Next Article in Journal
Photoelectrochemical Behavior of the Ternary Heterostructured Systems CdS/WO3/TiO2
Next Article in Special Issue
A Review on Quantum Dots Modified g-C3N4-Based Photocatalysts with Improved Photocatalytic Activity
Previous Article in Journal
Influence of Base-Catalyzed Organosolv Fractionation of Larch Wood Sawdust on Fraction Yields and Lignin Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Carbon Dioxide Photoreduction on the Bi2S3/MoS2 Catalyst

1
Department of Chemistry, College of Natural Sciences, Yeungnam University, Gyeongsan, Gyeongbuk 38541, Korea
2
Korea Institute of Energy Research, 152 Gajeong-ro, Yuseong-gu, Daejeon 34129, Korea
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(12), 998; https://doi.org/10.3390/catal9120998
Submission received: 8 November 2019 / Revised: 23 November 2019 / Accepted: 25 November 2019 / Published: 27 November 2019
(This article belongs to the Special Issue Advances in Solar- and Visible-Light Photocatalysis)

Abstract

:
The photocatalytic activity of a material is contingent on efficient light absorption, fast electron excitation, and control of the recombination rate by effective charge separation. Inorganic materials manufactured in unique shapes via controlled synthesis can exhibit significantly improved properties. Here, n-type Bi2S3 nanorods (with good optical activity) were wrapped with two-dimensional (2D) p-type MoS2 sheets, which have good light absorption properties. The designed p-n junction Bi2S3/MoS2 composite exhibited enhanced light absorption over the entire wavelength range, and higher carbon dioxide adsorption capacity and photocurrent density compared to the single catalysts. Consequently, the activity of the 1Bi2S3/1MoS2 composite catalyst for the photocatalytic reduction of carbon dioxide was more than 20 times higher than that of the single catalysts under visible-light irradiation at ≤400 nm, with partial selectivity for CO conversion. This is attributed to the p-n heterojunction Bi2S3/MoS2 composite designed in this study, the high light absorption of n-Bi2S3, accelerated electron excitation, and the electron affinity of the 2D sheet-p-MoS2, which quickly absorbed excited electrons, resulting in effective charge separation. This ultimately improved the catalytic performance by continuously supplying catalytically active sites to the heterojunction interfaces.

1. Introduction

The greatest challenge for an eco-friendly CO2-to-fuel system, which produces fuel by photocatalytically reducing carbon dioxide under visible light, is the development of stable and highly efficient photocatalysts. Chalcogenides are currently highly topical as candidate materials for this purpose and extensive studies of these materials have been conducted. Chalcogenides are an attractive research theme because their physicochemical properties vary with their unique shape and size. Furthermore, the shape and size of these materials can be easily controlled by the synthesis methods [1,2]. In particular, chalcogenides have small band-gaps and large extinction coefficients, and can be utilized in various applications such as batteries [3], gas sensors [4], and photodetectors [5]. In recent years, due to their wide light absorption ability, the application of chalcogenides as photocatalysts that are active under visible-light irradiation has been actively studied.
Chalcogen compounds, such as CdS [6], ZnS [7], SnS2 [8], NiS [9], Bi2S3 [10], CuS [11], and MoS2 [12], can be used for water decomposition or CO2 conversion, and these have produced quite reliable results. Specifically, CdS has been evaluated as one of the most promising catalyst candidates for water decomposition under visible light, which is thermodynamically advantageous, as the positions of the valence and conduction bands overlap with the oxidation and reduction potentials of water [13,14]. However, CdS suffers from the inherent disadvantage of low efficiency due to inefficient charge separation between the electrons and holes generated by the absorption of photons, and is self-decomposing due to its structural instability in water [15]. Therefore, the catalytic performance varies widely depending on the degree of overlap between the redox potential of the selected chemical reaction and the band-gap of the catalyst. Thus, designing a stable and highly active chalcogen photocatalyst for a given reaction is a very attractive challenge. Additionally, by controlling the morphology of chalcogenides, significantly improved properties can be derived from the unique shapes generated at various stages of synthesis.
Here, p-type MoS2 with excellent light absorption ability is prepared as 2D-structured nanosheets, and n-type Bi2S3 with excellent optical activity and a nanorod morphology is also prepared. A heterojunction comprising the two types of particles is designed to create a heterojunction composite system. Weng and coworkers suggested that the use of MoS2 nanosheets to coat the surface of Bi2S3 discoids by a facile anion-exchange strategy accelerates the light-harvesting efficiency and charge separation, and helps faster charge transport and collection, leading to higher activity in the photocatalytic reduction of Cr(VI) under visible-light irradiation (λ > 400 nm) [16]. As another example, Ke and coworkers reported that a prepared Bi2O3/Bi2S3/MoS2 nanocomposite exhibits improved photocatalytic activity for water oxidation (529.1 μmol h−1 g−1cat), with a value that is 1.5 and 12.5 times higher than that of pure Bi2O3 and MoS2, respectively, under simulated solar-light irradiation [17]. Ultimately, in this study, we propose that p-type MoS2 particles, which lack electrons, will easily accept electrons excited by n-type Bi2S3 particles that have a narrow band-gap (which facilitates excitation of the electrons by the absorbed visible light [18]). Consequently, it was expected that the cycle of electron flow would proceed smoothly and separation between the electrons and holes would progress effectively. Furthermore, we anticipated that sulfur deficient sites [19] would be formed at the new interfaces between the two particles, which would eventually act as reaction sites, leading to enhanced catalytic activity.

