deposition of CCS does not affect the crystalline structure of CMS.

Figure 1.

Figure 1

(a) XRD patterns of the as-prepared MnxCd1–xS (0.1 ≤ x ≤ 0.9); (b) XRD patterns at the lower-angle region (20° < 2θ < 35°) of figure (a); and (c) XRD patterns of CMS, CCS, and CMS/yCCS.

In order to further clarify the composition of CMS/yCCS, the inductively coupled plasma-optical emission spectroscopy test was carried out. It can be seen from Table S1 that the actual molar ratio of Mn/Cd/S in CMS/12CCS is 0.46:0.54:1, which is very close to 0.5:0.5:1. The actual loading amount of CCS is 11.9%, which is close to the theoretical 12%. These results indicate that the CMS/yCCS composite has been successfully synthesized.

2.2. Scanning Electron Microscopy, Energy-Dispersive Spectrometer, and Transmission Electron Microscopy Analysis

The surface morphologies of as-prepared CMS and CMS/12CCS composites were investigated by scanning electron microscopy (SEM) and the results are shown in Figure 2. It can be observed that the morphology of the CMS/12CCS nanocomposite consists of some irregular particles with an average size of 50–60 nm. It is similar to that of pure CMS, indicating that CCS loading has no effect on the morphology of CMS. The bright field micrograph in Figure 3 shows that the average particle size of the CMS/12CCS nanocomposite is 50–60 nm, which is consistent with the SEM results. The high-resolution transmission electron microscopy (HRTEM) image in Figure 3c shows three different kinds of lattice fringes. The lattice fringe with a spacing of about 0.328 nm can be assigned to the (002) crystal plane of CMS. The lattice fringes with a spacing of 0.533 and 0.285 nm are assigned to the (111) and (113) crystal plane of CCS, respectively. Therefore, CCS is in tight contact with the host material CMS, which can promote the transfer rate of photoinduced carriers between CMS and CCS. Moreover, Mn, Cd, Cu, Co, and S elements are all detected using an energy-dispersive spectrometer, demonstrating again that CCS is well dispersed in the surface of CMS. In Figure 4a–f, the EDS element mapping of CMS/12CCS displays the homogeneous distribution of Mn, Cd, S, Cu, and Co elements in the sample. All of these results confirm that there is an intimate interfacial contact between CMS and CCS. Such an intimate heterojunction can provide more active sites and promote the effective separation of photogenerated carriers, thus improving the activity of H2 evolution.

Figure 2.

Figure 2

(a) SEM images of CMS; (b) SEM images of CMS/12CCS; and (c) EDS pattern of CMS/12CCS.

Figure 3.

Figure 3

(a) TEM images of CMS. (b,c) TEM and HRTEM images of CMS/12CCS.

Figure 4.

Figure 4

(a) Original map of the CMS/12CCS mapping map. (b–f) EDS mapping images of Mn, Cd, S, Cu, and Co.

2.3. X-ray Photoelectron Spectroscopy Analysis

X-ray photoelectron spectroscopy (XPS) was further used to investigate the surface chemical status of CMS/12CCS. In Figure 5a, Mn, Cd, Cu, Co, and S elements can be detected obviously, which is in good agreement with the EDS results. Figure 5b–f shows the homologous high-resolution XPS spectra of Mn, Cd, S, Cu, and Co, respectively. Compared to pure CMS, slight shifts toward higher binding energies can be observed from the high-resolution spectra of CMS/12CCS in Figure 4b–d, suggesting that the binding energies of the core-level electrons of these metal and sulfide ions are affected because of the possible chemical bonding actions among the composite components.21,22 Compared to the pristine CMS, the binding energies of Cd 3d and Mn 2p in the CMS/12CCS slightly shift toward a high value, which suggests the presence of strong chemical and electronic coupling between CMS and CCS. In the Mn 2p XPS spectrum (Figure 5b), the peaks at 652.3 and 640.7 eV correspond to 2p1/2 and 2p3/2 of Mn2+, respectively.23 In Figure 5c, the Cd 3d spectrum displays two binding energy peaks at 411.6 and 404.8 eV, which are corresponding to Cd 3d5/2 and Cd 3d3/2 of Cd2+, respectively.24 The binding energy peaks of S 2p in Figure 5d are located at 162.2 and 161.1 eV, which can be attributed to 2p1/2 and 2p3/2 of S2– in CMS/12CCS nanocomposites, respectively. The Cu 2p peaks can be divided into two peaks, including Cu 2p1/2 (952.2 eV) and Cu 2p3/2 (932.3 eV) (Figure 5e),25 which indicates that the oxidation state of copper was +1.26 The Co 2p peak displayed in Figure 5f can be divided into four peaks. The peaks at 793.5 and 778.5 eV can be attributed to Co 2p1/2 and Co 2p3/2, respectively.27 The peaks at 782.1 and 797.2 eV are the satellite peaks, indicating that the Co element exists in the form of Co3+ and Co2+. All these results indicate the presence of chemical bonds between CCS and CMS rather than a simple physical mixture.14

Figure 5.

