Next Article in Journal
Ebola Virus Encodes Two microRNAs in Huh7-Infected Cells
Next Article in Special Issue
Antibacterial Vancomycin@ZIF-8 Loaded PVA Nanofiber Membrane for Infected Bone Repair
Previous Article in Journal
The Interaction of Anti-DNA Antibodies with DNA: Evidence for Unconventional Binding Mechanisms
Previous Article in Special Issue
Reducing Endogenous Labile Zn May Help to Reduce Smooth Muscle Cell Injury around Vascular Stents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploring the Effects of the Interaction of Carbon and MoS2 Catalyst on CO2 Hydrogenation to Methanol

National Center for International Research on Catalytic Technology, Key Laboratory of Chemical Engineering Process & Technology for High-Efficiency Conversion, College of Heilongjiang Province, School of Chemistry and Material Sciences, Heilongjiang University, Harbin 150080, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(9), 5220; https://doi.org/10.3390/ijms23095220
Submission received: 15 April 2022 / Revised: 2 May 2022 / Accepted: 5 May 2022 / Published: 7 May 2022
(This article belongs to the Collection State-of-the-Art Materials Science in China)

Abstract

:
Hydrogenation of CO2 to form methanol utilizing green hydrogen is a promising route to realizing carbon neutrality. However, the development of catalyst with high activity and selectivity to methanol from the CO2 hydrogenation is still a challenge due to the chemical inertness of CO2 and its characteristics of multi-path conversion. Herein, a series of highly active carbon-confining molybdenum sulfide (MoS2@C) catalysts were prepared by the in-situ pyrolysis method. In comparison with the bulk MoS2 and MoS2/C, the stronger interaction between MoS2 and the carbon layer was clearly generated. Under the optimized reaction conditions, MoS2@C showed better catalytic performance and long-term stability. The MoS2@C catalyst could sustain around 32.4% conversion of CO2 with 94.8% selectivity of MeOH for at least 150 h.

1. Introduction

Carbon dioxide is one of the important greenhouse gases contributing to global warming, glacial melting, sea-level rise, ocean acidification, and hypercapnia [1,2]. While CO2 is also an inexpensive, abundant, sustainable, and renewable C1 resource [3,4], it can be captured and utilized in a rational way to form high value-added chemicals, such as carbonate, methanol, formic acid, olefins, aromatics, and so on [5,6]. CO2 conversion to MeOH is the most direct route for synthesizing oxygenated compounds and has received great interest [7,8,9]. As a feedstock, methanol can be used as a precursor for synthesizing aromatics and low olefins; moreover, it is also considered to be a green hydrogen carrier and is used as a fuel additive and fuel substitute directly [10]. Although the synthesis of MeOH from CO2 and H2 is exothermic, CO2 conversion to MeOH is kinetically limited at low temperatures and thermodynamically limited at high temperatures. Due to a high activation energy barrier for the cleavage of the C-O bonds in CO2, which clearly results in a chief challenge to developing effective catalysts for the synthesis of methanol from CO2 at a low temperature.
Until now, there are numerous endeavors on different catalyst systems to address CO2 hydrogenation to methanol, such as Cu-based catalysts [11,12,13], precious metal catalysts [14,15,16,17,18], In2O3-based catalysts [19,20,21], solid solution catalysts [22,23,24], and so on. Among these catalysts, the Cu/ZnO/Al2O3 catalyst has been used as an industrial catalyst for methanol synthesis from CO2 hydrogenation. Therefore, Cu-based catalysts have been extensively investigated in CO2 hydrogenation to methanol and the Cu-ZnO composite is employed as the active species in more than 60% of related reports [25]. However, Cu-based catalysts showed lower selectivity and poor stability because of the competing reverse water-gas shift reaction and the sintering of the active phase, and it was exacerbated by the hydrophilicity of Al2O3, which could adsorb water generated from the CO2 hydrogenation. Therefore, there was an urgent need to develop highly efficient catalysts for CO2 hydrogenation to methanol.
As a typical two-dimensional lamellar material, MoS2 shows magic physical and chemical properties and is consequently applied in catalysts for hydrogen evolution reactions in electrocatalysis, hydrodesulfurization, and synthesis gas conversions. Early in 1981, Saito and Anderson used MoS2 as a catalyst for CO2 hydrogenation at 350 °C, 1 atm, and H2/CO2 = 3.74 [26], while CO was the sole product due to the water gas shift reaction. Combining the electrical conductivity of graphenes with the catalytic activity of MoS2, a few layers of MoS2 platelets supported on few-layers graphene exhibited high catalytic activity for CO2 hydrogenation, but the major product was methane, frequently with selectivity above 95% and in some cases close to 100% [27]. The catalytic performance of MoS2 for CO2 hydrogenation has been studied by density functional theory (DFT) and calculations suggest that MoS2 could promote the C-O scission of HxCO intermediates, thus explaining the high selectivity of hydrocarbons in the CO2 hydrogenation process by using molybdenum sulfides as a catalyst [28]. Interestingly, MoS2 has been used as a support for a single atom in the hydrogenation of CO2. The main product was methanol [29]. Zeng et al. have reported that the MoS2 supporting isolated Pt monomers favored the conversion of CO2 into methanol, and the selectivity of methanol arrived at 95.4% [30]. These results were thanks to the synergetic interaction between neighboring Pt monomers on MoS2. Recently, Wang et al. [31] found that the sulfur vacancy played a key role in the adsorption and activation of CO2 when the sulfur vacancy-rich MoS2 was used as a catalyst for the hydrogenation of CO2. At 180 °C, the selectivity of methanol was achieved at 94.3% with a 12.5% CO2 conversion at 3000 mL gcat.−1 h−1 and the catalyst exhibited high stability over 3000 h without any deactivation. However, improving the catalytic performance of molybdenum sulfide in CO2 hydrogenation to form methanol is still a challenging topic.
Due to anisotropy, the average slab length and layer stacking were important for describing any catalytic active edge sites of MoS2. Therefore, adjusting the slab length and layer stacking of MoS2 would be an effective strategy to generate more active edge sites. Abundant strategies were designed to develop nano-scaled MoS2 with highly exposed active edge sites to enhance its catalytic activity. In our previous work, MoCS@NSC has been prepared and showed the 97.3% selectivity of MeOH and a 20.0% conversion of CO2 in the CO2 hydrogenation [32]. In this catalyst, nano-sized MoS2 was in-situ generated in the process of preparing nano-sized Mo2C confined in carbon material by the pyrolysis of ionic liquid precursors, but its effects on the CO2 hydrogenation to form methanol were not clear. In the present work, carbon-confining molybdenum sulfide (MoS2@C) was designed and prepared using glucose as a carbon source by the in-situ pyrolysis method. On the one side, the carbon layer coating the surface of MoS2 could improve the adsorption quantity of CO2; on the other side, MoS2 with few layers and little size was prepared by the confinement effect of carbon, which exposed the more active edge sites. As a result, the catalytic performance of MoS2 in CO2 hydrogenation to form methanol would be improved. Moreover, the influence of the interaction between MoS2 and the carbon coating layer on the CO2 hydrogenation was investigated, and the activation and the conversion route of CO2 in the presence of MoS2@C were also discussed by in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS).

2. Experiment

2.1. Materials

Glucose, ammonium molybdate, and thiourea (NH4)6Mo7O24·4H2O, AMT were received from Tianjin Kemiou Chemical Reagent Co., Ltd. (Tianjin, China). CO2 (99.99%) and 5% Ar/95% H2 (99.99%) were obtained from Qing Hua Gas Company Limited (Harbin, China), and all reagents were unused and not further purified.