2. Results and Discussion

2.1. Physical Properties of Catalysts

Figure 1 shows the X-ray diffraction patterns of the as-synthesized Bi2S3, MoS2, 1Bi2S3/1MoS2, 2Bi2S3/1MoS2, and 1Bi2S3/2MoS2 particles. The XRD (X-ray diffraction) pattern of as-synthesized Bi2S3 indicated the formation of orthorhombic phase Bi2S3 (JCPDS No. 002-0391) [20]. No other crystalline phases were detected in the product. The XRD patterns show that the sample was mainly composed of rhombohedral MoS2, and the diffraction peaks at 2θ = 14.5°, 33°, 40°, and 58° are ascribed to the (002), (100), (103), and (110) lattice planes of crystalline MoS2 (JCPDS card No. 037-1492) [21]. Extremely broad reflection peaks were observed in the profile of the MoS2 particles, which indicated that the obtained MoS2 sample was nano-sized. The XRD profile of the heterojunction Bi2S3/MoS2 catalysts clearly showed the characteristic peaks of Bi2S3, but the peak assigned to the (002) plane of MoS2 was hardly visible. The expanded spacing of the MoS2 (002) plane resulted from the synthesis conditions. The spacing usually increases in the alkali solution, and in addition, the environmental temperature also has a critical effect. In particular, the expanding becomes large in both the alkali environment and the low temperature. At this time, it is reported that the new shifted peak (2θ = 9.3°) associated with the expanded (002) d-spacing was observed [22]. However, at temperatures higher than 220 °C, the spacing of the (002) plane of MoS2 did not be expanded [23]. In this study, perhaps both the alkali environment and the low temperature (180 °C) caused the expansion of the (002) plane of MoS2 in Bi2S3/MoS2 heterojunction materials. Furthermore, in the 1Bi2S3/2MoS2 sample, a new peak associated with the expanded (002) plane was also seen at 2θ = 9.3°. The intensity of the characteristic peaks of the Bi2S3 particles increased in proportion to the concentration added during the synthesis step. However, the crystallite sizes obtained from the Scherer equation [24] based on the (111) plane, which is the characteristic peak of Bi2S3 nanoparticles, decreased slightly with increasing addition of MoS2. The crystallite sizes were 10, 9, 8, and 7 nm, respectively, for Bi2S3, 1Bi2S3/1MoS2, 2Bi2S3/1MoS2, and 1Bi2S3/2MoS2. From these results, we predicted that the crystallite sizes of Bi2S3 in the Bi2S3/MoS2 heterojunction composites did not change and the shape was maintained, even when different amounts of Bi2S3 and MoS2 were incorporated into the heterocapsulated catalysts.
The Bi2S3 particles comprising 50–60 nm wide and 100–150 nm long rods are shown in the TEM (Transmission Electron Microscope) image in Figure 2. The MoS2 sample had a wide and thin sheet morphology. In the heterojunction Bi2S3/MoS2 samples, the rod-shaped Bi2S3 was wrapped by a thin sheet of MoS2. Increasing the amount of MoS2 led to complete wrapping of the Bi2S3 particles by MoS2. From these results, it was deduced that the two particles were uniformly structured and aligned for wrapping by electrostatic attraction during heterojunction formation.
HRTEM (High-resolution transmission electron microscope) (Figure 3A), selected area electron diffraction (SAED) (Figure 3B), and element mapping analysis (Figure 3C) were performed for 1Bi2S3/1MoS2 as a prototype of the heterojunction catalysts, and the results are shown in Figure 3. The fringe spacing parallel to the nanorod was estimated to be 3.55 and 6.34 Å in Bi2S3 and MoS2, which indicates that crystal growth occurred preferentially in the (111) and (002) directions, respectively. Additionally, some of the characteristic peaks identified in the XRD pattern could be seen in this figure. The SAED pattern of Bi2S3 recorded for the regularly spaced orthorhombic phase exhibited diffraction rings, thereby indicating that the orthorhombic phase was polycrystalline and could be readily indexed as the Bi2S3 phase. A circle attributed to the (002) plane was also observed, which also corresponds to the rhombohedral MoS2 nanosheet. Based on these results, we deduced that each of the crystallized particles, 1Bi2S3 and 1MoS2, existed stably in the 1Bi2S3/1MoS2 heterojunction particle with no crystal change. However, element mapping by TEM confirmed that all components (1Bi2S3 or 1MoS2) of the catalyst were uniformly dispersed in the catalyst; particularly, it was confirmed that MoS2 uniformly wrapped Bi2S3 as an outer layer. Moreover, there were almost no impurities such as C and O.
The EDS (Energy Dispersive X-ray Spectroscopy) spectrum in Figure 4 showed the presence of only Bi and S on the surface of Bi2S3, and the atomic ratio of Bi:S was close to the stoichiometric value of 2:3, as shown in the inset table. In the EDS spectrum of the MoS2 particles, only Mo and S components were observed, and no other impurities were found. The ratio of Mo:S, as shown in the inset table (Figure 4), was about 1:2. The atomic ratio of Bi2S3 and MoS2 once again suggested reliability of the synthesis. These ratios were almost the same as the theoretical values used in the synthesis step. On the other hand, all the components of the Bi2S3/MoS2 heterojunction particles were detected in the EDS spectrum, and the intensities of the peaks were proportional to the contents. The theoretical metal/S ratio of the 1Bi2S3/1MoS2 particles is 1:1.68. However, the actual surface ratio of Mo + Bi/S was 1:1.73, where the sulfur content was relatively high. In the 2Bi2S3/1MoS2 heterojunction particle, the theoretical ratio of Mo + Bi/S is 1:1.6, but the actual surface ratio was 1:1.9. For the 1Bi2S3/2MoS2 heterojunction particles, the theoretical ratio of Mo + Bi/S is 1:1.75, but the surface composition ratio of Mo + Bi/S was 1:1.99. This result suggests that more elemental sulfur is exposed on the surface of the Bi2S3/MoS2 heterojunction particles than on the surfaces of the pure particles. Ultimately, this result is also related to the disulfide peaks observed in the XPS (X-ray Photoelectron Spectroscopy) profile and may be responsible for defects in the edges of the heterojunction particles.
Typical high-resolution quantitative Bi 4f, Mo 3d, and S 2p in Bi2S3, MoS2, and 1Bi2S3/1MoS2 XPS spectra are presented in Figure 5. The characteristic orbital peaks occurred at different positions in the full scan for each sample (Figure 5a). The highest intensity peaks for the Mo 3d, Bi 4f, and S 2p states are presented in Figure 5b. The deconvoluted peaks of the Mo 3d5/2 and Mo 3d3/2 states, located at binding energies of 229.1 and 232.4 eV, respectively, are attributed to Mo4+ in the MoS2 sample [25]. A small peak due to Mo6+ [26] was observed at 235.5 eV, but this peak was negligibly small. The Mo 3d5/2 peak located at the lower binding energy of 228.4 eV in the profile of the 1Bi2S3/1MoS2 sample corresponds to the reduced Mo atoms due to sulfur vacancies [27]. Moreover, an overlapping peak of Mo6+ and a singlet S 1s peak was observed at 225.7 eV due to sulfur defects [27]. Additionally, Kwok and coworkers reported the presence of edge S-defect sites (S 2p) in MoS2 nanosheets within a mesoporous silica shell [28], such as MoV=O sites that are comparable to molybdenum oxysulfides (MoOxSy), bridging disulfides (S–S)br2−, shared disulfides (S–S)sh2−, and terminal disulfides (S–S)t2−. Notably, the disulfide ligands at these defect sites can be easily removed to expose unsaturated Mo4+ sites, and have been suggested as the active sites. In this study, the O 1s and S 2p peaks were also observed in the full range spectrum of MoS2 (Figure 5a). This result means that Mo present in the 1Bi2S3/1MoS2 heterojunction catalyst is not present as MoS2, but as MoOxSy. Eventually, this results in more defects in MoS2, which are believed to act as catalytically active sites. Two peaks respectively corresponding to the Bi 4f5/2 and Bi 4f7/2 states of Bi3+ appeared at 163.8 and 158.4 eV in the profile of the Bi2S3 sample (Figure 5b). Bi2S3 remain unchanged after the formation of Bi2S3/MoS2 heterojunction. The peaks at 161 and 162.3 eV could be assigned to the S 2p state, and the S 2s XPS peak at the higher binding energy of 225.7 eV is consistent with the Bi-S species [29]. From these results, we confirmed that after heterojunction formation, there was no significant change in the Bi2S3 crystal, but the oxidation state of Mo in the MoS2 crystal was lowered. Ultimately, this is believed to lead to the formation of S defects and to increase the capacity for carbon dioxide adsorption.