Figure 5

XPS spectra of the CMS and CMS/12CCS samples: (a) survey spectra; (b) Mn 2p; (c) Cd 3d; (d) S 2p; (e) Cu 2p; and (f) Co 2p.

2.4. BET Surface Areas

The Brunauer–Emmett–Teller (BET) specific surface areas (SBET) of pure CMS, CCS, and the as-prepared CMS/CCS composites were measured by nitrogen adsorption and are summarized in Table S2. The BET specific surface areas of the pristine CMS and CCS are 25.95 and 35.62 m2 g–1, respectively. Although CMS/12CCS has the smallest area (18.76 m2 g–1), it shows the highest activity. Consistent with Bao’s work,28SBET is not a key factor to restrict the H2 production rate of CMS/CCS. The different H2 evolution rate of CMS/CCS should be attributed to other factors, such as their crystal structure and band structure, because the separation of photogenerated electron–hole pairs and the proceeding of H2 evolution rate will be determined by these structures.

2.5. Photocatalytic H2 Evolution Activity and Stability

The photocatalytic H2 evolution activities of all samples were evaluated under visible light irradiation (λ ≥ 420 nm) using Na2S (0.35 M) and Na2SO3 (0.25 M) as sacrificial agents. Figure 6a shows the photocatalytic H2 evolution efficiency of MnxCd1–xS with different Mn/Cd ratios. The results show that the H2 evolution efficiency has enhanced with the increase in Mn content (x < 0.5). However, when x is more than 0.5, the H2 evolution efficiency of MnxCd1–xS decreases with a further increase in the x value. When x = 0.5, MnxCd1–xS has the highest H2 evolution activity (4.99 mmol h–1 g–1). Figure 6c,d shows the photocatalytic H2 evolution rate of CMS/yCCS composites with different contents of CCS. Almost no H2 is detected for bare CCS, suggesting its poor photocatalytic activity for H2 evolution. CMS exhibits a low H2 evolution rate of 4.99 mmol h–1 g–1, probably because of the high recombination of photogenerated electron–hole pairs. However, after loading CCS as a cocatalyst, H2 evolution performance of CMS/yCCS improved obviously, indicating that CCS is an effective cocatalyst for CMS. The H2 evolution performance enhances with the increased amount of CCS from 3 to 12 wt %. The optimal CCS loading amount is 12 wt %, with the highest H2 evolution rate of 15.74 mmol h–1 g–1, which is about 3.15 times and 2.89 times higher than that of CMS and Mn0.5Cd0.5S/3 wt % Pt (marked as CMS/3Pt), respectively. The photocatalytic H2 evolution activities of all samples were also evaluated under the same conditions, except that the 420 nm cutoff filter was replaced with a 510 nm cutoff filter. It can be seen from Figure S3 that when the wavelength is more than 510 nm, the H2 production activity of CMS/12CCS is 1.26 mmol h–1 g–1, which is 2.17 times than that of a pure CMS (0.58 mmol h–1 g–1). The AQE of CMS/12CCS was calculated to be 41.48% at 420 nm. The hydrogen performance and AQE of MnxCd1–xS-based catalysts are listed in Table S3. It can be seen from Table S3 that CMS/12CCS has a high hydrogen evolution rate and the highest AQE. However, with the further increase in CCS, the H2 evolution performance drops. The reason can be ascribed to the shielding effect that the excessive black CCS will partially block the light absorption of CMS.9 Besides, the active sites for H2 evolution can be covered by excessive CCS, thus blocking the effective contact between the active sites and reactants. Therefore, the overmuch amount of CCS can act as the recombination centers of photon-generated carriers and reduce the light absorption of CMS. What is more, to confirm the photostability of CMS/12CCS, five cycle tests have been carried out under the same conditions for 20 h. As shown in Figure 6f, after five cycle tests, the H2 evolution rate is 12.62 mmol h–1 g–1, which is 19.82% lower than the 15.74 mmol h–1 g–1 in the first cycle. Although the photocatalytic H2 evolution activity shows a little decrease after five-round cycles, it still maintains good stability. For the CMS/12CCS, XPS test was performed after the reaction. It can be seen from Figure S2 that the XPS spectra shift to higher binding energy after the reaction, which can be attributed to the electron–hole pair transfer after the photocatalytic reaction; however, no obvious changes are found in XRD and SEM before and after reaction (Figure S1), indicating that CMS/12CCS has good stability. We have synthesized five batches of CMS/CCS by the same method, and the activity test was carried out under the same conditions. As shown in Figure S4, the average H2 evolution rate of CMS/12CCS is basically stable at 15.7 mmol h–1 g–1, with certain repeatability. The results indicate that the CMS/12CCS photocatalyst is stable for photocatalytic H2 evolution.