2.2. Catalyst Preparation

2.2.1. Synthesis of MoS2

The MoS2 was synthesized according to the literature [31] and as follows: ammonium molybdate (1235.9 mg) and thiourea (2283.6 mg) were dissolved in 20 mL of distilled water. Then the resulting solution was placed in the glass evaporating dish at 80 °C for 12 h, and the dried mixture was calcined at 550 °C for 2 h in a nitrogen atmosphere, thus generating MoS2.

2.2.2. Synthesis of MoS2@C

Ammonium molybdate (1235.9 mg), thiourea (2283.6 mg), and glucose (9458.4 mg) (the molar ratio of carbon to molybdenum was 45) were dissolved in 35 mL of distilled water. Then the resulting solution was placed in the glass evaporating dish at 80 °C for 12 h, and the dried mixture was calcined at 550 °C for 2 h in a nitrogen atmosphere. Finally, the MoS2-45@C was obtained.
According to the same processes, the MoS2-5@C sample was prepared when the molar ratio of carbon to molybdenum was reduced to 5 in the precursor.

2.2.3. Synthesis of MoS2/C

MoS2/C catalyst was prepared by the isovolumic impregnation method. Ammonium molybdate (1235.9 mg) and thiourea (2283.6 mg) were dissolved in distilled water. The coconut shell charcoal was added to the solution to achieve a 5% wt. loading. Then the solid was dried at 80 °C and calcined at 550 °C for 2 h under a nitrogen atmosphere. After this process, the MoS2/C was prepared.

2.3. Catalyst Characterization

The crystalline phase of the catalyst was characterized by X-ray diffraction (XRD) on a D8 Advance with an acceleration voltage of 40 kV.
The microstructure of the catalyst was observed by transmission electron microscopy (TEM) on a JEM-2100 with an acceleration voltage of 200 kV.
The electronic properties of the catalyst surface were determined by X-ray photoelectron spectroscopy (XPS) with an ESCALAB 25, monochromatic Al Kα-rays as the X-ray source, and energy of 1486.6 eV.
The CO2-programmed temperature desorption (TPD) was performed. A 0.2 g sample was purged at 500 °C for 60 min under He (40 mL/min). It was naturally cooled to 50 °C and adsorbed for 60 min under CO2 (40 mL/min). The sample was then purged for 30 min in He (40 mL/min) and finally warmed from 50 °C to 400 °C at 10 °C/min in He (40 mL/min) for CO2 desorption.
In-situ diffuse reflectance and infrared Fourier transform spectroscopy (DRIFTS) measurements were carried out on a Frontier spectrometer by PerkinElmer. The sample was placed directly in the in-situ cell with a ZnSe window and pretreated at 400 °C for 60 min with an H2 flow of 20 mL/min, and then the background spectrum of the sample was collected from 500 to 4000 cm−1. The feed gas H2/CO2 (3/l, 60 mL/min H2, 20 mL/min CO2) was introduced into the cell. The in-situ DRIFTS were recorded with a resolution of 4 cm−1 and with an accumulation of 32 scans every 1 min.

2.4. Catalytic Performance Test

The activity measurements for CO2 hydrogenation were performed in a continuous flow high pressure fixed bed reactor (12 mm internal diameter). Prior to the reaction, the catalyst was pretreated in situ for 3 h at 400 °C in pure hydrogen (22 mL/min). After the reactor had cooled to 220 °C, feed gas with an H2/CO2 ratio of 3/l and a pressure of 3.0 MPa was introduced into the reactor. The effluent was quantified using a Tianmei GC-7900 F and a GC-7890-II gas chromatograph equipped with a flame ionization detector and a thermal conductivity detector, respectively.

2.5. Calculation of CO2 Conversion and Product Selectivity

The CO2 conversion was calculated by an internal normalization method, and the following Equations (1)–(5) were used for calculating CO2 conversion and product selectivity.
The CO2 conversion is expressed as Conv . CO 2 and the selectivity of the products CO, CH4, CH3OH and CH3OCH3 is expressed as Sel . CO , Sel . CH 4 , Sel . CH 3 OH   and   Sel . CH 3 OCH 3 respectively (Equations (1)–(5)). A CO 2 , out , A Ar , in , A Ar , out are the peak areas of the CO2 and Ar signals at the inlet and tail, in the following order. f CO 2 and f Ar are the correction factors for CO2 and Ar, in the order of precedence. A C O , o u t   and f C O , in turn, are the tailpipe signal response area and correction factor for CO. n C H 4 , o u t , n C H 3 O H , o u t , n C H 3 O C H 3 , o u t in order, represent the molarity at the tail gas of CH4, CH3OH, and CH3OCH3.   n C O 2 , i n   a n d   n C O 2 , o u t are the order of the moles of CO2 inlet and tail gas.
C o n v . C O 2 = A C O 2 , i n f C O 2 A A r , i n f A r A C O 2 , o u t f C O 2 A A r , o u t f A r A C O 2 , i n f C O 2 A A r , i n f A r × 100 %  
S e l . C O = A C O , o u t f C O A A r , o u t f A r A C O 2 , i n f C O 2 A A r , i n f A r A C O 2 , o u t f C O 2 A A r , o u t f A r × 100 %
S e l . C H 4 = 100 % × n C H 4 , o u t / ( n C O 2 , i n n C O 2 , o u t )
S e l . C H 3 O H = 100 % × n C H 3 O H , o u t / ( n C O 2 , i n n C O 2 , o u t )  
S e l . C H 3 O C H 3 = 100 % × 2 n C H 3 O C H 3 , o u t / ( n C O 2 , i n n C O 2 , o u t )
In addition, the space time yield (STY) of CH3OH was calculated according to the following Equation (6):
S T Y = F M Y V m W
where F is the volumetric flow rates of CO2,   M   is the molecular mass of CH3OH, V m is the molar volume of an ideal gas at standard temperature and pressure (22.414 L/mol), W is the mass of catalyst, and Y is the yield of CH3OH, respectively.