2.2. Carbon Dioxide Photoreduction on Catalysts

Before entering the full-scale CO2 photoreduction experiment, an isotopic experiment using 13CO2 as the substrate was performed under the same photocatalytic reaction conditions to further verify the source of the generated CO, and the products were analyzed by gas chromatography and mass spectrometry. The peak at m/z = 29 could be assigned to 13CO, indicating that the carbon source for the formation of CO was the CO2 used. When the CO2 was replaced with N2, no detectable products were formed. When CO2 was subsequently added to the reaction mixture with the clean photocatalyst (Ar gas/H2O/catalyst/CO2 gas/light on), the product formation yield increased, providing strong evidence of product formation from CO2. This means that the final carbon compounds produced were exclusively generated by the reduction of externally injected CO2, which was the only carbon source in this study. Additionally, isotopic study of 1Bi2S3/1MoS2 in the presence of moist 13CO2 was performed, as shown in Figure 6. Various signals of isotopic carbon compounds, 13C, 13CH2, 13CH3, 13CH4, 13CO, 13CH3OH, and H13COOH were observed at m/z = 13, 15, 16, 17, 29, 33, and 47, respectively, and the 13CO2 carbon source was detected at m/z = 45. Generally, the overall photoreduction reaction of CO2 to CH4 is a very complicated process that includes the transfer of eight electrons and eight protons, breaking of C–O bonds, and the formation of C–H bonds. There are two pathways, and both pathways preferentially follow the two-electron mechanism. Here, the formaldehyde pathway (fast hydrogenation pathway) follows the path CO2 → HCOOH → HCHO → CH3OH → CH4, which is thermodynamically feasible [30]. However, no HCHO was detected in the GC-MS analysis. The favorable production of HCOOH as an intermediate increases CO production, and the favorable production of HCHO increases CH4 production. Mass analysis showed that the 1Bi2S3/1MoS2 catalyst produced slightly more CH3OH than CH4. Overall, the production of CH4 from the complete reduction of CO2 in the presence of the 1Bi2S3/1MoS2 heterojunction catalysts seems to be unfavorable.
CO2 reduction was conducted under visible-light irradiation. As illustrated in Figure 7, the amount of CO generated increased almost linearly with the irradiation time for all of the catalysts. However, the amount of CO generated with the pure MoS2 and Bi2S3 catalysts was minimal, whereas the conversion to CO in the presence of the heterojunctioned Bi2S3/MoS2 catalysts increased surprisingly. There was no selectivity for CH4 when only Bi2S3 was used, whereas the heterojunctioned catalysts exhibited selectivity. The total amount of CO and CH4 generated from CO2 was similar for the 2Bi2S3/1MoS2 and 1Bi2S3/2MoS2 catalysts after 10 h. However, a high CO evolution of 40 μmol g−1 (42.5 μmol·g−1 for total amounts of CO and CH4) was achieved with the 1Bi2S3/1MoS2 catalyst, having a junction comprising 1:1 Bi2S3 and MoS2, under photo-illumination for 10 h. These results reveal that the activity of 1Bi2S3/1MoS2 for CO2 reduction was comparable to that of other catalysts. The total yield obtained with the 1Bi2S3/1MoS2 composite did not change significantly after five cycles (1 cycle corresponds to 10 h of reaction), as shown in Figure 7B), indicating the excellent stability of the 1Bi2S3/1MoS2 heterojunction catalyst.