Figure 6.

Figure 6

(a) H2 production activity of MnxCd1–xS products; (b) photocatalytic hydrogen production rate of MnxCd1–xS products; (c) H2 production activity of CMS/yCCS-based products; (d) photocatalytic hydrogen production rate of CMS/yCCS-based products; (e) comparison of H2 evolution of CMS, CMS/12CCS, and CMS/3Pt; and (f) cycling runs of the CMS/12CCS.

2.6. Optical Absorption

Ultraviolet–visible (UV–vis) diffuse reflectance spectra of samples are shown in Figure 7. As can be seen, the pure CMS sample displays an absorption edge at about 525 nm, corresponding to a band gap of 2.46 eV (Figure S5a). Pure CCS has a strong light response in the whole visible light region because of its narrow band gap of 1.75 eV (Figure S5b). For the CMS/yCCS composite, with the increase in CCS content from 3 to 15 wt %, the absorption edge of the composite displays an obvious red shift and enhanced absorption in the visible light region, which is attributed to the contribution of CCS on the surface of CMS. It is believed that the enhanced light absorption of CMS/yCCS composites is favorable to generate more available photogenerated electrons for participating in the photocatalytic reaction toward H2 evolution.

Figure 7.

Figure 7

UV–vis DRS spectra of CMS, CCS, and CMS/yCCS composites.

2.7. Photoluminescence Spectra and Photoelectrochemical Measurements

In order to study the processes of charge transfer, separation, and recombination in the CMS/CCS, the photoluminescence (PL) emission spectrum was measured under an excitation wavelength of 330 nm (Figure 8a).29,30 The lower the PL peak intensity, the lower the recombination rate of photogenerated electron–hole pairs. Obviously, CMS/12CCS has the lowest PL peak intensity compared to other samples, indicating that the formation of a heterojunction between the CMS solid solution and CCS can promote the photogenerated charge transfer/separation at the interface. To further evaluate the interface charge separation efficiency, the transient photocurrent for the pure CMS and CMS/CCS under visible light irradiation is measured (Figure 8b). The bare CMS shows a relatively low photocurrent density because of the rapid recombination of photoinduced carriers. After loading of the CCS cocatalyst, with the increase of the CCS (y ≤ 12), the sample shows an enhanced photocurrent intensity, suggesting that the separation of the photogenerated charge carriers is obviously promoted. When CCS is overloaded (y > 12), the photocurrent intensity of the sample decreases. Nyquist impedance spectroscopy of CMS and CMS/CCS is displayed in Figure 8c. The CMS/12CCS composite exhibits a smaller arc radius than other samples, indicating that the charge in CMS/12CCS has less resistance during the transportation process. These results indicate that the superior photocatalytic performance of CMS/12CCS is attributed to the effective separation and transfer of photogenerated charge carriers.

Figure 8.

Figure 8

(a) PL spectra of CMS, CMS/CCS, and CCS; (b) transient photocurrent responses; and (c) EIS spectra.

2.8. Proposed Mechanism

On the basis of the Results and Discussion mentioned above, a proposed photocatalysis mechanism over CMS/yCCS nanocomposites has been carried out. The detailed band structures, valence band (VB), and conduction band (CB) of CCS and CMS are studied to investigate the charge-transfer process and direction and the photocatalytic mechanism.

According to the equations

2.8.