3. Results and Discussions

The crystal phase structure of MoS2 samples was confirmed by the XRD patterns and is shown in Figure 1. The XRD characteristic diffraction peaks of the bulk MoS2 were shown at 14.4°, 32.9°, 39.5°, 49.8°, 58.8°, and 69.2°, which were assigned to (002), (100), (103), (105), (110), and (108) crystalline planes of MoS2 [33]. When MoS2 was supported on the coconut shell charcoal, except for the characteristic diffraction peak at 2θ of 26° attributed to graphitic carbon [34], there was also a weaker diffraction peak at 2θ of 44.2°, which was assigned to the (105) crystallographic plane of MoS2. It was suggested that MoS2 was successfully supported on coconut shell charcoal. When the molar ratio of carbon to molybdenum is 45 in the precursor, the resulting MoS2-45@C only shows a broad diffraction peak at 2θ of 26°, which corresponds to the (002) planar diffraction peak of graphite [34]. While the characteristic diffraction peaks of MoS2 were not apparent, it was due to the high dispersion of MoS2.
When the molar ratio of carbon to molybdenum was decreased to 5 in the precursor, the characteristic diffraction peaks of MoS2-5@C shown at 32.9°, 39.5°, and 58.8° were attributed to the (100), (103), and (110) crystallographic planes of MoS2, respectively. These results confirmed that the carbon, which was formed by in-situ pyrolysis, showed the confinement effect on the synthesis of MoS2.
It also can be seen that all patterns are characterized by broad reflections with low intensities, and these observations are clear hints for the poor crystallinity and sizes of coherent scattering domains within the nano regime. Compared with the bulk MoS2 sample, the intensity of the (002) reflection at 2θ of 14° disappeared for the samples MoS2/C, MoS2-5@C, and MoS2-45@C. This observation can be explained by the decreasing number of stacked MoS2 slabs in the products [35]. These results also indicated the confinement effects of in-situ formation carbon could suppress the growth of MoS2 grains.
To further enlighten the structure of MoS2, transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) analyses were also carried out. From Figure 2a, it can be seen that the bulk MoS2 has more layers on the edge, and the thicker part in the middle is not even visible. The clear lattice striations on the surface clearly show that the crystal plane spacing is 0.62 nm, corresponding to the (002) crystal plane of MoS2. Compared with bulk MoS2, MoS2/C is limited by the pores of the coconut shell charcoal (Figure 2b), and the layer number of MoS2 in MoS2/C was decreased to 7, which corresponded to the results of XRD. These results suggested that loaded on the support was beneficial to reduce the layers of MoS2 [36,37,38]. The image of Figure 2c showed MoS2 with fewer layers and the smaller practical was obtained successfully in MoS2-45@C when the carbon was in situ formed from the pyrolysis of glucose, and it was beneficial to expose the more active site at the edges for the CO2 hydrogenation. The SAED patterns (Figure 2d) demonstrate the low crystallinity of MoS2, which matches the XRD results.
An X-ray photoelectron spectroscopy (XPS) characterization was carried out to further study the element chemical states in MoS2 samples. Figure S1 and Table S1 showed the XPS survey spectra of the samples under study, and the numerical values of the surface composition obtained from these spectra were given in Table S1. The elements of Mo, S, C, and O were detected on the surface of all samples, and the N element was also found on the surface of MoS2/C and MoS2-45@C. Note that the ratio of the Mo and S in bulk MoS2 followed the chemical formula, while the ratio of the Mo and S in MoS2/C and MoS2-45@C was lower than the chemical formula. These results indicate that the carbon material was doped by the S element when the MoS2 was generated in MoS2/C and MoS2-45@C.
Figure 3a and Table S2 show the high-resolution XPS spectrum in the Mo 3d region, and the Mo 3d peaks of the bulk MoS2 sample with binding energies of 232.0 and 229.1 eV are indexed to Mo 3d3/2 and Mo 3d5/2, respectively, indicating the presence of Mo4+ of molybdenum disulfide [39]. In addition, a small peak at 226.1 eV was found and assigned to S 2s [39]. In XPS of MoS2/C, there are two prominent peaks assigned to Mo 3d3/2 and Mo 3d5/2 (232.0 eV and 229.1 eV), which demonstrate the existence of Mo4+ and the successful synthesis of MoS2 [40]. Additionally, the peaks at 232.8 and 235.9 eV can be ascribed to Mo6+ 3d5/2 and 3d3/2, which were formed by the surface oxidation of MoS2 [41,42,43]. Compared with MoS2/C, the characteristic peaks corresponding to Mo4+ and Mo6+ were also found, but the binding energy between the Mo6+ 3d5/2 and 3d3/2 in MoS2-45@C gave a negative shift (0.4 eV), suggesting that the electron interactions between carbon and the MoS2 surface in MoS2-45@C were stronger than in MoS2/C. It is worth mentioning that the two obvious peaks centered at 231.1 and 228.0 eV imply the existence of a C-Mo bond, further confirming the stronger interfacial interaction between the MoS2 surface and the carbon coating layer in MoS2-45@C [44], which could weaken the Mo-S bond [45].
As exhibited in Figure 3b and Table S3, the S 2p spectra of the bulk MoS2 sample showed two strong peaks at 163.2 eV and 161.8 eV for the S2p1/2 and S2p3/2 binding energies of S2− [46,47]. The electron binding energy of S 2p3/2 of S2− in MoS2/C and MoS2-45@C had a negative shift of about 0.3 eV compared with MoS2, respectively. It indicated the existence of electron interactions between the carbon coating layer and the MoS2 surface. Moreover, the doublet peaks at 163.7 and 164.8 eV were found, and they were assigned to S22− and apical S2−, which indicates the formation of sulfur vacancies on the catalyst surface [48]. The sulfur vacancies could induce the charge-density redistribution, thus producing much more active sites on the catalyst surface. Except, so far, the peak with a binding energy of 168.3 eV was detected in the XPS of MoS2/C and MoS2-45@C, which resulted from the surface oxidation of sulfur elements, and it was contributed to the presence of a sulfate group [49].
In addition, three peaks were present at 395.1 eV, 398.2–398.6 eV, and 400.1–400.5 eV in the N1s high-resolution XPS spectrum of MoS2-45@C and MoS2/C (Figure S2 and Table S4), which can be assigned to Mo3p, pyridinic-N, and pyrrolic-N [49]. It was mentioned that the binding energy of pyridinic-N and pyrrolic-N in MoS2-45@C was 0.4 eV lower than that in MoS2/C, demonstrating an increased electron cloud density around nitrogen in MoS2-45@C, which was benefited by the adsorption of CO2 on its surface.
Moreover, observing the C 1s spectra of MoS2-45@C, it was clearly divided into five fitted peaks at 284.5, 285.2, 286.1, 286.6, and 288.6 eV, which can be related to C-C, C-O, C-O-C/S, C-O-Mo, and C=O (Figure S3 and Table S5). It shows that MoS2 could also be tightly linked with carbon by the C-O-Mo bond, being conducive to the MoS2 confined in the carbon coating layer [44]. As shown in Figure S4 and Table S6, the peaks in the O1s spectrum at 533.3 eV, 532.5 eV, 531.7 eV, and 530.8 eV are for the Mo-O, C-OH, C-O/O-C-N, and O=C groups, respectively [50].
Due to the dual-site mechanism for CO2 hydrogenation, the adsorption and the activation of CO2 occur on the surface of the supporter, implying that the activity and conversion of CO2 are closely related to the surface basicity of the supporter [25,51,52,53,54]. The CO2-TPD experiments were carried out, and the results are shown in Figure 4 and Figure S5. The temperature of the CO2 desorption peak was increased with the order bulk MoS2 < MoS2/C < MoS2@C < NSC (N,S-codoping carbon), and the area of the CO2 desorption peak was also raised in the same order, suggesting the basic strength and the number of basic sites were enhanced with the order bulk MoS2 < MoS2/C < MoS2@C < NSC. It was due to the strong acid-base interactions between the basic S-C functional group and the acidic CO2 molecule in the carbon skeleton and the strong dipole-dipole interaction between the large quadrupole moment of the CO2 molecule and the polar sites associated with the sulfur functional group [55].