2.3. Optical Properties of Bi2S3/MoS2 Catalysts

The UV-vis diffuse reflectance spectra were used to analyze the light absorption and energy band features of the Bi2S3/MoS2 catalysts, as illustrated in Figure 8a. The absorption edge of Bi2S3 was close to 970 nm; thus, the compound exhibited a photo-response from the visible to near-infrared region. In comparison, the pure MoS2 nanosheets also presented a broad absorption band around 670 nm, which indicates that they can absorb photons over almost all of the visible region. Furthermore, the Bi2S3/MoS2 heterojunction catalysts absorbed light over the full spectral region due to the transition of MoS2 and Bi2S3. Other factors affect the observed photocatalytic efficiency, such as the optical range, containment of the charge transfer, specific surface area, and the suppression of the recombination of electron–hole pairs in the MoS2/Bi2S3 heterojunction. Nevertheless, the highest absorbance point was around 670 nm. To determine the relative band-gap and the energy levels of all the catalysts, the band-gap energy (Eg) was determined by applying the following formula based on the differential reflectance spectroscopy (DRS) data [31]: the hetero-junction particles are direct band-gap semiconductors. The reflectance data can be converted to absorption according to the Kubelka−Munk (K-M) theory: F(R) = (1−R)2/2R, where, R is reflectance and F(R) is the K-M function [32]. The band gap of the samples can be estimated by using Tauc plot: (αhν)1/n = A (hν−Eg), where, α, hν, A, and Eg are the absorption coefficient, incident light frequency, proportionality constant, and band gap, respectively. The value of exponent n determines the nature of electronic transition; n = 1/2 for direct transition and n = 2 for indirect transition. The linear extrapolation of (αhν)1/n to zero give the band gap energies of the samples. According to Tauc's relation, the plotting of (αhυ)1/2 versus the photon energy (hυ) gives a straight line in a certain region. The extrapolation of this straight line will intercept the (hυ)-axis to give the value of the indirect optical energy gap (Eg). From the Tauc plots presented in Figure 8B, the Egs of Bi2S3 and MoS2 (determined from the plot of (ahν)2 versus (hν)) were found to be 1.32 and 1.58 eV, respectively. The band-gap of the heterogeneous catalysts was intermediate between the band-gap of these two particles.
To determine the valence band (VB) and conduction band (CB) edge potential of the semiconductors, the VB positions were determined from XPS, as shown in Figure 9a; the VBs of MoS2 and Bi2S3 were 0.4 and 1.05 eV, respectively. Thus, by combining the XPS spectral data and band-gaps in Figure 8, the CB position was determined by applying the following equation, Eg = CB − VB [33]. In conclusion, the conductance bands of MoS2 and Bi2S3 were −0.62 and −0.88 eV, respectively, as illustrated in Figure 9b. Based on these findings, it can be seen that the Bi2S3 rod and MoS2 sheet can form a perfect heterojunction.
To understand the charge carrier transfer behavior in the catalysts, photocurrent experiments were carried out and the photocurrent responses are shown in Figure 10. For the photocurrent density cycle, the analytical samples were prepared as pastes and coated on the cell-type FTO (Fluorine doped tin oxide) glass. This cell was used as the working electrode, and the Pt-coated FTO glass was used as the counter-electrode. The two electrodes were connected to form a single system. An iodolyte electrolyte (AN-50, Solarnonix) was used and sufficiently immersed in the electrode, and a potential of 0 V was applied. The current was measured under one-sun illumination with a 2000 Solar Simulator (IVIUM STAT, ABET Technologies). After the initial 1 min stabilization period, the current was measured during periodic irradiation at intervals of 30 s. The cell area (0.4 cm2) was divided by the measured current, and the current density was calculated. The photocurrent was 30 nA for Bi2S3 and 1185 nA for MoS2. When the amount of MoS2 in the heterojunction catalyst was increased, the photocurrent increased with the active photoirradiation (light on) time; when the amount of Bi2S3 was increased, the current was small but stable during active photoirradiation. The 1Bi2S3/1MoS2 heterojunction showed the best photocurrent response, which was 1.5 and 58 times that of pristine MoS2 and Bi2S3, respectively. This result confirms slower recombination of the charge carriers in the 1Bi2S3/1MoS2 heterojunction, indicating significant improvement of the photocatalytic performance. Moreover, the steady and reproducible photocurrent response after several cycles suggests that the 1Bi2S3/1MoS2 heterojunction possessed good photostability.

2.4. Catalytic Mechanism for 1Bi2S3/1MoS2

Figure 11 shows the CO2 gas adsorption capacity (CO2–TPD (Temperature programmed desorption) curve) of the Bi2S3, MoS2, and 1Bi2S3/1MoS2 particles. In order to remove impurities, the particles were treated at 300 °C for 30 min, and CO2 gas was adsorbed at 50 °C for 2 h.
The temperature was increased to 550 °C at a rate of 5 °C min−1 to determine the desorption temperature and curves in order to determine the amount of CO2 gas adsorbed on the catalysts. The results indicated that CO2 gas was desorbed above 300 °C; the amount of CO2 desorbed was small, but the desorption temperature was high in the case of Bi2S3. The CO2 desorption curves were separated into peaks at 324.2 and 484.4 °C for the MoS2 sample, and the amount of CO2 gas desorbed by the MoS2 sample was significantly greater than that desorbed by Bi2S3. As noted above, this was attributed to the amount of sulfur vacancies in MoS2, as shown in the XPS profile. CO2 is expected to be adsorbed in the vacant sulfur sites, and therefore, more sulfur vacancies lead to stronger attraction between CO2 gas and the particles. However, in the case of 1Bi2S3/1MoS2, desorption occurred at lower temperatures and the highest desorption was achieved at 392.6 °C. However, the desorption amount decreased relative to that achieved with the single MoS2 catalyst. This is because the amount of MoS2 present per unit weight was reduced. Nevertheless, the amount of CO2 adsorbed was high enough to react with water. Moreover, the lower CO2 desorption temperature suggests that the adsorbed CO2 can easily be converted in the next step. This also means that even after conversion to the product by the reaction with water, the product was easily desorbed, resulting in less coking on the catalyst. This in turn extended the catalytic performance.
To comprehend the method of transfer of the photo-induced charge, a possible mechanism for photocatalysis by the Bi2S3/MoS2 heterojunction was proposed, as shown in Scheme 1.
When the Bi2S3/MoS2 heterojunction is excited by visible light, the electrons move from the CB of the Bi2S3 rods to the CB of the MoS2 nanosheets. The holes are then transferred from the VB of MoS2 to the VB of Bi2S3. Excited electrons naturally flow through the CB surface of Bi2S3 to the CB of MoS2 at lower potential (−0.62 eV vs. NHE (Normal Hydrogen Electrode)); thus, the density of electrons at the interface of these two particles will plausibly increase. Finally, CO2 gas is adsorbed on the electron-rich junction surface or the VB surface of MoS2 based on the gas adsorption data presented in Figure 11, and these electrons can reduce CO2 to produce CO. This leads to efficient separation between the electrons and holes in this Bi2S3/MoS2 heterojunction catalytic system. In theory, the photo-induced electrons in the VB of Bi2S3 (0.4 eV vs. NHE) are not effective for the decomposition of water to produce H+ ions; thus, the CB of Bi2S3 cannot participate in reduction to yield H radicals or H2 gas. This reduces the possibility of interaction of the CO intermediate with hydrogen, which in turn leads to lower CH4 yields. Overall, the yield of CO was increased and the yield of CH4 was very low.