where EVB and ECB are the VB and CB of semiconductors, respectively. X is the electronegativity of the semiconductor. Ee is the energy of free electrons on the hydrogen scale (4.5 eV).31Eg is the band gap energy of the semiconductor. The X value of CMS is 4.996,32 so the EVB and ECB of CMS are calculated to be 1.73 and −0.73 eV, respectively (Eg of CMS is 2.46 eV). The X value of CCS is 5.32 eV,33 so the EVB and ECB of CCS are 1.69 and −0.06 eV, respectively (Eg of CCS is 1.75 eV). According to Figure S6a, the Mott–Schottky plots of CMS have a positive slope, showing that it is an n-type semiconductor.34 In contrast, the Mott–Schottky plots of CCS have a negative slope, corresponding to a p-type semiconductor.35 The position of the flat-band potential (Efb) can be considered as the position of the Fermi level. The ECB of n-type semiconductors and EVB of p-type semiconductors are believed to be approximately 0.1 eV above and below Efb, respectively.36 Hence, the Fermi levels of CMS and CCS are estimated to be −0.63 and 1.59 eV, respectively. When n-type CMS is in contact with p-type CCS, the large Fermi energy level difference between them will transfer electrons from CMS to CCS and holes from CCS to CMS until the two Fermi levels are balanced.37 Because of the carrier migration at the contact interface, a built-in electric field can be formed when p-type and n-type semiconductors combine to form a heterojunction. Therefore, when p-type CCS is coupled with n-type CMS, a new p–n heterojunction will be formed. According to the obtained results, both CMS and CCS can be excited under visible light and engender photogenerated electron–hole pairs. Because of the built-in electric field in the interface, the photogenerated electrons transfer from the CB of CCS to the CB of CMS, while the photogenerated holes of CMS transfer to the VB of CCS. In this way, the photogenerated electrons and holes of CMS and CCS will be separated effectively. The accumulated electrons in the CB of CMS can directly reduce H+ to H2; the holes remaining on the VB of CCS can be consumed by the sacrificial agent, further improving the separation of electron–hole pairs. Constructing a heterojunction is a crucial method for the efficient separation of photogenerated carriers.38 The p–n heterojunction system formed between CCS and CMS can promote the separation of photogenerated charge carriers (Figure 9).

Figure 9.

Figure 9

Proposed mechanism of the photocatalytic H2 reduction over CMS/12CCS under visible light irradiation.

3. Conclusions

In summary, we have synthesized a series of CMS/yCCS photocatalysts via the hydrothermal impregnation method. CMS/12CCS loading with 12 wt % CCS exhibits the best H2 production efficiency of 15.74 mmol h–1 g–1 under visible light (λ ≥ 420 nm) irradiation, which is about 3.15 times and 15.28 times higher than that of bare CMS (4.99 mmol h–1 g–1) and CCS (1.03 mmol h–1 g–1), respectively. Additionally, CMS/12CCS shows an excellent AQE of 41.48% at 420 nm and good photostability over 20 h. The significantly photocatalytic performance can be attributed to the effective separation and transfer of photogenerated charge carriers because of the formation of the p–n junction. This work reveals that the heterojunction between CMS and CCS can improve the photocatalytic hydrogen evolution and provides a suitable suggestion for the synthesis of composite materials without noble metals to improve H2 production activity.

4. Experimental Section

4.1. Materials

Manganese acetate tetrahydrate (Mn(CH3COO)2·4H2O), cadmium nitrate tetrahydrate (Cd(NO3)2·4H2O), thioacetamide, ethylenediamine (EDA), copper nitrate trihydrate (Cu(NO3)2·3H2O), cobalt nitrate hexahydrate (Co(NO3)2·6H2O), tertiary butanol (TBA), ammonia water (NH3·H2O), sodium sulfide nonahydrate (Na2S·9H2O), and sodium sulfite (Na2SO3) are analytic reagents and used without further purification.

4.2. Synthesis

4.2.1. Synthesis of MnxCd1–xS

MnxCd1–xS solid solutions were synthesized through a hydrothermal method. First, 2x mmol Mn(CH3COO)2·4H2O, 2(1 – x) mmol Cd(NO3)2·4H2O and excessive amounts of thioacetamide were dissolved in 60 mL of EDA under continuous stirring. Then, the mixture was stirred for 30 min and ultrasonicated for another 30 min. Finally, the mixture was transferred into a 100 mL Teflon-lined autoclave and kept at 180 °C for 24 h. After cooling to room temperature naturally, MnxCd1–xS solid solutions were obtained by washing with ethanol and deionized water alternately several times and then dried at 60 °C for 6 h. The obtained Mn0.5Cd0.5S sample was marked as CMS.

4.2.2. Synthesis of CuCo2S4

A total of 1 mmol Cu(NO3)2·3H2O, 2 mmol Co(NO3)2·6H2O, and 6 mmol thiourea (CH4N2S) were dissolved in 60 mL of deionized water. Then, 15 mL of TBA and 2.5 mL of ammonia water were added into the abovementioned mixed solution and stirred for 1 h. Next, the mixture was transferred to a 100 mL Teflon-lined autoclave and kept at 180 °C for 10 h. After cooling naturally, a black precipitate was obtained by filtration and washed with ethanol and deionized water several times. The collected powder was denoted as CCS.