Under 220 °C, 3 MPa, and a gas hourly space velocity (GHSV) of 5670 mL h−1 gcat.−1, the catalytic performance was evaluated in a fixed-bed reactor, and the results are listed in Table 1. As a reference catalyst, the catalytic performance of NSC was investigated first, and the result showed CO2 could not be converted, suggesting that NSC had no catalytic activity in the CO2 hydrogenation, although it exhibited the highest CO2 adsorption capacity. Employing MoS2 as a catalyst, the selectivity of methanol arrived at 66.9% with the 18.3% conversion of CO2, and methane, as the major byproduct, was found with a 32.7% selectivity. Meanwhile, dimethyl ether (DME) was also detected with a 0.4% selectivity. These results indicated that MoS2 displayed catalytic activity for CO2 hydrogenation, but the selectivity of methanol was lower, and the high selectivity of methane was given. Using MoS2-45@C as an alternative catalyst, the selectivity of methanol was improved to 95.8% with the 27.3% conversion of CO2, the selectivity of methane was reduced to 4.2%, and DME was not detected. Additionally, the STY of methanol arrived at 0.538 ggcat.−1 h−1 in the presence of MoS2-45@C. On the one hand, exposing the more active sites to thin and small MoS2 and the higher carbon dioxide adsorption capacity of MoS2-45@C were beneficial for accelerating the CO2 conversion due to the dual-site mechanism for CO2 hydrogenation, and the complete decomposition of CO2 on the surface of MoS2-45@C was inhibited by the higher carbon dioxide adsorption capacity, which was conducive to improving methanol selectivity [56,57,58]. On the other hand, the stronger interaction of MoS2 and the carbon coating layer in MoS2-45@C was also beneficial to CO2 hydrogenation by decreasing the Gibbs free energy of hydrogen adsorption [43]. Moreover, the high selectivity of methanol was also thanks to the additional S-vacancy in the MoS2-45@C catalyst (Figure S6) [31]. When the molar ratio of carbon to molybdenum was decreased to 5 in the precursor, the MoS2-5@C gave a 79.9% selectivity of methanol with an 18.4% conversion of CO2. These results were probably because the larger size of MoS2 and the less active site were exposed. The compared catalyst, MoS2/C, showed a 78.5% selectivity of methanol, while the conversion was very low (only 4.2%). This was due to the fact that the 5% loading capacity of MoS2 on the coconut shell carbon was too low, which resulted in a large amount of adsorbed CO2 that could not be efficiently converted.
In the presence of MoS2-45@C, the reaction conditions were optimized, and the results are shown in Figure 5. Firstly, the effects of reaction temperature on the catalytic performance were investigated (Figure 5a).
It can be seen that the CO2 conversion was increased by enhancing the reaction temperature, and it was improved from 26.3% to 32.6% by raising the temperature from 140 °C to 240 °C. These results indicated that MoS2-45@C exhibited higher catalytic activity at a low temperature. When the reaction was performed at 140 °C, 89.0% selectivity of methanol was given, and the methane selectivity reached 11.0%. With increasing reaction temperature to 160 °C, the selectivity of methanol was improved to 95.8%. Further raising the reaction temperature to 240 °C, the selectivity of methanol was reduced, and an 82.6% selectivity of methanol was obtained. It was probably because the too high reaction temperature contributed to the full decomposition of CO2 on the catalyst surface and enhanced the catalytic hydrogenation activity, which contributed to the high selectivity of methane. Additively, CO was found when the reaction was performed at 240 °C in the presence of MoS2-45@C.
Controlling the reaction temperature at 160 °C, the effects of pressure on the CO2 hydrogenation were evaluated by using MoS2-45@C as a catalyst (Figure 5b). With increasing the reaction pressure, the CO2 conversion was gradually increased with no significant change in methanol selectivity. These results suggest the higher reaction pressure has a positive effect on the CO2 conversion. It was because the CO2 hydrogenation was a volume reduction reaction. At the same time, the higher reaction pressure was advantageous for the adsorption of CO2 on the surface of MoS2-45@C, which was favorable for the conversion of CO2. In Figure 5b, it is worth noting that the 12.2% conversion of CO2 and the 92.2% selectivity of methanol were given when the reaction pressure was reduced to 1 MPa. These results suggested that MoS2-45@C displayed a highly catalytic performance under the low reaction pressure. Following that, the influence of the ratio of H2 and CO2 in the feed gas on the CO2 hydrogenation was also investigated (Figure 5c). The results showed the CO2 conversion was susceptible to the ratio of H2 and CO2 in the feed gas, while the selectivity of methanol was kept almost unchanged.
Stability is a fatal issue for the catalysts, which were used in CO2 hydrogenation. Under the optimal reaction conditions, the stability of MoCS-45@C was investigated in a fixed-bed reactor, and the results are shown in Figure 6. In the first 20 h, the CO2 conversion, the methanol selectivity, and STY were increased with prolonged reaction time, while the methane selectivity was decreased. When the reaction time was more than 20 h, the catalytic performance of MoCS-45@C was kept stable. The CO2 conversion was stabilized at around 32.4%, with about 94.8% selectivity of methanol. During the reaction period of 150 h, the catalytic performance of MoCS-45@C showed almost no attenuation, which suggests a promising prospect for industrial applications.
To propose a reaction sequence and a surface reaction mechanism, in situ diffuse reflectance infrared Fourier transform spectroscopy (in situ DRIFTS) was used to identify the evolution of surface species on the surface of MoS2-45@C. The in-situ drift spectra for the hydrogenation of CO2 to methanol over MoS2-45@C at 180 °C with time were shown in Figure 7, and detailed information on the evolution of intermediate species could be found.
Firstly, the pure CO2 was introduced into the MoS2-45@C catalyst (Figure 7a), and the IR bands at 1420, 1437, 1458, 1522, 1542, and 1575 cm−1 were observed, which were assigned to adsorbed *CO2 species [59,60,61,62,63]. Moreover, IR bands at 1474 and 1558 cm−1 also came into the formation, assigned to bidentate carbonate [59,64], and the signals of monodentate carbonate species were prevalent at 1490 and 1508 cm−1 [64,65]. Additionally, their intensities were increased by prolonging the contact time of CO2 and MoS2-45@C from 1 min to 8 min. These results indicated that CO2 could be adsorbed on the surface of the MoS2-45@C catalyst, and the adsorption capacity of CO2 was improved by prolonging the contact time of CO2 and MoS2-45@C. It was worth noting that the IR bands at 2078 and 2094 cm−1 were found when the MoS2-45@C catalyst was exposed to the CO2 atmosphere. These results indicated that CO2 was dissociated to yield surface-bound CO* on the catalytic surface [56,58,66], and their intensities were increased as time went on. These results would be beneficial for increasing the selective synthesis of methanol from the CO2 hydrogenation [31]. Then, the feeding gas was switched from pure CO2 to H2 (Figure 7b), and the CO* peaks from the dissociation of CO2 gradually disappeared with the rise of CH3O* peaks (2864 and 2917 cm−1) [62,67], and the intensity of which decreased as time went on, thereby indicating the hydrogenation of CO* to CH3O* and then the formation of CH3OH. At the same time, a weak shoulder peak that appeared at 2957 cm−1 in the ν (CH) region was also detected, and it was a combination of the CH bending and asymmetric OCO stretching modes of formate species (HCOO*) [60,68]. It was indicated the carbonate species adsorbed on the surface of the MoS2-45@C catalyst were also hydrogenated. These results explain the decrease in the peaks intensity of the carbonate species when H2 was introduced. Moreover, the IR bonds at 2850 cm–1 were also found, and they were assigned to the symmetric and asymmetric H2CO* stretching vibrations [60,69], respectively, which might derive from both formate and CO-hydro pathways. These results confirmed that both CO and formate were significant intermediate species for CO2 hydrogenation to methanol over the MoS2-45@C catalyst. Exposing MoS2-45@C to the feed gas (CO2 + H2), similar in-situ drift IR spectra were obtained (Figure 3c), and these results suggested that the HCOO* hydrogenation route and the CO* hydrogenation route were performed simultaneously in the presence of the MoS2-45@C catalyst in the CO2 hydrogenation. Hence, combining the results of in-situ DRIFTS with the literature [56,57,58,59,60,61,62,63,64,65,66,67,68,69], we held the opinion that the hydrogenation of HCOO* and CO* were all carried out when MoS2-45@C was employed as a catalyst in the CO2 hydrogenation (Figure 8).