3. Materials and Methods

3.1. Preparation of Catalysts

As shown in Scheme 2, the catalysts were prepared by a hydrothermal method.
Bi2S3 was synthesized as follows: 0.25 g (3.28 mmol) of thiourea (SC(NH2)2, MW = 76.12 g mol−1, Aldrich) was dissolved by stirring in 50 mL of deionized (DI) water; 0.61 g (1.26 mmol) of bismuth nitrate pentahydrate (Bi(NO3)3 5H2O, MW = 485.07 g mol−1, Aldrich) was slowly added to this solution. The homogenous solution was stirred for 4 h, then transferred to a Teflon-lined autoclave, heated to 140 °C at a rate of 10 °C min−1, and maintained for 10 h. The solution was then slowly cooled at room temperature to obtain a precipitate that was separated by centrifugation and washed several times with absolute ethanol. The washed sample was dried at 70 °C for 8 h to obtain the final compound. MoS2 was synthesized in the following order: 0.41 g (5.4 mmol) of thiourea was uniformly dissolved in 35 mL of DI water with stirring. Hexaammonium heptamolybdate tetrahydrate ((NH4)6Mo7O24 4H2O, MW = 1235.9 g mol−1, Aldrich) was slowly added to the above solution and stirred for 4 h. This solution was transferred to a Teflon-lined autoclave, heated to 220 °C at a rate of 10 °C min−1, and then held for 18 h. At the end of the reaction, the autoclave was cooled to room temperature, and the precipitate obtained was centrifuged and washed several times with anhydrous ethanol. The compound was dried at 70 °C for 8 h to obtain the final product. Finally, heterojunction Bi2S3/MoS2 catalysts were synthesized in the following manner: thiourea was dissolved in 35 mL of DI (Deionized) water. To this solution, the appropriate amounts of (NH4)6Mo7O24 4H2O and Bi(NO3)3 5H2O were slowly added and then stirred for 4 h for homogenization. The solution was transferred to a Teflon-lined autoclave, heated to 180 °C at a rate of 10 °C min−1, and held for 14 h. After the reaction, the autoclave was cooled to room temperature, and the precipitate obtained by centrifugation was washed several times with anhydrous ethanol. Finally, the compound was dried at 70 °C for 8 h to obtain the final product. The molar ratios of Bi2S3:MoS2 in the obtained samples, were 1:1, 2:1, and 1:2. In total, five types of catalysts, Bi2S3, MoS2, 1Bi2S3/1MoS2, 2Bi2S3/1MoS2, and 1Bi2S3/2MoS2, were synthesized and their optical properties and activities were investigated.

3.2. Characterization

Powder XRD patterns of the prepared catalysts were collected with an XPert Pro MPD PANalytical 2-circle diffractometer using Cu-Kα radiation (λ = 1.5406 Å), a working voltage of 30 kV, and a current of 15 mA in the angle of diffraction (2θ) range 10–80°. The morphologies and compositions of the as-prepared samples were characterized by transmission electron microscopy (TEM, H-7600, Hitachi, operating at 200 kV, Japan), energy dispersive spectrometric element mapping (EDS, EX-250, Horiba), and high-resolution transmission electron microscopy (HRTEM, JEM-2100 F, Jeol). The optical properties were confirmed from the UV-visible diffuse reflectance spectra collected on a Neosys-2000 (UV–vis, Scinco Co. Korea) spectrometer using BaSO4 as the reference sample, photoluminescence spectroscopy (PL, Perkin Elmer, He–Cd laser source, wavelength of 300 nm), and a photocurrent–time response system (IVIUM STAT, ABET Technologies). A 150 W Xe lamp was used as the light source. A saturated calomel electrode (SCE), and saturated 0.1 M Na2SO4 solution were used as the counter-electrode and electrolyte, respectively. The working electrode was prepared via a dip-coating method. Time-dependent photocurrent curves were constructed by the i–t curve method.

3.3. Photocatalytic Carbon Dioxide Conversion Reaction

The photoreduction of CO2 in the presence of the synthesized Bi2S3, MoS2, 1Bi2S3/1MoS2, 2Bi2S3/1MoS2, and 1Bi2S3/2MoS2 particles with H2O was performed in a closed cylinder-type quartz vessel (length: 15.0 cm; diameter: 1.0 cm; total volume: 12.50 mL). A schematic diagram of the photoreactor is provided in a previous report [34]. Here, 0.2 g of the catalyst and 40.0 L of distilled water were placed in the photoreactor. High-purity CO2 gas (99.99%) was used as the reactant. The chamber was purged with the CO2 gas before irradiation. The water was injected and the reactor was sealed; the outside of the reactor was wrapped with a heat tape and kept at 80 °C for 30 min to vaporize the water in the reactor. The reaction was performed after cooling the reactor to 25–30 °C. At this time, the pressure in the reactor was close to atmospheric pressure. Here, in general, the activity of a photocatalyst that decomposes organic molecules in a solution shows a tendency to increase steadily up to a reaction temperature of 60 °C, but decreases above it [35]. When the temperature is increased, it causes the bubbles formation in the solution which results in the generation of free radicals. Additionally, the increase in temperature helps the degradation reaction to overcome electron-hole recombination. In addition to that, the increasing temperature may enhance the oxidation rate of organic molecules at the interface. Zhang et al. also reported [36] when temperature increased from 323 K to 343 K, the yield of the photocatalytic reaction increased by two times. Although raising temperature is an effective strategy to obtain higher yields, the photoreduction of carbon dioxide has been reported to have the best performance at the low reaction temperature of 25 °C in some researches [37]. Thus, there is always an optimum temperature range to get the best. Here, the photoreactor in this study was an open system in which a one sun lamp illuminates a quartz photoreactor from above, with a diameter of up to 10 cm × 10 cm radiated from one sun. The distance from the end of one sun lamp to the reactor was 1.0 cm and the inner diameter of the reactor was 1.0 cm, thus the distance between the lamp and the photocatalyst sample was 2.0 cm. Moreover, the laboratory was maintained at 18 °C by the air conditioner to suppress the photoreactor temperature from rising above 25 °C by the one sun lamp. The reactor chamber was then closed and the lamp was switched on. A natural solar simulator (a xenon lamp, 150 mW cm−2, 620 nm, 2000 Solar Simulator, ABET Tech) was used as the irradiation source. The solar light intensity of the solar simulator was calculated with a digital lux meter and the average light intensity was determined as approximately 1,200,000 Lux. Photoreduction was carried out at room temperature and atmospheric pressure. The product gases were analyzed using a gas chromatography (GC; Master GC, Scinco, Korea) instrument. Hydrocarbons such as methane, ethane, and propane were quantified by flame ionization detection (FID), and quantitative analysis of CO, H2, O2, CH3OH, etc., was performed with thermal conductivity detectors (TCD). The minimum detection limit of the GC was 1.5 pg carbon/s for FID, 2.5 ng mL−1 (standard), and 400 pg mL−1 (μ-TCD) for TCD. The product selectivity was calculated using the following equation: Ci (%) = Ci moles of the product/total moles of C produced × 100%. Furthermore, trace amounts of gases that are difficult to analyze by GC were analyzed by using a mass spectrometer (BELMASS, BEL Japan Inc., Japan).
When the catalyst is repeatedly done to run-off rather than continuously running the reaction for a long time, the catalyst structure is more damaged and so the activity of the catalyst is lowered. That is, more important than the lifetime of the catalyst is to confirm the recycling of the catalyst. Generally, in the thermal catalyst system, a catalyst used in the reaction is calcined at a high temperature, and a regeneration method for removing carbon coking accumulated on the catalyst surface is performed. However, a general regeneration method is not standardized in the photocatalyst system. Therefore, some researchers, including us, choose their own special regeneration method. In this study, the recycling experiment was carried out in the following manner: After each run of photocatalysis reactions for 10 h, the light was switched off, the reactor was disassembled, and then the photocatalyst was separated from reactor. The used catalyst was washed with distilled water for three times, filtered and dried at 60 °C. It was then thermal-treated for 2 h under N2 at 150 °C. The catalyst was then stored in the dark for 24 h to minimize contact with external light until it was repacked into the photoreactor. This is to ensure that all regenerated catalysts are in the same environment as much as possible. The reactor was again installed, and then the gas inside the reactor was ejected. The stored catalyst was again fixed in the reactor, and the reactor was purged with CO2 to displace the air inside the reactor. At this time, the reactor valve is closed/opened and repeated three times while purging CO2 into the reactor. After ejecting all the air gas inside the reactor, both valves of the reactor are closed, water was injected in, vaporized it, and the reaction was initiated by switching on the light source. This experiment was repeated five times.