4.2.3. Preparation of the Mn0.5Cd0.5S/CuCo2S4 Composite

The Mn0.5Cd0.5S/CuCo2S4 nanocomposite was prepared by the hydrothermal impregnation method. In a typical process, 0.1 g of Mn0.5Cd0.5S was dispersed in 50 mL of absolute ethanol. After ultrasonic treatment for 1 h, the quantitative CuCo2S4 was added into the abovementioned solution and ultrasonicated for another 1 h. Finally, the mixture was stirred overnight and dried in an oven at 80 °C. The mass ratio of CuCo2S4 to Mn0.5Cd0.5S was 3, 7, 10, 12, and 15 wt %. The total preparation process is shown in Scheme 1. The final products were marked as CMS/yCCS (y is the mass percentage of CCS, y = 3, 7, 10, 12, and 15).

Scheme 1. Preparation Process of CMS/CCS.

Scheme 1

4.3. Characterizations

The crystal structure of the samples was characterized using a Rigaku SmartLab X-ray diffractometer using Cu Kα radiation. The surface morphologies of the samples were evaluated using a scanning electron microscope (FEI Inspect F50), transmission electron microscope (JEOL model JEM-2010F), and high-resolution transmission electron microscope (JEM-2010F). The XPS measurement was carried out to identify the elemental composition and chemical state of the samples. UV–vis diffuse reflectance absorption spectra of the samples (with BaSO4 as a reference) were studied using a Shimadzu UV 2401 spectrophotometer. The PL was analyzed on an F-7100 fluorescence spectrometer. Nitrogen adsorption data obtained on a NOVA Surface Area Analyzer Station A were used to calculate the BET surface area.

The photoelectrochemical tests were obtained using a CHI model 618C electrochemical analyzer and a standard three-electrode system; the Ag/AgCl (saturated KCl) and Pt foil were used as the reference electrode and counter electrode, respectively. A 300 W Xe lamp with a 420 nm cutoff filter was used as the light source. Mott–Schottky measurements were evaluated in a 0.5 mol/L Na2SO4 electrolyte (pH = 6.2) at a frequency of 1 kHz in the dark. The preparation method of the working electrode is as follows: a total of 30 mg of the samples was dispersed in absolute ethanol and grinded for 30 min and the suspension was spin-coated on cleaned FTO glass and dried at 80 °C for 1 h.

4.4. Photocatalytic Test

The photocatalytic H2 production experiment was completed in the photocatalytic activity evaluation system. In a typical process, 20 mg of the sample was well dispersed in a 100 mL mixed solution containing Na2S (0.35 M) and Na2SO3 (0.25 M) as sacrificial agents. The system was evacuated using a vacuum pump to drive away the dissolved gases. A 300 W Xe lamp with a cutoff filter (λ ≥ 420 nm) was used as the visible light source. The temperature of the reaction system was maintained at about 6 °C. Before irradiation, the photoreactor was evacuated for 30 min to thoroughly exhaust the air. The produced hydrogen amount was detected on line by gas chromatography (GC-7920) every 1 h. The total reaction time was 4 h. The irradiation area is 19.6 cm2 and the average energy density is 59.506 mW/cm2. The AQE for H2 evolution was measured under the same conditions, except that the UV cutoff filter was replaced by a 420 nm band-pass filter. The AQE was calculated using the following equation

4.4.

Acknowledgments

This work was financially supported by the Key Research Projects of Science and Technology Department of Sichuan Province (2020YJ0260).

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.0c05106.

  • XRD, SEM, and XPS of the samples after reaction; H2 production rate (λ ≥ 510 nm); H2 production rate (λ ≥ 420 nm) of different batches of CMS/12CCS samples; Tauc plots; Mott–Schottky plots; table of element content in CMS/12CCS; BET; and comparison of H2 evolution rate and AQE of MnCdS-based catalysts (PDF)

The authors declare no competing financial interest.

Supplementary Material

ao0c05106_si_001.pdf (1.2MB, pdf)