4. Conclusions

In conclusion, comparing MoS2/C and bulk MoS2 samples, carbon-confining MoS2 in MoS2-45@C samples prepared by the in-situ pyrolysis method had the characteristics of few layers and small size, which were beneficial to exposing the more active sites. The strong interaction between MoS2 and the carbon coating layer in the MoS2-45@C catalyst was formed, which was also favorable to the CO2 hydrogenation by decreasing the Gibbs free energy of hydrogen adsorption. Moreover, the adsorption capacity of CO2 on the MoS2-45@C surface was improved when the carbon coating layer was doped with sulfur and nitrogen, which also contributed to the CO2 conversion and the methanol selectivity. Under the optimal reaction conditions, the MoS2-45@C showed excellent catalytic performance and catalytic stability, and there was no deactivation in CO2 hydrogenation for more than 150 h on stream at least, which indicates a promising potential for industrial applications.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms23095220/s1 [70,71,72,73,74].

Author Contributions

P.C.: Investigation, Formal analysis, Writing—original draft. R.S.: characterization. L.X.: Conceptualization, Methodology, Writing—review and editing. W.W.: Visualization, Writing—review and editing, Supervision, Project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the financial supports from the National Key Research and Development Project, Intergovernmental International Science and Technology Innovation Cooperation Key Project (2018YFE0108800).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors acknowledge the financial supports from the National Key Research and Development Project, Intergovernmental International Science and Technology Innovation Cooperation Key Project (2018YFE0108800).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wei, J.; Yao, R.; Han, Y.; Ge, Q.; Sun, J. Towards the development of the emerging process of CO2 heterogenous hydrogenation into high-value unsaturated heavy hydrocarbons. Chem. Soc. Rev. 2021, 50, 10764–10805. [Google Scholar] [CrossRef] [PubMed]
  2. Jiang, X.; Nie, X.; Guo, X.; Song, C.; Chen, J.G. Recent advances in carbon dioxide hydrogenation to methanol via heterogeneous catalysis. Chem. Rev. 2020, 120, 7984–8034. [Google Scholar] [CrossRef] [PubMed]
  3. Xu, D.; Wang, Y.Q.; Ding, M.Y.; Hong, X.L.; Liu, G.L.; Tsang, S.C.E. Advances in higher alcohol synthesis from CO2 hydrogenation. Chem 2021, 7, 849–881. [Google Scholar] [CrossRef]
  4. Bediako, B.B.A.; Qian, Q.; Han, B.X. Synthesis of C2+ Chemicals from CO2 and H2 via C-C Bond Formation. Acc. Chem. Res. 2021, 54, 2467–2476. [Google Scholar] [CrossRef] [PubMed]
  5. Sancho-Sanz, I.; Korili, S.A.; Gil, A. Catalytic valorization of CO2 by hydrogenation: Current status and future trends. Catal. Rev. 2021. [Google Scholar] [CrossRef]
  6. Wang, D.; Xie, Z.H.; Porosoff, M.D.; Chen, J.G. Recent advances in carbon dioxide hydrogenation to produce olefins and aromatics. Chem 2021, 7, 2277–2311. [Google Scholar] [CrossRef]
  7. Navarro-Jaén, S.; Virginie, M.; Bonin, J.; Robert, M.; Wojcieszak, R.; Khodakov, A.Y. Highlights and challenges in the selective reduction of carbon dioxide to methanol. Nat. Rev. Chem. 2021, 5, 564–579. [Google Scholar] [CrossRef]
  8. Bai, S.T.; De Smet, G.; Liao, Y.H.; Sun, R.Y.; Zhou, C.; Beller, M.; Maes, B.U.W.; Sels, B.F. Homogeneous and heterogeneous catalysts for hydrogenation of CO2 to methanol under mild conditions. Chem. Soc. Rev. 2021, 50, 4259–4298. [Google Scholar] [CrossRef]
  9. Zhang, X.B.; Zhang, G.H.; Song, C.S.; Guo, X.W. Catalytic conversion of carbon dioxide to methanol: Current status and future perspective. Front. Energy Res. 2021, 8, 621119. [Google Scholar] [CrossRef]
  10. Sharma, P.; Sebastian, J.; Ghosh, S.; Creaser, D.; Olsson, L. Recent advances in hydrogenation of CO2 into hydrocarbons via methanol intermediate over heterogeneous catalysts. Catal. Sci. Technol. 2021, 11, 1665–1697. [Google Scholar] [CrossRef]
  11. Murthy, P.S.; Liang, W.B.; Jiang, Y.J.; Huang, J. Cu-Based Nanocatalysts for CO2 hydrogenation to methanol. Energy Fuels 2021, 35, 8558–8584. [Google Scholar] [CrossRef]
  12. Shi, Z.; Tan, Q.; Wu, D. Enhanced CO2 hydrogenation to methanol over TiO2 nanotubes-supported CuO-ZnO-CeO2 catalyst. Appl. Catal. A Gen. 2019, 581, 58–66. [Google Scholar] [CrossRef]
  13. Zhang, G.; Fan, G.; Yang, L.; Li, F. Tuning surface-interface structures of ZrO2 supported copper catalysts by in situ introduction of indium to promote CO2 hydrogenation to methanol. Appl. Catal. A Gen. 2020, 605, 117805. [Google Scholar] [CrossRef]
  14. Snider, J.L.; Streibel, V.; Hubert, M.A.; Choksi, T.S.; Valle, E.; Upham, D.C.; Schumann, J.; Duyar, M.S.; Gallo, A.; Abild-Pedersen, F.; et al. Revealing the synergy between oxide and alloy phases on the performance of bimetallic In-Pd catalysts for CO2 hydrogenation to methanol. ACS Catal. 2019, 9, 3399–3412. [Google Scholar] [CrossRef]
  15. Gholinejad, M.; Khosravi, F.; Afrasi, M.; Sansano, J.M.; Nájera, C. Applications of bimetallic PdCu catalysts. Catal. Sci. Technol. 2021, 11, 2652–2702. [Google Scholar] [CrossRef]
  16. Sha, F.; Han, Z.; Tang, S.; Wang, J.; Li, C. Hydrogenation of carbon dioxide to methanol over non-Cu-based heterogeneous catalysts. ChemSusChem 2020, 13, 6160–6181. [Google Scholar] [CrossRef] [PubMed]
  17. Khobragade, R.; Roškarič, M.; Žerjav, G.; Košiček, M.; Zavašnik, J.; Van de Velde, N.; Jerman, I.; Tušar, N.N.; Pintar, A. Exploring the effect of morphology and surface properties of nanoshaped Pd/CeO2 catalysts on CO2 hydrogenation to methanol. Appl. Catal. A Gen. 2021, 667, 118394. [Google Scholar] [CrossRef]