4. Conclusions

Bi2S3/MoS2 heterojunction catalysts comprising Bi2S3 rods wrapped with MoS2 nanosheets were prepared via a hydrothermal method. The optimal photocatalytic performance was achieved with the 1Bi2S3/1MoS2 heterojunction, for which CO2 reduction to CO after 10 h was almost 40 and 16 times higher than that achieved with pure MoS2 and Bi2S3, respectively. The novelty of this study is that CO was generated with sufficiently high selectively to preclude the need for secondary gas separation. Additionally, the enhanced photocatalytic efficiency could be attributed to the combination of Bi2S3 and MoS2, which extends the light absorption ability and formation of the Bi2S3/MoS2 heterojunction with a tight face-to-face connection and a good energy band match between MoS2 and Bi2S3. The interfacial interaction between MoS2 and Bi2S3 can accelerate the separation of the photo-induced carriers and increase the sulfur defects.

Author Contributions

Conceptualization, M.K.; Data curation, R.K. and J.Y.D.; Formal analysis, R.K., J.K. and J.Y.D.; Investigation, R.K. and J.K.; Methodology, M.W.S. and J.Y.D.; Writing—original draft, M.K.

Funding

This research was funded by a Yeungnam University research grant 2019 (Grant No. 219A345005).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, Y.; Ajmal, S.; Zheng, X.; Zhang, L. Efficient nanomaterials for harvesting clean fuels from electrochemical and photoelectrochemical CO2 reduction. Sustain. Energy Fuels 2018, 2, 510–537. [Google Scholar] [CrossRef]
  2. Wang, L. Synthetic methods of CuS nanoparticles and their applications for imaging and cancer therapy. RSC Adv. 2016, 6, 82596–82615. [Google Scholar] [CrossRef]
  3. Alcaraz, L.; García-Díaz, I.; González, L.; Rabanal, M.E.; Urbieta, A.; Fernández, P.; López, F.A. New photocatalytic materials obtained from the recycling of alkaline and Zn/C spent batteries. J. Mater. Res. Technol. 2019, 8, 2809–2818. [Google Scholar] [CrossRef]
  4. Liu, B.; Xu, J.; Ran, S.; Wang, Z.; Chen, D.; Shen, G. High-performance photodetectors, photocatalysts, and gas sensors based on polyol reflux synthesized porous ZnO nanosheets. CrystEngComm 2012, 14, 4582–4588. [Google Scholar] [CrossRef]
  5. Yang, J.; Jiang, Y.; Li, L.; Muhire, E.; Gao, M. High-performance photodetectors and enhanced photocatalysts of two-dimensional TiO2 nanosheets under UV light excitation. Nanoscale 2016, 8, 8170–8177. [Google Scholar] [CrossRef] [PubMed]
  6. Chava, R.K.; Do, J.Y.; Kang, M. Enhanced photoexcited carrier separation in CdS–SnS2 heteronanostructures: A new 1D–0D visible-light photocatalytic system for the hydrogen evolution reaction. J. Mater. Chem. A 2019, 7, 13614–13628. [Google Scholar] [CrossRef]
  7. Lee, G.; Wu, J.J. Recent developments in ZnS photocatalysts from synthesis to photocatalytic applications—A review. Powder Technol. 2017, 318, 8–22. [Google Scholar] [CrossRef]
  8. Shown, I.; Samireddi, S.; Chang, Y.; Putikam, R.; Chang, P.; Sabbah, A.; Fu, F.; Chen, W.; Wu, C.; Yu, T.; et al. Carbon-doped SnS2 nanostructure as a high-efficiency solar fuel catalyst under visible light. Nat. Commun. 2018, 9, 169. [Google Scholar] [CrossRef]
  9. Lee, J.H.; Kim, S.; Park, S.; Kang, M. A p-n heterojunction NiS-sensitized TiO2 photocatalytic system for efficient photoreduction of carbon dioxide to methane. Ceram. Int. 2017, 43, 1768–1774. [Google Scholar] [CrossRef]
  10. Kim, J.; Do, J.Y.; Park, N.; Lee, S.J.; Hong, J.; Kang, M. Photoreduction of CO2 into CH4 using Bi2S3-TiO2 double-layered dense films. Korean J. Chem. Eng. 2018, 35, 1089–1098. [Google Scholar] [CrossRef]
  11. Do, J.Y.; Chava, R.K.; Kim, S.K.; Nahm, K.; Park, N.; Hong, J.; Lee, S.J.; Kang, M. Fabrication of core@interface:shell structured CuS@CuInS2:In2S3 particles for highly efficient solar hydrogen production. Appl. Surf. Sci. 2018, 451, 86–98. [Google Scholar] [CrossRef]
  12. Sabarinathan, M.; Harish, S.; Archana, J.; Navaneethan, M.; Ikeda, H.; Hayakawa, Y. Highly efficient visible-light photocatalytic activity of MoS2–TiO2 mixtures hybrid photocatalyst and functional properties. RSC Adv. 2017, 7, 24754–24763. [Google Scholar] [CrossRef]
  13. Li, J.; Wu, N. Semiconductor-based photocatalysts and photoelectrochemical cells for solar fuel generation: A review. Catal. Sci. Technol. 2015, 5, 136–1384. [Google Scholar] [CrossRef]
  14. Xu, Y.; Zhao, W.; Xu, R.; Shi, Y.; Zhang, B. Synthesis of ultrathin CdS nanosheets as efficient visible-light-driven water splitting photocatalysts for hydrogen evolution. Chem. Commun. 2013, 49, 9803–9805. [Google Scholar] [CrossRef]
  15. Kumar, P.; Boukherroub, R.; Shankar, K. Sunlight-driven water-splitting using two-dimensional carbon based semiconductors. J. Mater. Chem. A 2018, 6, 12876–12931. [Google Scholar] [CrossRef]