References

  1. Dawood F.; Anda M.; Shafiullah G. M. Hydrogen production for energy: An overview. Int. J. Hydrogen Energy 2020, 45, 3847–3869. 10.1016/j.ijhydene.2019.12.059. [DOI] [Google Scholar]
  2. Nicita A.; Maggio G.; Andaloro A. P. F.; Squadrito G. Green hydrogen as feedstock: Financial analysis of a photovoltaic-powered electrolysis plant. Int. J. Hydrogen Energy 2020, 45, 11395–11408. 10.1016/j.ijhydene.2020.02.062. [DOI] [Google Scholar]
  3. Wang J.; Zhang Y.; Wang X.; Su W. Simultaneous enhancements in photoactivity and anti-photocorrosion of Z-scheme Mn0.25Cd0.75S/WO3 for solar water splitting. Appl. Catal., B 2020, 268, 118444. 10.1016/j.apcatb.2019.118444. [DOI] [Google Scholar]
  4. Kojima Y.; Yamaguchi M. Ammonia storage materials for nitrogen recycling hydrogen and energy carriers. Int. J. Hydrogen Energy 2020, 45, 10233–10246. 10.1016/j.ijhydene.2020.01.145. [DOI] [Google Scholar]
  5. Xiang Z.; Nan J.; Deng J.; Shi Y.; Zhao Y.; Zhang B.; Xiang X. Uniform CdS-decorated carbon microsheets with enhanced photocatalytic hydrogen evolution under visible-light irradiation. J. Alloys Compd. 2019, 770, 886–895. 10.1016/j.jallcom.2018.08.214. [DOI] [Google Scholar]
  6. Liu Y.; Gong Z.; Lv H.; Ren H.; Xing X. Rational design of Au decorated Mn0.5Cd0.5S/WO3 step-scheme heterostructure with multichannel charge transfer and efficient H2 generation. Appl. Surf. Sci. 2020, 526, 146734. 10.1016/j.apsusc.2020.146734. [DOI] [Google Scholar]
  7. Zhai H.; Liu X.; Wang P.; Huang B.; Zhang Q. Enhanced photocatalytic H 2 production of Mn 0.5 Cd 0.5 S solid solution through loading transition metal sulfides XS (X = Mo, Cu, Pd) cocatalysts. Appl. Surf. Sci. 2018, 430, 515–522. 10.1016/j.apsusc.2017.08.056. [DOI] [Google Scholar]
  8. Wang J.; Luo J.; Liu D.; Chen S.; Peng T. One-pot solvothermal synthesis of MoS2-modified Mn0.2Cd0.8S/MnS heterojunction photocatalysts for highly efficient visible-light-driven H2 production. Appl. Catal., B 2019, 241, 130–140. 10.1016/j.apcatb.2018.09.033. [DOI] [Google Scholar]
  9. Huang Q.-Z.; Tao Z.-J.; Ye L.-Q.; Yao H.-C.; Li Z.-J. Mn0.2Cd0.8S nanowires modified by CoP3 nanoparticles for highly efficient photocatalytic H2 evolution under visible light irradiation. Appl. Catal., B 2018, 237, 689–698. 10.1016/j.apcatb.2018.06.040. [DOI] [Google Scholar]
  10. Ikeue K.; Shinmura Y.; Machida M. Ag-doped Mn–Cd sulfide as a visible-light-driven photocatalyst for H2 evolution. Appl. Catal., B 2012, 123–124, 84–88. 10.1016/j.apcatb.2012.04.019. [DOI] [Google Scholar]
  11. Liu Y.; Ren H.; Lv H.; Gong Z.; Cao Y. An insight into the trifunctional roles of Cu2(OH)2CO3 cocatalyst in boosting the photocatalytic H2 evolution activity over Zn0.5Cd0.5S nanoparticles. Appl. Surf. Sci. 2019, 484, 1061–1069. 10.1016/j.apsusc.2019.04.148. [DOI] [Google Scholar]
  12. Liu H.; Meng J.; Zhang J. Self-assembled three-dimensional flowerlike Mn0.8Cd0.2S microspheres as efficient visible-light-driven photocatalysts for H2 evolution and CO2 reduction. Catal. Sci. Technol. 2017, 7, 3802–3811. 10.1039/c7cy01102d. [DOI] [Google Scholar]
  13. Huang Q.-Z.; Wang J.-C.; Ye L.-Q.; Zhang Q.; Yao H.-C.; Li Z.-J. A novel p–n heterojunction Mn 0.25 Cd 0.75 S/Co 3 O 4 for highly efficient photocatalytic H 2 evolution under visible light irradiation. J. Taiwan Inst. Chem. Eng. 2017, 80, 570–577. 10.1016/j.jtice.2017.08.030. [DOI] [Google Scholar]