  18. Lin, F.; Jiang, X.; Boreriboon, N.; Wang, Z.; Song, C.; Cen, K. Effects of supports on bimetallic Pd-Cu catalysts for CO2 hydrogenation to methanol. Appl. Catal. A Gen. 2019, 585, 117210. [Google Scholar] [CrossRef]
  19. Nie, X.W.; Li, W.H.; Jiang, X.; Guo, X.W.; Song, C.S. Recent advances in catalytic CO2 hydrogenation to alcohols and hydrocarbons. Adv. Catal. 2020, 65, 121–233. [Google Scholar]
  20. Shen, C.Y.; Bao, Q.Q.; Xue, W.J.; Sun, K.H.; Zhang, Z.T.; Jia, X.Y.; Mei, D.H.; Liu, C.J. Synergistic effect of the metal-support interaction and interfacial oxygen vacancy for CO2 hydrogenation to methanol over Ni/In2O3 catalyst: A theoretical study. J. Energy Chem. 2022, 65, 623–629. [Google Scholar] [CrossRef]
  21. Wang, J.Y.; Zhang, G.H.; Zhu, J.; Zhang, X.B.; Ding, F.S.; Zhang, A.F.; Guo, X.W.; Song, C.S. CO2 hydrogenation to methanol over In2O3-based catalysts: From mechanism to catalyst development. ACS Catal. 2021, 11, 1406–1423. [Google Scholar] [CrossRef]
  22. Wang, J.J.; Tang, C.Z.; Li, G.N.; Han, Z.; Li, Z.L.; Liu, H.; Cheng, F.; Li, C. High-performance MaZrOx (Ma = Cd, Ga) solid-solution catalysts for CO2 hydrogenation to methanol. ACS Catal. 2019, 9, 10253–10259. [Google Scholar] [CrossRef]
  23. Wang, J.J.; Li, G.N.; Li, Z.L.; Tang, C.Z.; Feng, Z.C.; An, H.Y.; Liu, H.L.; Liu, T.F.; Li, C. A highly selective and stable ZnO-ZrO2 solid solution catalyst for CO2 hydrogenation to methanol. Sci. Adv. 2017, 3, e1701290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Xu, D.; Hong, X.L.; Liu, G.L. Highly dispersed metal doping to ZnZr oxide catalyst for CO2 hydrogenation to methanol: Insight into hydrogen spillover. J. Catal. 2021, 393, 207–214. [Google Scholar] [CrossRef]
  25. Qiao, S.C.; Liu, X.Y.; Zhu, R.R.; Xue, D.M.; Liu, X.Q.; Sun, L.B. Causation of catalytic activity of Cu-ZnO for CO2 hydrogenation to methanol. Chem. Eng. J. 2022, 430, 132784. [Google Scholar] [CrossRef]
  26. Saito, M.; Anderson, R.B. The activity of several molybdenum compounds for the methanation of CO2. J. Catal. 1981, 67, 296–302. [Google Scholar] [CrossRef]
  27. Primo, A.; He, J.; Jurca, B.; Cojocaru, B.; Bucur, C.; Parvulescu, V.I.; Garcia, H. CO2 methanation catalyzed by oriented MoS2 nanoplatelets supported on few layers graphene. Appl. Catal. B Environ. 2019, 245, 351–359. [Google Scholar] [CrossRef]
  28. Liu, P.; Choi, Y.M.; Yang, Y.; White, M.G. Methanol synthesis from H2 and CO2 on a Mo6S8 cluster: A density functional study. J. Phys. Chem. A 2010, 114, 3888–3895. [Google Scholar] [CrossRef]
  29. Lu, Z.; Cheng, Y.; Li, S.; Yang, Z.; Wu, R. CO2 thermoreduction to methanol on the MoS2 supported single Co atom catalyst: A DFT study. Appl. Sur. Sci. 2020, 528, 147047. [Google Scholar] [CrossRef]
  30. Li, H.; Wang, L.; Dai, Y.; Pu, Z.; Lao, Z.; Chen, Y.; Wang, M.; Zheng, X.; Zhu, J.; Zhang, W.; et al. Synergetic interaction between neighbouring platinum monomers in CO2 hydrogenation. Nat. Nanotechnol. 2018, 13, 411–417. [Google Scholar] [CrossRef]
  31. Hu, J.; Yu, L.; Deng, J.; Wang, Y.; Cheng, K.; Ma, C.; Zhang, Q.; Wen, W.; Yu, S.; Pan, Y.; et al. Sulfur vacancy-rich MoS2 as a catalyst for the hydrogenation of CO2 to methanol. Nat. Catal. 2021, 4, 242–250. [Google Scholar] [CrossRef]
  32. Han, H.; Cui, P.P.; Xiao, L.F.; Wu, W. MoCS@NSC with interfacial heterojunction nanostructure: A highly selective catalyst for synthesizing methanol from CO2 at low temperature. J. Environ. Chem. Eng. 2021, 9, 106354. [Google Scholar] [CrossRef]
  33. Panigrahi, P.K.; Pathak, A. Aqueous medium synthesis route for randomly stacked molybdenum disulfide. J. Nanoparticl. 2013, 2013, 671214. [Google Scholar] [CrossRef] [Green Version]
  34. Xia, Y.; Zhu, Y.; Tang, Y. Preparation of sulfur-doped microporous carbons for the storage of hydrogen and carbon dioxide. Carbon 2012, 50, 5543–5553. [Google Scholar] [CrossRef]
  35. Djamil, J.; Segler, S.A.; Dabrowski, A.; Bensch, W.; Lotnyk, A.; Schürmann, U.; Kienle, L.; Hansen, S.; Bewerie, T. The influence of carbon content on the structure and properties of MoSxCy photocatalysts for light-driven hydrogen generation. Dalton Trans. 2013, 42, 1287–1292. [Google Scholar] [CrossRef]
  36. Tsai, C.; Abild-Pedersen, F.; Norskov, J.K. Tuning the MoS2 edge-site activity for hydrogen evolution via support interactions. Nano Lett. 2014, 14, 1381–1387. [Google Scholar] [CrossRef]
  37. Hinnemann, B.; Moses, P.G.; Bonde, J.; Jørgensen, K.P.; Nielsen, J.H.; Horch, S.; Chorkendorff, I.; Nørskov, J.K. Biomimetic hydrogen evolution: MoS2 nanoparticles as catalyst for hydrogen evolution. J. Am. Chem. Soc. 2005, 127, 5308–5309. [Google Scholar] [CrossRef]
  38. Prins, R.; De Beer, V.H.J.; Somorjai, G.A. Structure and function of the catalyst and the promoter in Co-Mo Hydrodesulfurization catalysts. Catal. Rev. 1989, 31, 1–41. [Google Scholar] [CrossRef]
  39. Gao, S.; Yang, L.; Shao, J.; Qu, Q.; Wu, Y.; Holze, R. Construction of hierarchical hollow MoS2/carbon microspheres for enhanced lithium storage performance. J. Electrochem. Soc. 2020, 167, 100525. [Google Scholar] [CrossRef]
  40. Cao, H.; Wang, H.; Huang, Y.; Sun, Y.; Shi, S.; Tang, M. Quantification of gold(III) in solution and with a test stripe via the quenching of the fluorescence of molybdenum disulfide quantum dots. Microchimi. Acta 2016, 184, 91–100. [Google Scholar] [CrossRef]
  41. Zhang, J.; Cui, P.; Gu, Y.; Wu, D.; Tao, S.; Qian, B.; Chu, W.; Song, L. Encapsulating carbon-coated MoS2 nanosheets within a nitrogen-doped graphene network for high-performance potassium-ion storage. Adv. Mater. Interfaces 2019, 6, 1901066. [Google Scholar] [CrossRef]