  16. Weng, B.; Zhang, X.; Zhang, N.; Tang, Z.; Xu, Y. Two-Dimensional MoS2 Nanosheet-Coated Bi2S3 Discoids: Synthesis, Formation Mechanism, and Photocatalytic Application. Langmuir 2015, 31, 4314–4322. [Google Scholar] [CrossRef]
  17. Ke, J.; Liu, J.; Sun, H.; Zhang, H.; Duan, X.; Liang, P.; Li, X.; Tade, M.O.; Liu, S.; Wang, S. Facile assembly of Bi2O3/Bi2S3/MoS2 n-p heterojunction with layered n-Bi2O3 and p-MoS2 for enhanced photocatalytic water oxidation and pollutant degradation. Appl. Catal. B Environ. 2017, 200, 47–55. [Google Scholar] [CrossRef]
  18. Liang, N.; Zai, J.; Xu, M.; Zhu, Q.; Wei, X.; Qian, X. Novel Bi2S3/Bi2O2CO3 heterojunction photocatalysts with enhanced visible light responsive activity and wastewater treatment. J. Mater. Chem. A 2014, 2, 4208–4216. [Google Scholar] [CrossRef]
  19. Tsai, C.; Li, H.; Park, S.; Park, J.; Han, H.S.; Nørskov, J.K.; Zheng, X.; Abild-Pedersen, F. Electrochemical generation of sulfur vacancies in the basal plane of MoS2 for hydrogen evolution. Nat. Commun. 2017, 8, 15113–15121. [Google Scholar] [CrossRef]
  20. Dashtian, K.; Porhemat, S.; Rezvani, A.R.; Ghaedi, M.; Sabzehmeidani, M.M. Adsorption of semisoft pollutants onto Bi2S3/Ag2S-AC under the influence of ultrasonic waves as external filed. J. Ind. Eng. Chem. 2018, 60, 390–400. [Google Scholar] [CrossRef]
  21. Cheah, A.J.; Chiu, W.S.; Khiew, P.S.; Nakajima, H.; Saisopa, T.; Songsiriritthigul, P.; Radiman, S.; Hamid, M.A.A. Facile synthesis of a Ag/MoS2 nanocomposite photocatalyst for enhanced visible-light driven hydrogen gas evolution. Catal. Sci. Technol. 2015, 5, 4133–4143. [Google Scholar] [CrossRef]
  22. Wang, F.Z.; Zheng, M.J.; Zhang, B.; Zhu, C.Q.; Li, L.; Ma, L.; Shen, W.Z. Ammonia intercalated flower-like MoS2 nanosheet film as electrocatalyst for high efficient and stable hydrogen evolution. Sci. Rep. 2016, 6, 31092. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, S.; Yu, X.; Yu, H.; Chen, Y.; Gao, P.; Li, C.; Zhu, C. Growth of Ultrathin MoS2 Nanosheets with Expanded Spacing of (002) Plane on Carbon Nanotubes for High-Performance Sodium-Ion Battery Anodes. ACS Appl. Mater. Interfaces 2014, 6, 21880–21885. [Google Scholar] [CrossRef] [PubMed]
  24. Holzwarth, U.; Gibson, N. The Scherrer equation versus the ‘Debye-Scherrer equation’. Nat. Nanotechnol. 2011, 6, 534–1534. [Google Scholar] [CrossRef] [PubMed]
  25. Gao, D.; Si, M.; Li, J.; Zhang, J.; Zhang, Z.; Yang, Z.; Xue, D. Ferromagnetism in freestanding MoS2 nanosheets. Nanoscale Res. Lett. 2013, 8, 129. [Google Scholar] [CrossRef] [PubMed]
  26. Zhao, L.; Jia, J.; Yang, Z.; Yu, J.; Wang, A.; Sang, Y.; Zhou, W.; Liu, H. One-step synthesis of CdS nanoparticles/MoS2 nanosheets heterostructure on porous molybdenum sheet for enhanced photocatalytic H2 evolution. Appl. Catal. B Environ. 2017, 210, 290–296. [Google Scholar] [CrossRef]
  27. Kondekar, N.P.; Boebinger, M.G.; Woods, E.V.; McDowell, M.T. In situ XPS investigation of transformations at crystallographically oriented MoS2 interfaces. ACS Appl. Mater. Interfaces 2017, 9, 32394–32404. [Google Scholar] [CrossRef]
  28. Kwok, K.M.; Ong, S.W.D.; Chen, L.; Zeng, H.C. Constrained growth of MoS2 nanosheets within a mesoporous silica shell and its effects on defect sites and catalyst stability for H2S decomposition. ACS Catal. 2018, 8, 714–724. [Google Scholar] [CrossRef]
  29. Bhoi, Y.P.; Behera, C.; Majhi, D.; Equeenuddin, S.M.; Mishra, B.G. Visible light-assisted photocatalytic mineralization of diuron pesticide using novel type II CuS/Bi2W2O9 heterojunctions with a hierarchical microspherical structure. New J. Chem. 2018, 42, 281–292. [Google Scholar] [CrossRef]
  30. Ji, Y.; Luo, Y. Theoretical study on the mechanism of photoreduction of CO2 to CH4 on the anatase TiO2(101) surface. ACS Catal. 2016, 6, 2018–2025. [Google Scholar] [CrossRef]
  31. Patil, S.A.; Hwang, Y.; Jadhav, V.V.; Kim, H.; Kim, K.H. Solution processed growth and photoelectrochemistry of Bi2S3 nanorods thin film. J. Photochem. Photobiol. A 2017, 332, 174–181. [Google Scholar] [CrossRef]
  32. D’Amato, C.A.; Giovannetti, R.; Zannotti, M.; Rommozzi, E.; Ferraro, S.; Seghetti, C.; Minicucci, M.; Gunnella, R.; Cicco, A.D. Enhancement of visible-light photoactivity by polypropylene coated plasmonic Au/TiO2 for dye degradation in water solution. Appl. Surf. Sci. 2018, 441, 575–587. [Google Scholar] [CrossRef]
  33. Do, J.Y.; Chava, R.K.; Mandari, K.K.; Park, N.; Ryu, H.; Seo, M.W.; Lee, D.; Senthil, T.S.; Kang, M. Selective methane production from visible-light-driven photocatalytic carbon dioxide reduction using the surface plasmon resonance effect of superfine silver nanoparticles anchored on lithium titanium dioxide nanocubes (Ag@LixTiO2). Appl. Catal. B Environ. 2018, 237, 895–910. [Google Scholar] [CrossRef]
  34. Park, M.; Kwak, B.S.; Jo, S.W.; Kang, M. Effective CH4 production from CO2 photoreduction using TiO2/xmol% Cu–TiO2 double-layered films. Energy Convers. Manag. 2015, 103, 431–438. [Google Scholar] [CrossRef]
  35. Barakat, N.A.M.; Kanjwal, M.A.; Chronakis, I.S.; Kim, H.Y. Influence of temperature on the photodegradation process using Ag-doped TiO2 nanostructures: Negative impact with the nanofibers. J. Mol. Catal. A Chem. 2013, 366, 333–340. [Google Scholar] [CrossRef] [Green Version]
  36. Zhang, Q.H.; Han, W.D.; Hong, Y.J.; Yu, J.G. Photocatalytic reduction of CO2 with H2O on Pt-loaded TiO2 catalyst. Catal. Today 2009, 148, 335–340. [Google Scholar] [CrossRef]
  37. Nogueira, A.E.; Oliveira, J.A.; Silva, G.T.S.T.; Ribeiro, C. Insights into the role of CuO in the CO2 photoreduction process. Sci. Rep. 2019, 9, 1316. [Google Scholar] [CrossRef] [Green Version]
Figure 1. XRD patterns of all of catalysts.
Figure 1. XRD patterns of all of catalysts.
Catalysts 09 00998 g001
Figure 2. TEM images of all of catalysts.
Figure 2. TEM images of all of catalysts.
Catalysts 09 00998 g002
Figure 3. HRTEM image (A) of 1Bi2S3/1MoS2 catalyst, selected area electron diffraction (SAED) pattern (B), and element mapping images (C) of each element.
Figure 3. HRTEM image (A) of 1Bi2S3/1MoS2 catalyst, selected area electron diffraction (SAED) pattern (B), and element mapping images (C) of each element.
Catalysts 09 00998 g003
Figure 4. EDS patterns for all of catalysts and their compositional ratios.
Figure 4. EDS patterns for all of catalysts and their compositional ratios.
Catalysts 09 00998 g004
Figure 5. XPS curves at full range (A) and binding energy curves (B) for Mo 3d, Bi 4f, and S 2s.
Figure 5. XPS curves at full range (A) and binding energy curves (B) for Mo 3d, Bi 4f, and S 2s.
Catalysts 09 00998 g005
Figure 6. Product distributions observed in 13C-mass spectrum during CO2 photoconversion over Bi2S3/MoS2 heterojunction catalyst.
Figure 6. Product distributions observed in 13C-mass spectrum during CO2 photoconversion over Bi2S3/MoS2 heterojunction catalyst.
Catalysts 09 00998 g006
Figure 7. Carbon dioxide photoconversion in the presence of catalysts under visible-light (A) and recycling test for Bi2S3/MoS2 heterojunction catalyst (B).
Figure 7. Carbon dioxide photoconversion in the presence of catalysts under visible-light (A) and recycling test for Bi2S3/MoS2 heterojunction catalyst (B).
Catalysts 09 00998 g007
Figure 8. DR (Diffuse Reflectance)-UV-Visible spectra (A) and Tauc plots (B) for all catalysts. Differential reflectance spectroscopy (DRS).
Figure 8. DR (Diffuse Reflectance)-UV-Visible spectra (A) and Tauc plots (B) for all catalysts. Differential reflectance spectroscopy (DRS).
Catalysts 09 00998 g008
Figure 9. Valence band (VB) locations recorded by XPS (A) and energy diagrams for pure Bi2S3 and MoS2 particles (B).
Figure 9. Valence band (VB) locations recorded by XPS (A) and energy diagrams for pure Bi2S3 and MoS2 particles (B).
Catalysts 09 00998 g009
Figure 10. Photocurrent density cycles of all catalysts.
Figure 10. Photocurrent density cycles of all catalysts.
Catalysts 09 00998 g010
Figure 11. CO2–TPD curves of Bi2S3, MoS2, and Bi2S3/MoS2 heterojunction catalysts.
Figure 11. CO2–TPD curves of Bi2S3, MoS2, and Bi2S3/MoS2 heterojunction catalysts.
Catalysts 09 00998 g011
Scheme 1. Plausible mechanism for carbon dioxide photo-conversion on Bi2S3/MoS2 heterojunction catalysts. Conduction band (CB).
Scheme 1. Plausible mechanism for carbon dioxide photo-conversion on Bi2S3/MoS2 heterojunction catalysts. Conduction band (CB).
Catalysts 09 00998 sch001
Scheme 2. Preparation of Bi2S3/MoS2 heterojunction catalyst using the hydrothermal method.
Scheme 2. Preparation of Bi2S3/MoS2 heterojunction catalyst using the hydrothermal method.
Catalysts 09 00998 sch002

Share and Cite

MDPI and ACS Style

Kim, R.; Kim, J.; Do, J.Y.; Seo, M.W.; Kang, M. Carbon Dioxide Photoreduction on the Bi2S3/MoS2 Catalyst. Catalysts 2019, 9, 998. https://doi.org/10.3390/catal9120998

AMA Style

Kim R, Kim J, Do JY, Seo MW, Kang M. Carbon Dioxide Photoreduction on the Bi2S3/MoS2 Catalyst. Catalysts. 2019; 9(12):998. https://doi.org/10.3390/catal9120998

Chicago/Turabian Style

Kim, Raeyeong, Junyeong Kim, Jeong Yeon Do, Myung Won Seo, and Misook Kang. 2019. "Carbon Dioxide Photoreduction on the Bi2S3/MoS2 Catalyst" Catalysts 9, no. 12: 998. https://doi.org/10.3390/catal9120998

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