  14. Liu H.; Xu Z.; Zhang Z.; Ao D. Novel visible-light driven Mn0.8Cd0.2S/g-C3N4 composites: Preparation and efficient photocatalytic hydrogen production from water without noble metals. Appl. Catal., A 2016, 518, 150–157. 10.1016/j.apcata.2015.08.026. [DOI] [Google Scholar]
  15. Yang J.; Wang D.; Han H.; Li C. Roles of Cocatalysts in Photocatalysis and Photoelectrocatalysis. Acc. Chem. Res. 2013, 46, 1900–1909. 10.1021/ar300227e. [DOI] [PubMed] [Google Scholar]
  16. Kumar A.; Rana A.; Sharma G.; Naushad M.; Al-Muhtaseb A.; Guo C.; Iglesias-Juez A.; Stadler F. J. High-Performance Photocatalytic Hydrogen Production and Degradation of Levofloxacin by Wide Spectrum-Responsive Ag/Fe3O4 Bridged SrTiO3/g-C3N4 Plasmonic Nanojunctions: Joint Effect of Ag and Fe3O4. ACS Appl. Mater. Interfaces 2018, 10, 40474–40490. 10.1021/acsami.8b12753. [DOI] [PubMed] [Google Scholar]
  17. Jin J.; Wang C.; Ren X.-N.; Huang S.-Z.; Wu M.; Chen L.-H.; Hasan T.; Wang B.-J.; Li Y.; Su B.-L. Anchoring ultrafine metallic and oxidized Pt nanoclusters on yolk-shell TiO2 for unprecedentedly high photocatalytic hydrogen production. Nano Energy 2017, 38, 118–126. 10.1016/j.nanoen.2017.05.025. [DOI] [Google Scholar]
  18. Pan B.; Luo S.; Su W.; Wang X. Photocatalytic CO2 reduction with H2O over LaPO4 nanorods deposited with Pt cocatalyst. Appl. Catal., B 2015, 168–169, 458–464. 10.1016/j.apcatb.2014.12.046. [DOI] [Google Scholar]
  19. Liu Y.; Gong Z.; Xie Y.; Lv H.; Zhang B. Revealing the synergetic effects of graphene and MoS2 on boosted photocatalytic H2 production of Mn0.5Cd0.5S photocatalyst. Appl. Surf. Sci. 2020, 505, 144637. 10.1016/j.apsusc.2019.144637. [DOI] [Google Scholar]
  20. Liu X.; Liang X.; Wang P.; Huang B.; Qin X.; Zhang X.; Dai Y. Highly efficient and noble metal-free NiS modified MnxCd1-xS solid solutions with enhanced photocatalytic activity for hydrogen evolution under visible light irradiation. Appl. Catal., B 2017, 203, 282–288. 10.1016/j.apcatb.2016.10.040. [DOI] [Google Scholar]
  21. Liao L.; Zhu J.; Bian X.; Zhu L.; Scanlon M. D.; Girault H. H.; Liu B. MoS2 Formed on Mesoporous Graphene as a Highly Active Catalyst for Hydrogen Evolution. Adv. Funct. Mater. 2013, 23, 5326–5333. 10.1002/adfm.201300318. [DOI] [Google Scholar]
  22. Biroju R. K.; Das D.; Sharma R.; Pal S.; Mawlong L. P. L.; Bhorkar K.; Giri P. K.; Singh A. K.; Narayanan T. N. Hydrogen Evolution Reaction Activity of Graphene–MoS2 van der Waals Heterostructures. ACS Energy Lett. 2017, 2, 1355–1361. 10.1021/acsenergylett.7b00349. [DOI] [Google Scholar]
  23. Ghasemi Z.; Abdi V.; Sourinejad I. Green fabrication of Ag/AgCl@TiO2 superior plasmonic nanocomposite: Biosynthesis, characterization and photocatalytic activity under sunlight. J. Alloys Compd. 2020, 841, 155593. 10.1016/j.jallcom.2020.155593. [DOI] [Google Scholar]
  24. Liang Q.; Zhang C.; Xu S.; Zhou M.; Li Z. Nanocomposites based on 3D honeycomb-like carbon nitride with Cd0.5Zn0.5S quantum dots for efficient photocatalytic hydrogen evolution. Int. J. Hydrogen Energy 2019, 44, 29964–29974. 10.1016/j.ijhydene.2019.09.180. [DOI] [Google Scholar]
  25. Zhang X.; Guo Y.; Tian J.; Sun B.; Liang Z.; Xu X.; Cui H. Controllable growth of MoS2 nanosheets on novel Cu2S snowflakes with high photocatalytic activity. Appl. Catal., B 2018, 232, 355–364. 10.1016/j.apcatb.2018.03.074. [DOI] [Google Scholar]
  26. Moosavifard S. E.; Fani S.; Rahmanian M. Hierarchical CuCo2S4 hollow nanoneedle arrays as novel binder-free electrodes for high-performance asymmetric supercapacitors. Chem. Commun. 2016, 52, 4517–4520. 10.1039/c6cc00215c. [DOI] [PubMed] [Google Scholar]