  42. Zhang, X.; Ma, T.; Fang, T.; Gao, Y.; Gao, S.; Wang, W.; Liao, L. A novel MoS2@C framework architecture composites with three-dimensional cross-linked porous carbon supporting MoS2 nanosheets for sodium storage. J. Alloy. Compd. 2020, 818, 152821. [Google Scholar] [CrossRef]
  43. Cai, Y.; Yang, H.; Zhou, J.; Luo, Z.; Fang, G.; Liu, S.; Pan, A.; Liang, S. Nitrogen-doped hollow MoS2/C nanospheres as anode for long-life sodium-ion batteries. Chem. Eng. J. 2017, 327, 522–529. [Google Scholar] [CrossRef]
  44. Hu, R.; Fang, Y.; Zhu, K.; Yang, X.; Yin, J.; Ye, K.; Yan, J.; Cao, D.; Wang, G. Carbon coated MoS2 Hierarchical Microspheres enabling fast and durable potassium ion storage. Appl. Sur. Sci. 2021, 564, 150387. [Google Scholar] [CrossRef]
  45. Zhao, Z.; Qin, F.; Kasiraju, S.; Xie, L.; Alam, M.K.; Chen, S.; Wang, D.; Ren, Z.; Wang, Z.; Grabow, L.C.; et al. Vertically aligned MoS2/Mo2C hybrid nanosheets grown on carbon paper for efficient electrocatalytic hydrogen evolution. ACS Catal. 2017, 7, 7312–7318. [Google Scholar] [CrossRef]
  46. Lin, Q.; Dong, X.; Wang, Y.; Zheng, N.; Zhao, Y.; Xu, W.; Ding, T. Molybdenum disulfide with enlarged interlayer spacing decorated on reduced graphene oxide for efficient electrocatalytic hydrogen evolution. J. Mater. Sci. 2020, 55, 6637–6647. [Google Scholar] [CrossRef]
  47. Yin, X.; Sun, G.; Song, A.; Wang, L.; Wang, Y.; Dong, H.; Shao, G. A novel structure of Ni-(MoS2/GO) composite coatings deposited on Ni foam under supergravity field as efficient hydrogen evolution reaction catalysts in alkaline solution. Electrochimi. Acta 2017, 249, 52–63. [Google Scholar] [CrossRef]
  48. Yang, S.; Wang, Y.; Zhang, H.; Zhang, Y.; Liu, L.; Fang, L.; Yang, X.; Gu, X.; Wang, Y. Unique three-dimensional Mo2C@MoS2 heterojunction nanostructure with S vacancies as outstanding all-pH range electrocatalyst for hydrogen evolution. J. Catal. 2019, 371, 20–26. [Google Scholar] [CrossRef]
  49. Tang, Y.; Wang, Y.; Wang, X.; Li, S.; Huang, W.; Dong, L.; Liu, C.; Li, Y.; Lan, Y. Molybdenum disulfide/nitrogen-doped reduced graphene oxide nanocomposite with enlarged interlayer spacing for electrocatalytic hydrogen evolution. Adv. Energy Mater. 2016, 6, 1600116. [Google Scholar] [CrossRef]
  50. Liu, H.; Hu, H.; Wang, J.; Niehoff, P.; He, X.; Paillard, E.; Eder, D.; Winter, M.; Li, J. Hierarchical ternary MoO2/MoS2/heteroatom-doped carbon hybrid materials for high-performance lithium-ion storage. ChemElectroChem 2016, 3, 922–932. [Google Scholar] [CrossRef]
  51. Ma, Y.; Wang, J.; Goodman, K.R.; Head, A.R.; Tong, X.; Stacchiola, D.J.; White, M.G. Reactivity of a zirconia-copper inverse catalyst for CO2 hydrogenation. J. Phys. Chem. C 2020, 124, 22158–22172. [Google Scholar] [CrossRef]
  52. Gao, P.; Li, F.; Zhan, H.J.; Zhao, N.; Xiao, F.K.; Wei, W.; Zhong, L.S.; Wang, H.; Sun, Y.H. Influence of Zr on the performance of Cu/Zn/Al/Zr catalysts via hydrotalcite-like precursors for CO2 hydrogenation to methanol. J. Catal. 2013, 298, 51–60. [Google Scholar] [CrossRef]
  53. Xiao, J.; Mao, D.S.; Guo, X.M.; Yu, J. Effect of TiO2, ZrO2, and TiO2-ZrO2 on the performance of CuO-ZnO catalyst for CO2 hydrogenation to methanol. Appl. Surf. Sci. 2015, 338, 146–153. [Google Scholar] [CrossRef]
  54. Liu, T.; Xu, D.; Wu, D.; Liu, G.; Hong, X. Spinel ZnFe2O4 regulates copper sites for CO2 hydrogenation to methanol. ACS Sustain. Chem. Eng. 2021, 9, 4033–4041. [Google Scholar] [CrossRef]
  55. Kiciński, W.; Szala, M.; Bystrzejewski, M. Sulfur-doped porous carbons: Synthesis and applications. Carbon 2014, 68, 1–32. [Google Scholar] [CrossRef]
  56. Posada-Pérez, S.; Ramírez, P.J.; Gutiérrez, R.A.; Stacchiola, D.J.; Viñes, F.; Liu, P.; Illas, F.; Rodriguez, J.A. The conversion of CO2 to methanol on orthorhombic β-Mo2C and Cu/β-Mo2C catalysts: Mechanism for admetal induced change in the selectivity and activity. Catal. Sci. Technol. 2016, 6, 6766–6777. [Google Scholar] [CrossRef]
  57. Posada-Pérez, S.; Viñes, F.; Rodriguez, J.A.; Illas, F. Fundamentals of methanol synthesis on metal carbide based catalysts: Activation of CO2 and H2. Top. Catal. 2015, 58, 159–173. [Google Scholar] [CrossRef]
  58. Posada-Perez, S.; Politi, J.R.d.S.; Vines, F.; Illas, F. Methane capture at room temperature: Adsorption on cubic δ-MoC and orthorhombic β-Mo2C molybdenum carbide (001) surfaces. RSC Adv. 2015, 5, 33737–33746. [Google Scholar] [CrossRef]
  59. Alves, L.M.N.C.; Almeida, M.P.; Ayala, M.; Watson, C.D.; Jacobs, G.; Rabelo-Neto, R.C.; Noronha, F.B.; Mattos, L.V. CO2 methanation over metal catalysts supported on ZrO2: Effect of the nature of the metallic phase on catalytic performance. Chem. Eng. Sci. 2021, 239, 116604. [Google Scholar] [CrossRef]
  60. Yu, J.; Yang, M.; Zhang, J.; Ge, Q.; Zimina, A.; Pruessmann, T.; Zheng, L.; Grunwaldt, J.D.; Sun, J. Stabilizing Cu+ in Cu/SiO2 catalysts with a shattuckite-like structure boosts CO2 hydrogenation into methanol. ACS Catal. 2020, 10, 14694–14706. [Google Scholar] [CrossRef]
  61. Zhao, K.; Wang, L.; Moioli, E.; Calizzi, M.; Züttel, A. Identifying reaction species by evolutionary fitting and kinetic analysis: An example of CO2 hydrogenation in DRIFTS. J. Phys. Chem. C 2021, 123, 8785–8792. [Google Scholar] [CrossRef]
  62. Fisher, I.A.; Bell, A.T. In situ infrared study of methanol synthesis from H2/CO over Cu/SiO2 and Cu/ZrO2/SiO2. J. Catal. 1998, 178, 153–173. [Google Scholar] [CrossRef]
  63. Fisher, I.A.; Bell, A.T. In-situ infrared study of methanol synthesis from H2/CO2 over Cu/SiO2 and Cu/ZrO2/SiO2. J. Catal. 1997, 172, 222–237. [Google Scholar] [CrossRef]
  64. Xu, M.; Yao, S.; Rao, D.; Niu, Y.; Liu, N.; Peng, M.; Zhai, P.; Man, Y.; Zheng, L.; Wang, B.; et al. Insights into interfacial synergistic catalysis over Ni@TiO2−x catalyst toward Water-Gas shift reaction. J. Am. Chem. Soc. 2018, 140, 11241–11251. [Google Scholar] [CrossRef]
  65. Cerdá-Moreno, C.; Chica, A.; Keller, S.; Rautenberg, C.; Bentrup, U. Ni-sepiolite and Ni-todorokite as efficient CO2 methanation catalysts: Mechanistic insight by operando DRIFTS. Appl. Catal. B Environ. 2020, 264, 118546. [Google Scholar] [CrossRef]
  66. Malik, A.S.; Zaman, S.F.; Al-Zahrani, A.A.; Daous, M.A.; Driss, H.; Petrov, L.A. Selective hydrogenation of CO2 to CH3OH and in-depth DRIFT analysis for PdZn/ZrO2 and CaPdZn/ZrO2 catalysts. Catal. Today 2020, 357, 573–582. [Google Scholar] [CrossRef]
  67. Luo, L.; Wang, M.; Cui, Y.; Chen, Z.; Wu, J.; Cao, Y.; Luo, J.; Dai, Y.; Li, W.; Bao, J.; et al. Surface iron species in palladium-iron intermetallic nanocrystals that promote and stabilize CO2 methanation. Angew. Chem. Int. Ed. 2020, 59, 14434–14442. [Google Scholar] [CrossRef]
  68. Kattel, S.; Yan, B.; Yang, Y.; Chen, J.G.; Liu, P. Optimizing binding energies of key intermediates for CO2 hydrogenation to methanol over oxide-supported copper. J. Am. Chem. Soc. 2016, 138, 12440–12450. [Google Scholar] [CrossRef]
  69. Romero-Sarria, F.; Bobadilla, L.F.; Barrera, E.M.J.; Odriozola, J.A. Experimental evidence of HCO species as intermediate in the fischer tropsch reaction using operando techniques. Appl. Catal. B Environ. 2020, 272, 119032. [Google Scholar] [CrossRef]
  70. Wang, X.; Zhang, Y.; Si, H.; Zhang, Q.; Wu, J.; Gao, L.; Wei, X.; Sun, Y.; Liao, Q.; Zhang, Z.; et al. Single-atom vacancy defect to trigger high-efficiency hydrogen evolution of MoS2. J. Am. Chem. Soc. 2020, 142, 4298–4308. [Google Scholar] [CrossRef]
  71. Mu, X.; Zhu, Y.; Gu, X.; Dai, S.; Mao, Q.; Bao, L.; Li, W.; Liu, S.; Bao, J.; Mu, S. Awakening the oxygen evolution activity of MoS2 by exophilic-metal induced surface reorganization engineering. J. Energy Chem. 2021, 62, 546–551. [Google Scholar] [CrossRef]
  72. Gong, F.; Ye, S.; Liu, M.; Zhang, J.; Gong, L.; Zeng, G.; Meng, E.; Su, P.; Xie, K.; Zhang, Y.; et al. Boosting electrochemical oxygen evolution over yolk-shell structured O-MoS2 nanoreactors with sulfur vacancy and decorated Pt nanoparticles. Nano Energy 2020, 78, 105284. [Google Scholar] [CrossRef]
  73. Liu, G.; Robertson, A.W.; Li, M.M.; Kuo, W.C.H.; Darby, M.T.; Muhieddine, M.H.; Lin, Y.C.; Suenaga, K.; Stamatakis, M.; Warner, J.H.; et al. MoS2 monolayer catalyst doped with isolated Co atoms for the hydrodeoxygenation reaction. Nat. Chem. 2017, 9, 810–816. [Google Scholar] [CrossRef]
  74. Cai, L.; He, J.; Liu, Q.; Yao, T.; Chen, L.; Yan, W.; Hu, F.; Jiang, Y.; Zhao, Y.; Hu, T.; et al. Vacancy-induced ferromagnetism of MoS2 nanosheets. J. Am. Chem. Soc. 2015, 137, 2622–2627. [Google Scholar] [CrossRef] [PubMed]
Figure 1. XRD patterns of MoS2 samples.
Figure 1. XRD patterns of MoS2 samples.
Ijms 23 05220 g001
Figure 2. Catalyst characterization: (a) MoS2 TEM (20 nm); (b) MoS2/C TEM (20 nm); (c) MoS2-45@C HRTEM (10 nm); (d) MoS2-45@C SAED (5 l/nm).
Figure 2. Catalyst characterization: (a) MoS2 TEM (20 nm); (b) MoS2/C TEM (20 nm); (c) MoS2-45@C HRTEM (10 nm); (d) MoS2-45@C SAED (5 l/nm).
Ijms 23 05220 g002aIjms 23 05220 g002b
Figure 3. XPS of different catalyst: (a) Mo3d and (b) S2p.
Figure 3. XPS of different catalyst: (a) Mo3d and (b) S2p.
Ijms 23 05220 g003
Figure 4. CO2-TPD curves of MoS2-45@C, MoS2/C and MoS2.
Figure 4. CO2-TPD curves of MoS2-45@C, MoS2/C and MoS2.
Ijms 23 05220 g004
Figure 5. Optimized reaction conditions: (a) reaction temperatures, (b) reaction pressures, (c) ratio of H2 to CO2.
Figure 5. Optimized reaction conditions: (a) reaction temperatures, (b) reaction pressures, (c) ratio of H2 to CO2.
Ijms 23 05220 g005
Figure 6. Stability of the MoS2-45@C catalyst with granule stacking.
Figure 6. Stability of the MoS2-45@C catalyst with granule stacking.
Ijms 23 05220 g006
Figure 7. In situ DRIFTS spectra of CO2 hydrogeantion on MoS2-45@C: different feed introduced to catalyst (a) CO2; (b) H2; (c) CO2 + H2.
Figure 7. In situ DRIFTS spectra of CO2 hydrogeantion on MoS2-45@C: different feed introduced to catalyst (a) CO2; (b) H2; (c) CO2 + H2.
Ijms 23 05220 g007
Figure 8. Reaction route for CO2 hydrogenation to methanol over MoS2-45@C.
Figure 8. Reaction route for CO2 hydrogenation to methanol over MoS2-45@C.
Ijms 23 05220 g008
Table 1. The performance of catalyst in CO2 hydrogeantion [a].
Table 1. The performance of catalyst in CO2 hydrogeantion [a].
CatalystConversion/%Selectivity/%STY
CH3OHCH4CH3OCH3/gMeOH gcat.−1 h−1
NSC-----
MoS218.366.932.70.40.252
MoS2-45@C27.395.84.200.538
MoS2-5@C 18.479.920.100.302
MoS2/C4.278.521.500.068
[a] Reaction conditions: 160 °C, 3 MPa, GHSV 5670 mL gcat.−1 h−1, VH2/VCO2 = 3.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cui, P.; Sun, R.; Xiao, L.; Wu, W. Exploring the Effects of the Interaction of Carbon and MoS2 Catalyst on CO2 Hydrogenation to Methanol. Int. J. Mol. Sci. 2022, 23, 5220. https://doi.org/10.3390/ijms23095220

AMA Style

Cui P, Sun R, Xiao L, Wu W. Exploring the Effects of the Interaction of Carbon and MoS2 Catalyst on CO2 Hydrogenation to Methanol. International Journal of Molecular Sciences. 2022; 23(9):5220. https://doi.org/10.3390/ijms23095220

Chicago/Turabian Style

Cui, Pingping, Ruyu Sun, Linfei Xiao, and Wei Wu. 2022. "Exploring the Effects of the Interaction of Carbon and MoS2 Catalyst on CO2 Hydrogenation to Methanol" International Journal of Molecular Sciences 23, no. 9: 5220. https://doi.org/10.3390/ijms23095220

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