  27. Jin R.; Cui Y.; Gao S.; Zhang S.; Yang L.; Li G. CNTs@NC@CuCo2S4 nanocomposites: An advanced electrode for high performance lithium-ion batteries and supercapacitors. Electrochim. Acta 2018, 273, 43–52. 10.1016/j.electacta.2018.03.122. [DOI] [Google Scholar]
  28. Bao N.; Shen L.; Takata T.; Domen K.; Gupta A.; Yanagisawa K.; Grimes C. A. Facile cd-thiourea complex thermolysis synthesis of phase-controlled CdS nanocrystals for photocatalytic hydrogen production under visible light. J. Phys. Chem. C 2007, 111, 17527–17534. 10.1021/jp076566s. [DOI] [Google Scholar]
  29. Liu X.; Wang W.; Liu Y.; Huang B.; Dai Y.; Qin X.; Zhang X. In situ synthesis of Bi2S3/Bi2SiO5 heterojunction photocatalysts with enhanced visible light photocatalytic activity. RSC Adv. 2015, 5, 55957–55963. 10.1039/c5ra09070a. [DOI] [Google Scholar]
  30. Zhong Y.; Yuan J.; Wen J.; Li X.; Xu Y.; Liu W.; Zhang S.; Fang Y. Earth-abundant NiS co-catalyst modified metal-free mpg-C3N4/CNT nanocomposites for highly efficient visible-light photocatalytic H2 evolution. Dalton Trans. 2015, 44, 18260–18269. 10.1039/c5dt02693h. [DOI] [PubMed] [Google Scholar]
  31. Zhou W.; Jia J.; Lu J.; Yang L.; Hou D.; Li G.; Chen S. Recent developments of carbon-based electrocatalysts for hydrogen evolution reaction. Nano Energy 2016, 28, 29–43. 10.1016/j.nanoen.2016.08.027. [DOI] [Google Scholar]
  32. Liu X.; Liu Q.; Wang P.; Liu Y.; Huang B.; Rozhkova E. A.; Zhang Q.; Wang Z.; Dai Y.; Lu J. Efficient photocatalytic H2 production via rational design of synergistic spatially-separated dual cocatalysts modified Mn0.5Cd0.5S photocatalyst under visible light irradiation. Chem. Eng. J. 2018, 337, 480–487. 10.1016/j.cej.2017.12.125. [DOI] [Google Scholar]
  33. Li L.; Xu J.; Ma J.; Liu Z.; Li Y. A bimetallic sulfide CuCo2S4 with good synergistic effect was constructed to drive high performance photocatalytic hydrogen evolution. J. Colloid Interface Sci. 2019, 552, 17–26. 10.1016/j.jcis.2019.05.036. [DOI] [PubMed] [Google Scholar]
  34. Li H.; Wang Z.; He Y.; Meng S.; Xu Y.; Chen S.; Fu X. Rational synthesis of MnxCd1-xS for enhanced photocatalytic H2 evolution: Effects of S precursors and the feed ratio of Mn/Cd on its structure and performance. J. Colloid Interface Sci. 2019, 535, 469–480. 10.1016/j.jcis.2018.10.018. [DOI] [PubMed] [Google Scholar]
  35. Zhang D.; Tang Y.; Qiu X.; Yin J.; Su C.; Pu X. Use of synergistic effects of the co-catalyst, p-n heterojunction, and porous structure for improvement of visible-light photocatalytic H2 evolution in porous Ni2O3/Mn0.2Cd0.8S/Cu3P@Cu2S. J. Alloys Compd. 2020, 845, 155569. 10.1016/j.jallcom.2020.155569. [DOI] [Google Scholar]
  36. Li X.; Yu J.; Low J.; Fang Y.; Xiao J.; Chen X. Engineering heterogeneous semiconductors for solar water splitting. J. Mater. Chem. A 2015, 3, 2485–2534. 10.1039/c4ta04461d. [DOI] [Google Scholar]
  37. Kronik L.; Shapira Y. Surface photovoltage phenomena: theory, experiment, and applications. Surf. Sci. Rep. 1999, 37, 1–206. 10.1016/s0167-5729(99)00002-3. [DOI] [Google Scholar]
  38. Jiang J.; Zhang X.; Sun P.; Zhang L. ZnO/BiOI Heterostructures: Photoinduced Charge-Transfer Property and Enhanced Visible-Light Photocatalytic Activity. J. Phys. Chem. C 2011, 115, 20555–20564. 10.1021/jp205925z. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ao0c05106_si_001.pdf (1.2MB, pdf)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES