Next Article in Journal
Exploring Nitrogen-Functionalized Graphene Composites for Urinary Catheter Applications
Previous Article in Journal
Nanodiamond Decorated PEO Oxide Coatings on NiTi Alloy
Previous Article in Special Issue
Preparation of Ce-MnOx Composite Oxides via Coprecipitation and Their Catalytic Performance for CO Oxidation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Carbonization Conditions on Structural and Surface Properties of K-Doped Mo2C Catalysts for the Synthesis of Methyl Mercaptan from CO/H2/H2S

1
Faculty of Environmental Science and Engineering, Kunming University of Science and Technology, Kunming 650500, China
2
Xishuangbanna Prefecture Comprehensive Inspection Center of Quality and Technical Supervision, Jinghong 666100, China
3
Faculty of Chemical Engineering, Kunming University of Science and Technology, Kunming 650500, China
4
Yunnan Research Academy of Eco-Environmental Sciences, Kunming 650093, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(18), 2602; https://doi.org/10.3390/nano13182602
Submission received: 1 August 2023 / Revised: 17 August 2023 / Accepted: 22 August 2023 / Published: 21 September 2023
(This article belongs to the Special Issue Nanocatalysts for Air Purification)

Abstract

:
The cooperative transition of sulfur-containing pollutants of H2S/CO/H2 to the high-value chemical methyl mercaptan (CH3SH) is catalyzed by Mo-based catalysts and has good application prospects. Herein, a series of Al2O3-supported molybdenum carbide catalysts with K doping (denoted herein as K-Mo2C/Al2O3) are fabricated by the impregnation method, with the carbonization process occurring under different atmospheres and different temperatures between 400 and 600 °C. The CH4-K-Mo2C/Al2O3 catalyst carbonized by CH4/H2 at 500 °C displays unprecedented performance in the synthesis of CH3SH from CO/H2S/H2, with 66.1% selectivity and a 0.2990 g·gcat−1·h−1 formation rate of CH3SH at 325 °C. H2 temperature-programmed reduction, temperature-programmed desorption, X-ray diffraction and Raman and BET analyses reveal that the CH4-K-Mo2C/Al2O3 catalyst contains more Mo coordinatively unsaturated surface sites that are responsible for promoting the adsorption of reactants and the desorption of intermediate products, thereby improving the selectivity towards and production of CH3SH. This study systematically investigates the effects of catalyst carbonization and passivation conditions on catalyst activity, conclusively demonstrating that Mo2C-based catalyst systems can be highly selective for producing CH3SH from CO/H2S/H2.

Graphical Abstract

1. Introduction

Coal is the main energy source in China. This produces large quantities of sulfur-containing pollutants emitted by coal combustion, which has resulted in serious atmospheric environmental pollution [1,2,3]. Therefore, the removal and utilization of hydrogen sulfide resources have become research hotspots in the coal chemical industry [4,5]. In this industry, the Claus method and the improved Claus method are generally used to recover H2S to produce sulfur or sulfuric acid, and the added value of the products is low. The preparation of the high-value chemical methyl mercaptan (CH3SH) from H2S, along with CO and H2 in the process of coal gasification, has become a valuable development direction [6,7,8,9]. This technology can not only effectively remove H2S but also provide a new route for downstream product development. The key to realizing the efficient utilization of raw materials and the high yield of products is to build a high-performance catalyst system. Until now, molybdenum sulfide based catalysts have been widely preferred because of their excellent sulfur resistance, but the reactant conversion and product selectivity are poor [10,11], which greatly restricts the development and application of molybdenum sulfide based catalysts.
In the industry, CH3SH is mainly synthesized from the thiolation of methanol over alkali-promoted transition metal catalysts [12,13], but methanol needs to be prepared with syngas (CO + H2), which leads to complex processes, massive energy consumption and high economic costs. Olin et al. [14] developed a one-step synthesis method of CH3SH using syngas raw materials and reduced sulfur species (hydrogen sulfide) generated during coal gasification that will greatly reduce the production cost of CH3SH. Therefore, the preparation of CH3SH from CO/H2/H2S has received widespread attention in recent years [15,16,17], and most research has focused on tungsten-based and molybdenum-based catalysts.
Among these, sulfurized K-Mo-based materials have received widespread attention due to their good catalytic activity and excellent sulfur resistance [18,19,20,21]. Many studies have been focused on the reactive phase, the reaction path and improving the performance of the catalysts by regulating catalyst carriers, additives, preparation methods and so on. Liu et al. [7] found that the addition of alkali metals (Na, K, Cs) can lead to different sulfidation degrees of Mo oxidized species owing to the different electric effects of alkali metals, which result in the formation of different concentrations of alkali-metal-promoted Mo containing coordinatively unsaturated sites, dramatically improving the CH3SH selectivity and CO conversion. Lu et al. [22] synthesized mesoporous silica (SBA-15)-supported, 2–4 layered, ordered and K-promoted MoS2 nanosheets. K promotion played an important role by stabilizing the C-S bond during the adsorption of the intermediate, COS, to avoid its conversion to side products in the gas phase, so the CO conversion and the selectivity towards CH3SH were obviously improved. Yu et al. [23] also found not only that K can enhance the MoS2-catalyzed synthesis of methanethiol, whereby synergy increases with alkali cation size, but also that alkali sulfides themselves were the active sites. However, there are still obstacles in CO/H2/H2S synthesis of CH3SH using a molybdenum sulfide based catalyst, such as low reactant conversion and low selectivity towards the product, CH3SH. Therefore, it is still of great significance to explore high-performance catalysts.
The transition metal carbide (TMC) molybdenum carbide (Mo2C), which acts as an electron donor, exhibits high catalytic activity and selectivity due to the redistribution of electrons after carbon introduction. It has been widely applied for selective C-C, C-O and C-H bond cleavage and for producing fuel molecules from oxygenated biomass [24,25,26,27]. Moreover, it has good sulfur resistance and is expected to replace noble metal catalysts, which are limited by their low abundance and high cost [28,29]. The current synthesis methods of Mo2C including solid–gas reactions [30,31], solid–liquid reactions [32,33,34] and solid–solid reactions [35,36,37], among which solid–gas reactions are the most commonly used and involve the carbonization reaction of molybdenum oxide (MoO3 or MoO2) with hydrogen and alkane/aromatics. During the carbonization process, different carbonization atmospheres can lead to the formation of different crystal forms of Mo2C [27]. At present, there are no reports of the synthesis of CH3SH from CO/H2/H2S over molybdenum carbide based catalysts. Due to their unique structures, molybdenum carbide based materials are expected to become emerging catalysts for the synthesis of CH3SH from CO/H2/H2S.
Here, K-Mo/Al2O3 catalysts are prepared by the impregnation method, and a series of characterizations, including H2 temperature-programmed reduction (H2-TPR), Brunauer–Emmett–Teller (BET) analysis, temperature-programmed desorption (TPD), powder X-ray diffraction (XRD) and Raman spectroscopy, are conducted to investigate the effects of catalyst carbonization and passivation conditions on catalyst activity. We show that the regulation of carbonization and passivation conditions dramatically impacts the active sites and surface chemistry of the catalysts, and we explain how this can be exploited to greatly propel the production of CH3SH from CO/H2/H2S.

2. Experimental Section

2.1. Catalyst Preparation

K-Mo/Al2O3 catalysts with a Mo loading of 10 wt% were prepared by the impregnation method, wherein γ-Al2O3 was used as the support and ammonium molybdate tetrahydrate ((NH4)6Mo7O24·4H2O) and potassium carbonate (K2CO3) were used as the precursors of Mo and K, respectively. Firstly, a certain amount of (NH4)6Mo7O24·4H2O was fully dissolved in 3.5 mL deionized water, and then K2CO3 was added to the solution at a K:Mo molar ratio of 2:1. Next, 2 g support was added to the mixed solution and stirred. The mixture was allowed to stand overnight for aging and was then placed in an oven and dried at 110 °C for 6 h. Finally, the samples were transferred to a muffle furnace and heated from room temperature to 550 °C at 5 °C/min, with calcination in air for 5 h. The resulting catalyst was labelled as K-Mo/Al2O3.

2.2. Catalyst Characterization

Powder X-ray diffraction (XRD) of catalysts was performed on a Bruker Germany D8 ADVANCE instrument with Cu Kα-radiation (λ = 0.15418 nm) operating at 40 kV and 30 Ma. Raman spectroscopy (Jobin-Yvon France LabRam HR 800) was performed under 532 nm laser excitation, and the collection range was 200–1200 cm−1. BET surface areas and pore structures of catalysts were determined by N2 physisorption at 77 K using an America Quantachrome NOVA4200e analyzer. The temperature-programmed reduction of the sulfided sample (0.05 g) by hydrogen (H2-TPR) was performed using a thermal conductivity detector (TCD) at a heating rate of 10 °C/min until reaching 800 °C. The reduction atmosphere comprised a 10% H2/Ar (v/v) gas at a flow rate of 30 mL/min. The temperature-programmed desorption of carbon monoxide/hydrogen/hydrogen sulfide (CO/H2/H2S-TPD) was conducted in a fixed-bed flow reactor equipped with a thermal conductivity detector (TCD) as the detector. Taking CO-TPD as an example, prior to the analysis, the sample (0.05 g, 40–60 mesh) was first heated to 300 °C at a rate of 5 °C/min and kept at 300 °C for 0.5 h to remove adsorbed matter from the catalyst surface. After the sample was cooled down to room temperature, the gas stream was switched to 10% CO/He (30 mL/min). The adsorption was carried out at room temperature for 1 h, and then the sample was purged with Ar to remove the physically adsorbed CO. The reactor was heated from room temperature to 900 °C at a rate of 10 °C/min for temperature-programmed desorption. With other conditions remaining unchanged, H2-TPD and H2S-TPD were performed by replacing 10% CO/He with 10% H2/Ar and 10% H2S/He, respectively.

2.3. Catalyst Activity Test

2.3.1. Catalytic Performance Tests

Before the performance test, 0.8 g K-Mo/Al2O3 (40–60 mesh) was placed in a quartz tube and a pre-treatment gas (carbonization gas: CH4/H2, C2H6/H2, C3H8/H2) was introduced. Taking CH4/H2 as an example, the ratio of CH4:H2 was 1:9 (v/v, 40 mL/min); the sample was first heated from room temperature to 300 °C under a carbonization atmosphere at a rate of 1 °C/min and was maintained for 2 h; the sample was then further heated to the carbonization temperature stage (400 °C, 500 °C, 600 °C) at a rate of 1 °C/min and held for 2 h. K-Mo/Al2O3 samples were carbonized in CH4/H2, C2H6/H2 or C3H8/H2 atmospheres and named as CH4-K-Mo2C/Al2O3 or K-Mo2C/Al2O3, C2H6-K-Mo2C/Al2O3 and C3H8-K-Mo2C/Al2O3, respectively.
The activity evaluation of the catalysts (0.4 g) was carried out on a fixed-bed reactor. A gas mixture of CO/H2/H2S (gas volume concentration: CO = 10%, H2 = 40%, H2S = 50%) was used. N2 was used as the equilibrium component and was introduced to the quartz tube with a pressure of 0.2 MPa and a flow rate of 40 mL/min. The catalytic activity was measured in the range of 275–400 °C. After reaching a steady state at each temperature, the products were analyzed online using three gas chromatographs (GC9790, China FULI INSTRUMENTS) equipped with FPD, TCD and FID.

2.3.2. Analysis of the Products

The CO conversion rate and product selectivity were calculated using the following equation:
C o n C O = C CO , in   -   C CO , out C in × 100 % Sel X =   X i / i = 1 n X i × 100 % = C i C CH 3 SH + C COS + C CO 2 + C CH 4 + C CS 2 × 100 %
where C C O , i n denotes the concentration of CO in the feed gas and C CO , out denotes the concentration of CO in the product; the product selectivity is calculated based on carbon balance, where i is the target product (CH3SH, COS, CO2, CH4, CS2). Ci is the product concentration.
The CO consumption rate and CH3SH formation rate were calculated using the following equation:
r C O = M C O × C C O , i n × c o n C O × Q V m × m c a t r C H 3 S H = M C H 3 S H × C C H 3 S H × Q V m × m c a t
where MCO is CO molar mass (g/mol); MCH3SH is the molar mass of CH3SH (g/mol). Q is the flow rate (L/h), Vm is the molar volume of gas under standard condition (L/mol) and mcat is the catalyst amount (g).

3. Results

3.1. Performance of Catalysts Prepared at Different Carbonization Temperatures

Figure 1 shows the effect of different carbonization temperatures on the catalytic performance of the K-Mo/Al2O3 catalyst for the synthesis of CH3SH from a CO/H2/H2S mixture. From Figure 1A,B, it can be seen that the CO consumption rate first increased and then decreased with increasing temperature over 325 °C. Our former studies showed that the equilibrium constant of the CO consumption reaction (CO + H2S → COS + H2) decreases with increasing temperature, which indicates that rising temperature is adverse to the occurrence of the CO consumption reaction. Therefore, the decrease in CO conversion at high temperatures is due to equilibrium limitations [7]. The CH3SH formation rate also showed a similar tendency to that of CO consumption. It can be obviously seen from the tendency that when the oxidation state material was carbonized at 500 °C, the catalyst was found to have the optimum catalytic activity, and the CO conversion (35.5%) and selectivity towards CH3SH (66.1%) reached maximum values when the reaction temperature was 325 °C. At this temperature, the CO consumption rate reached 0.2660 g·gcat−1·h−1 and the CH3SH formation rate reached 0.2990 g·gcat−1·h−1. Recently, Lu et al. investigated the synthesis of CH3SH using CO/H2/H2S as reactants over K-Mo-type catalysts. COS was demonstrated to be generated first via the reaction between CO and H2S (as shown by Equation (1)), and then CH3SH was formed via two reaction pathways, which were the hydrogenation of COS and CS2 (as shown in Equations (3) and (4)).
CO + H2S → COS + H2
COS + 3H2 → CH3SH + H2O
2COS → CO2 + CS2
CS2 + 3H2 → CH3SH + H2S
CH3SH + H2 → CH4 + H2S
From Figure 1C, it can be seen that there was no significant difference in the main products formed over K-Mo/Al2O3 after being carbonized at different temperatures for the preparation of CH3SH from CO/H2/H2S; CH3SH, COS, CO2, CH4 and CS2 were the main products in this reaction system.
The H2-TPR curves of K-Mo/Al2O3 catalysts treated at different carbonization temperatures are shown in Figure 1D. The three materials with different carbonization temperatures exhibited similar reduction peak characteristics, with a low temperature reduction peak at 200–350 °C and a moderate temperature reduction peak at 450–600 °C. Moreover, the catalyst carbonized at 500 °C exhibited a reduction peak at a high temperature (above 700 °C), which was attributed to the reduction of Mo4+ to Mo0, and the hydrogen consumption peaks were obviously lower than those of the other two materials at the same reduction temperature. These results indicated that K-Mo/Al2O3 carbonized at 500 °C was more easily reduced under the same reduction conditions, resulting in more Mo with coordinatively unsaturated surface sites that were conducive to the adsorption of reactants and the desorption of intermediate products such as COS and CH3SH. Thus, the catalyst obtained by carbonization at 500 °C delivered the best catalytic performance, and the catalysts were carbonized at 500 °C in subsequent experiments.

3.2. The Role of Passivation

According to the literature, fresh molybdenum carbide materials are prone to violent reactions with air to form molybdenum oxide [38,39,40]. Therefore, fresh molybdenum carbide materials need to be passivated. After passivation, a passivation layer is formed on the surface of the material, which makes it more stable (passivation treatment with O2/inert gas). To reveal the role of passivation in the catalytic performance of K-Mo/Al2O3 for the synthesis of CH3SH from CO/H2/H2S, the catalytic activity of the catalysts before/after passivation was conducted. As shown in Figure 2, there was a significant change in the CO consumption rate within the reaction temperature range of 275–400 °C after passivation that decreased with increasing temperature. However, the formation rate trend of CH3SH remained almost unchanged, first increasing and then decreasing with increasing temperature. After comparing with the product analysis diagram, it was found that the formation of main products before/after passivation did not change (CH3SH, COS, CO2, CH4, and CS2), but the selectivity towards CH3SH was significantly reduced and the selectivity of by-product COS was increased after passivation. These changes were likely due to the passivation layer on the passivated catalyst surface, which covered the active site for direct hydrogenation of some COS to CH3SH, resulting in the reduction of selectivity towards CH3SH. The result revealed that passivation treatment was not effective for the synthesis of CH3SH from CO/H2/H2S, as the existing of the passivation layer would inhibit the formation of CH3SH.

3.3. Performance of Catalysts Carbonized under Different Atmospheres

Different carbonization atmospheres can lead to the formation of different crystal forms of molybdenum carbide. For example, MoO3 will be converted into cubic molybdenum carbide with H2/toluene as the reaction gas at 673 K [41]; using H2/butane as the reaction gas, MoO3 is initially converted into face-centered cubic (fcc) molybdenum carbide, then the face-centered cubic molybdenum carbide gradually transforms into a hexagonal closely packed structure (fcp) with an increase in carbonization temperature [42]. Using alkanes with different carbon atomic numbers as a carbonization gas will generally lead to different carbonation degrees and then generate different numbers of active sites [27]. K-Mo/Al2O3 was carbonized by different atmospheres (CH4/H2, C2H6/H2, C3H8/H2), and the catalytic performance of different catalysts in the synthesis of CH3SH from CO/H2/H2S is shown in Figure 3. Remarkably, K-Mo/Al2O3 samples treated with different atmospheres all showed similar tendencies of catalytic performance, whereby the CO consumption rate first increased and then decreased. Catalysts treated with C3H8/H2 had the highest CO consumption rate, followed by catalysts treated with C2H6/H2, and catalysts treated with CH4/H2 had the worst CO consumption rate. From Figure 3B, it can be seen that the CH3SH generation rate of the catalyst treated with C3H8/H2, which had the highest CO consumption rate, was not optimum. Combined with the product analysis (Figure 3E), the selectivity towards by-product CO2 of this catalyst was high, which might have been caused by the hydrolysis of COS or the water–gas shift reaction of CO. Compared with that of the K-Mo2C/Al2O3 catalyst treated with C3H8/H2, the CO consumption rate of the K-Mo2C/Al2O3 catalyst treated with C2H6/H2 was slightly lower. It can be seen intuitively in Figure 3B that the CH3SH generation rate decreased sharply at high temperatures. And it was found that the selectivity of by-products of the catalyst was higher at high temperatures (COS, CH4) and the CH3SH selectivity was lower, which might be due to easy carbon deposition on the catalyst surface at high temperatures, further leading to a decrease in COS hydrogenation to generate CH3SH and excessive hydrogenation of CH3SH to generate CH4; thus, CH3SH selectivity decreased. Although the K-Mo2C/Al2O3 catalyst treated with CH4/H2 exhibited a poor CO consumption rate, it had a relatively stable CH3SH generation rate and high CH3SH selectivity. To reveal the role of carbonization atmospheres in catalytic performance, the characterization of catalysts was conducted.

3.4. Characterization of Catalysts Carbonized under Different Atmospheres

The N2 adsorption–desorption isotherms and physical properties of the catalysts assessed using N2 physisorption are displayed in Figure 4 and Table 1. As shown in Figure 4, the oxidation state sample was a typical mesoporous structure with an IV curve and an H1 hysteresis ring, and the surface area and pore volume were 116.1 m2/g and 0.269 cc/g, respectively. After carbonization, the specific surface area of all samples decreased slightly, with CH4-K-Mo2C/Al2O3 decreasing to 103.9 m2/g, C2H6-K-Mo2C/Al2O3 decreasing to 101.3 m2/g and C3H8-K-Mo2C/Al2O3 decreasing to 113.4 m2/g. However, the average pore volume of CH4-K-Mo2C/Al2O3, C2H6-K-Mo2C/Al2O3 and C3H8-K-Mo2C/Al2O3 increased to 0.321 cc/g, 0.290 cc/g and 0.390 cc/g, respectively. It can be therefore be concluded that the small decrease in specific surface area and the increase in pore volume after carbonization were caused by the collapse of some small pores during the carbonization process.
The phase compositions of the catalysts of different carbonization atmospheres were characterized by XRD and Raman spectroscopy. As shown in Figure 5A, the diffraction peaks at 2θ = 38.1°, 39.4°, 61.7° and 75.7° were attributed to β-Mo2C (JCPDS card No.65-8766) and correspond to the (002), (101), (110) and (201) crystal planes, respectively. Small diffraction peaks of K2MoO4 species (JCPDS card No.29-1021) were also detected. In the Raman spectrum (Figure 5B), diffraction peaks were detected at 325 cm−1, 896 cm−1 and 917 cm−1, which were attributed to monomolybdate MoO42− species and K2MoO4 on the sample surface [43,44], indicating an interaction between the precursor potassium and molybdenum. The diffraction peak was detected at 215 cm−1, assigned to well-dispersed AlMo6O24H63+ anderson heteropolymetalate anions (AlMo6) formed by an alumina carrier and molybdenum-based aqueous solution [45,46,47]. The diffraction peaks at 661 cm−1, 818 cm−1 and 990 cm−1 were also assigned to β-Mo2C [48,49]; among these, the peaks at 818 cm−1 and 990 cm−1 correspond to the stretching vibrations of Mo-C-Mo and Mo-C, respectively. The results of XRD and Raman spectroscopy showed that there was no significant difference in the phase structure of the catalysts under different carbonization atmospheres.
H2-TPR was carried out on the catalysts under different carbonization atmospheres to investigate redox properties; the results for the catalysts are displayed in Figure 6. It can be seen that all K-Mo2C/Al2O3 catalysts had three obvious hydrogen consumption peaks, two of which appeared at 200–400 °C and 400–600 °C, with larger hydrogen consumption peaks at >600 °C. Generally, the hydrogen consumption peak at low temperatures (200–400 °C) was the reduction of the passivation layer on the surface of the molybdenum carbide [50]. The moderate temperature reduction peak (400–600 °C) can be attributed to the reduction of Mo6+ to Mo4+ or other high-valent Mo species [51]. The high-temperature reduction peak (>600 °C) can be attributed to the reduction of Mo4+ or Mo2+ to Mo atoms [52]. However, the sample was not subjected to passivation treatment, and the presence of Mo6+ was not detected in the XRD spectrum. According to the literature [39,52], the hydrogen consumption peak at low temperatures (200–400 °C) is the reduction of surface high-valent molybdenum oxides, while the consumption peak at 650–800 °C is the reduction of Mo4+ or Mo2+ to Mo atoms. Therefore, in this work, the three hydrogen consumption peaks of 200–400 °C, 400–600 °C and >600 °C can attributed to the reduction of Mo4+, the reduction of potassium carbide species and the reduction of Mo2+ to Mo atoms, respectively. K-Mo2C/Al2O3 catalysts synthesized under different carbonization atmospheres exhibited similar H2-TPR reduction peaks, but the CH4-K-Mo2C/Al2O3 catalyst showed larger hydrogen consumption peaks at low and medium temperatures, indicating that more Mo coordinatively unsaturated surface sites were formed on the catalyst, which is more conducive to the adsorption of reactants and the desorption of products, consistent with the above catalytic performances.
The TPD technique is commonly used to characterize the surface properties and chemical reactions of a material. As an important step in a catalytic reaction, the adsorption of reactant molecules by the catalyst plays a crucial role in the catalytic performance [53,54]. To investigate the adsorption and activation abilities of K-Mo2C/Al2O3 catalysts under different carbonization atmospheres for reactants, the catalysts were characterized by carbon monoxide, hydrogen and hydrogen sulfide adsorption and desorption; the results of CO-TPD, H2-TPD and H2S-TPD of the catalysts are displayed in Figure 7. As shown in Figure 7A, the CH4-K-Mo2C/Al2O3 catalyst exhibited three CO desorption peaks, indicating three different CO adsorption sites and on the catalyst with different adsorption strengths. According to the literature, CH4-K-Mo2C/Al2O3 has a low-temperature desorption peak at around 120 °C, which was assigned to the desorption of CO and physical adsorption of Mo2C. In the high-temperature region, the desorption peak was the dissociation adsorption peak formed by the strong chemical adsorption between the catalyst and CO [52]. C2H6-K-Mo2C/Al2O3 and C3H8-K-Mo2C/Al2O3 showed one and two CO desorption peaks, respectively, which were obviously lower than those of CH4-K-Mo2C/Al2O3. According to the related research, the Mo coordinatively unsaturated sites serve as the adsorption sites for the activation of CO [6,7,55]. Therefore, the CH4/H2 atmosphere promotes the generation of more Mo coordinatively unsaturated sites on the catalyst and thus facilitates CO adsorption and activation, which also enhances the selectivity of CH3SH.
In H2-TPD, the CH4-K-Mo2C/Al2O3 catalyst exhibited slightly strong hydrogen activation ability, producing a large amount of desorbed H at medium temperature, and there was an obvious H2 negative peak, indicating that hydrogen might be dissociated into adsorbed H* at this temperature. Moreover, as shown in the H2S-TPD results (Figure 7C), the CH4-K-Mo2C/Al2O3 catalyst showed that it had strong H2S adsorption and dissociation abilities, very similar to the other two catalysts in the low temperature range but slightly stronger above 500 °C. In short, the use of CH4/H2 as a carbonization atmosphere facilitates the adsorption and activation of reactants of K-Mo2C/Al2O3, thus delivering a high activity and selectivity towards CH3SH.

4. Conclusions

In conclusion, a series of K-Mo2C/Al2O3 catalysts were successfully obtained through the impregnation method with a carbonization process at temperatures of 400–600 °C. Through careful control of the carbonization temperature, passivation process and carbonization atmosphere (500 °C, unpassivated and CH4/H2 being optimal), a CH4-K-Mo2C/Al2O3 catalyst with unusually high activity and selectivity towards CH3SH produced from CO/H2/H2S at 325 °C could be obtained (e.g., 66.1% selectivity and 0.2990 g·gcat−1·h−1 formation rate to CH3SH). Structural characterization studies revealed that passivation of the catalysts would lead to a layer formed on the surface of the material, covering the active sites of the reaction and reducing the activity of the catalyst. Under different carbonization temperatures and atmospheres (CH4/H2, C2H6/H2, C3H8/H2), the catalyst carbonized by CH4/H2 at 500 °C exhibited higher catalytic activity because it was more easily reduced under the same reduction conditions and resulted in more Mo coordinatively unsaturated sites, which was conducive to the adsorption of reactants and the desorption of intermediate products, thus showing the best overall performance. This work identifies K-Mo2C/Al2O3 as a promising new Mo-based catalyst for the production of CH3SH from CO/H2/H2S.

Author Contributions

Conceptualization, T.A. and Y.L. (Yongming Luo); data curation, H.J. and Y.L. (Yubei Li); writing-original draft preparation, X.Z.; software, Z.X. and H.J.; writing-review and editing, T.A. and Y.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Natural Science Foundation of China (No. 42207127, 42030712, 21966018), the Key Project of Natural Science Foundation of Yunnan Province (Grant No. 202101AS070026) and the Applied Basic Research Foundation of Yunnan Province (Grant No. 202101BE070001-027, 2022J0069 and 202105AE160019).

Data Availability Statement

All data used to support the findings of this study are included within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Midilli, A.; Kucuk, H.; Topal, M.E.; Akbulut, U.; Dincer, I. A comprehensive review on hydrogen production from coal gasification: Challenges and Opportunities. Int. J. Hydrogen Energy 2021, 46, 25385–25412. [Google Scholar] [CrossRef]
  2. Claure, M.T.; Chai, S.-H.; Dai, S.; Unocic, K.A.; Alamgir, F.M.; Agrawal, P.K.; Jones, C.W. Tuning of higher alcohol selectivity and productivity in CO hydrogenation reactions over K/MoS2 domains supported on mesoporous activated carbon and mixed MgAl oxide. J. Catal. 2015, 324, 88–97. [Google Scholar] [CrossRef]
  3. Wagner, N.J.; Coertzen, M.; Matjie, R.H.; van Dyk, J.C. Chapter 5—Coal Gasification. In Applied Coal Petrology; Suárez-Ruiz, I., Crelling, J.C., Eds.; Elsevier: Burlington, MA, USA, 2008; pp. 119–144. [Google Scholar]
  4. Renda, S.; Barba, D.; Palma, V. Recent Solutions for Efficient Carbonyl Sulfide Hydrolysis: A Review. Ind. Eng. Chem. Res. 2022, 61, 5685–5697. [Google Scholar] [CrossRef]
  5. He, B.; Xu, Z.; Cao, X.; Lu, J.; Fang, J.; Li, Y.; Feng, S.; Luo, Y. The influence of lanthanum (La) promoter on Mo/Al2O3 for the catalytic synthesis of methyl mercaptan (CH3SH) from COS/H2/H2S. Fuel Process. Technol. 2022, 238, 107488. [Google Scholar] [CrossRef]
  6. Lu, J.; Liu, P.; Xu, Z.; He, S.; Luo, Y. Investigation of the reaction pathway for synthesizing methyl mercaptan (CH3SH) from H2S-containing syngas over K–Mo-type materials. RSC Adv. 2018, 8, 21340–21353. [Google Scholar] [CrossRef] [PubMed]
  7. Liu, P.; Lu, J.; Xu, Z.; Liu, F.; Chen, D.; Yu, J.; Liu, J.; He, S.; Wan, G.; Luo, Y. The effect of alkali metals on the synthesis of methanethiol from CO/H2/H2S mixtures on the SBA-15 supported Mo-based catalysts. Mol. Catal. 2017, 442, 39–48. [Google Scholar] [CrossRef]
  8. Cordova, A.; Blanchard, P.; Lancelot, C.; Frémy, G.; Lamonier, C. Probing the Nature of the Active Phase of Molybdenum-Supported Catalysts for the Direct Synthesis of Methylmercaptan from Syngas and H2S. ACS Catal. 2015, 5, 2966–2981. [Google Scholar] [CrossRef]
  9. Gutiérrez, O.Y.; Kaufmann, C.; Hrabar, A.; Zhu, Y.; Lercher, J.A. Synthesis of methyl mercaptan from carbonyl sulfide over sulfide K2MoO4/SiO2. J. Catal. 2011, 280, 264–273. [Google Scholar] [CrossRef]
  10. Gutierrez, O.Y.; Zhong, L.; Zhu, Y.; Lercher, J.A. Synthesis of Methanethiol from CS2 on Ni-, Co-, and K-Doped MoS2/SiO2 Catalysts. Chemcatchem 2013, 5, 3249–3259. [Google Scholar] [CrossRef]
  11. Gutiérrez, O.Y.; Kaufmann, C.; Lercher, J.A. Synthesis of Methanethiol from Carbonyl Sulfide and Carbon Disulfide on (Co)K-Promoted Sulfide Mo/SiO2 Catalysts. ACS Catal. 2011, 1, 1595–1603. [Google Scholar] [CrossRef]
  12. Paskach, T.J.; Schrader, G.L.; McCarley, R.E. Synthesis of Methanethiol from Methanol over Reduced Molybdenum Sulfide Catalysts Based on the Mo6S8 Cluster. J. Catal. 2002, 211, 285–295. [Google Scholar]
  13. Kramer, R.L.; Reid, E.E. The catalytic preparation of mercaptans.1. J. Am. Chem. Soc. 1921, 43, 880–890. [Google Scholar] [CrossRef]
  14. Olin, J.F.; Bernard, B.; Goshorn, R.H.J.F. Process for Preparation of Methyl mercaptan. US Patent 3070632, 25 December 1962. [Google Scholar]
  15. Cordova, A.; Blanchard, P.; Salembier, H.; Lancelot, C.; Frémy, G.; Lamonier, C. Direct synthesis of methyl mercaptan from H2/CO/H2S using tungsten based supported catalysts: Investigation of the active phase. Catal. Today 2017, 292, 143–153. [Google Scholar] [CrossRef]
  16. Zhang, B.; Taylor, S.H.; Hutchings, G.J. Catalytic synthesis of methanethiol from CO/H2/H2S mixtures using α-Al2O3. New J. Chem. 2004, 28, 471–476. [Google Scholar] [CrossRef]
  17. Zhang, B.; Taylor, S.H.; Hutchings, G.J. Synthesis of Methyl Mercaptan and Thiophene from CO/H2/H2S Using α-Al2O3. Catal. Lett. 2003, 91, 181–183. [Google Scholar] [CrossRef]
  18. Wang, Q.; Hao, Y.; Chen, A.; Yang, Y. Effect of Thermal Treatment on Structure and Catalytic Performance of K2MoO4-NiO/SiO2 Catalyst for One-Step Synthesis of Methanethiol from High H2S-Containing Syngas. Chin. J. Catal. 2010, 31, 242–247. [Google Scholar]
  19. Chen, A.; Wang, Q.; Hao, Y.; Fang, W.; Yang, Y. The promoting effect of tellurium on K2MoO4/SiO2 catalyst for methanethiol synthesis from high H2S-Containing syngas. Catal. Lett. 2007, 118, 295–299. [Google Scholar] [CrossRef]
  20. Yang, Y.Q.; Dai, S.J.; Yuan, Y.Z.; Lin, R.C.; Tang, D.L.; Zhang, H.B. The promoting effects of La2O3 and CeO2 on K2MoS4/SiO2 catalyst for methanthiol synthesis from syngas blending with H2S. Appl. Catal. A Gen. 2000, 192, 175–180. [Google Scholar] [CrossRef]
  21. Dai, S.J.; Yang, Y.Q.; Yuan, Y.Z.; Tang, D.L.; Lin, R.C.; Zhang, H.B. On methanethiol synthesis from H2S-containing syngas over K2MoS4/SiO2 catalysts promoted with transition metal oxides. Catal. Lett. 1999, 61, 157–160. [Google Scholar] [CrossRef]
  22. Lu, J.; Fang, J.; Xu, Z.; He, D.; Feng, S.; Li, Y.; Wan, G.; He, S.; Wu, H.; Luo, Y. Facile synthesis of few-layer and ordered K-promoted MoS2 nanosheets supported on SBA-15 and their potential application for heterogeneous catalysis. J. Catal. 2020, 385, 107–119. [Google Scholar] [CrossRef]
  23. Yu, M.; Chang, M.-W.; Kosinov, N.; Hensen, E.J.M. Alkali catalyzes methanethiol synthesis from CO and H2S. J. Catal. 2022, 405, 116–128. [Google Scholar] [CrossRef]
  24. Hua, W.; Sun, H.-H.; Xu, F.; Wang, J.-G. A review and perspective on molybdenum-based electrocatalysts for hydrogen evolution reaction. Rare Met. 2020, 39, 335–351. [Google Scholar] [CrossRef]
  25. Alexander, A.-M.; Hargreaves, J.S.J. Alternative catalytic materials: Carbides, nitrides, phosphides and amorphous boron alloys. Chem. Soc. Rev. 2010, 39, 4388–4401. [Google Scholar] [CrossRef]
  26. Hwu, H.H.; Chen, J.G.G. Surface chemistry of transition metal carbides. Chem. Rev. 2005, 105, 185–212. [Google Scholar] [CrossRef]
  27. Xiao, T.-C.; York, A.P.E.; Al-Megren, H.; Williams, C.V.; Wang, H.-T.; Green, M.L.H. Preparation and Characterisation of Bimetallic Cobalt and Molybdenum Carbides. J. Catal. 2001, 202, 100–109. [Google Scholar] [CrossRef]
  28. Schaidle, J.A.; Lausche, A.C.; Thompson, L.T. Effects of sulfur on Mo2C and Pt/Mo2C catalysts: Water gas shift reaction. J. Catal. 2010, 272, 235–245. [Google Scholar] [CrossRef]
  29. Lewandowski, M.; Szymańska-Kolasa, A.; Da Costa, P.; Sayag, C. Catalytic performances of platinum doped molybdenum carbide for simultaneous hydrodenitrogenation and hydrodesulfurization. Catal. Today 2007, 119, 31–34. [Google Scholar] [CrossRef]
  30. Zhang, S.; Shi, C.; Chen, B.; Zhang, Y.; Zhu, Y.; Qiu, J.; Au, C. Catalytic role of β-Mo2C in DRM catalysts that contain Ni and Mo. Catal. Today 2015, 258, 676–683. [Google Scholar] [CrossRef]
  31. Zhang, A.; Zhu, A.; Chen, B.; Zhang, S.; Au, C.; Shi, C. In-situ synthesis of nickel modified molybdenum carbide catalyst for dry reforming of methane. Catal. Commun. 2011, 12, 803–807. [Google Scholar] [CrossRef]
  32. Giordano, C.; Antonietti, M. Synthesis of crystalline metal nitride and metal carbide nanostructures by sol–gel chemistry. Nano Today 2011, 6, 366–380. [Google Scholar] [CrossRef]
  33. Giordano, C.; Erpen, C.; Yao, W.; Antonietti, M. Synthesis of Mo and W Carbide and Nitride Nanoparticles via a Simple “Urea Glass” Route. Nano Lett. 2008, 8, 4659–4663. [Google Scholar] [CrossRef]
  34. Gu, Y.; Li, Z.; Chen, L.; Ying, Y.; Qian, Y. Synthesis of nanocrystalline Mo2C via sodium co-reduction of MoCl5 and CBr4 in benzene. Mater. Res. Bull. 2003, 38, 1119–1122. [Google Scholar] [CrossRef]
  35. Liao, L.; Wang, S.; Xiao, J.; Bian, X.; Zhang, Y.; Scanlon, M.D.; Hu, X.; Tang, Y.; Liu, B.; Girault, H.H. A nanoporous molybdenum carbide nanowire as an electrocatalyst for hydrogen evolution reaction. Energy Environ. Sci. 2014, 7, 387–392. [Google Scholar] [CrossRef]
  36. Morishita, T.; Soneda, Y.; Hatori, H.; Inagaki, M. Carbon-coated tungsten and molybdenum carbides for electrode of electrochemical capacitor. Electrochim. Acta 2007, 52, 2478–2484. [Google Scholar] [CrossRef]
  37. Nartowski, A.M.; Parkin, I.P.; Mackenzie, M.; Craven, A.J. Solid state metathesis: Synthesis of metal carbides from metal oxides. J. Mater. Chem. 2001, 11, 3116–3119. [Google Scholar] [CrossRef]
  38. Lin, Z.; Wan, W.; Yao, S.; Chen, J.G. Cobalt-modified molybdenum carbide as a selective catalyst for hydrodeoxygenation of furfural. Appl. Catal. B Environ. 2018, 233, 160–166. [Google Scholar] [CrossRef]
  39. Liu, X.; Kunkel, C.; Ramírez de la Piscina, P.; Homs, N.; Viñes, F.; Illas, F. Effective and Highly Selective CO Generation from CO2 Using a Polycrystalline α-Mo2C Catalyst. ACS Catal. 2017, 7, 4323–4335. [Google Scholar] [CrossRef]
  40. Sullivan, M.M.; Held, J.T.; Bhan, A. Structure and site evolution of molybdenum carbide catalysts upon exposure to oxygen. J. Catal. 2015, 326, 82–91. [Google Scholar] [CrossRef]
  41. Vitale, G.; Frauwallner, M.L.; Hernandez, E.; Scott, C.E.; Pereira-Almao, P. Low temperature synthesis of cubic molybdenum carbide catalysts via pressure induced crystallographic orientation of MoO3 precursor. Appl. Catal. A Gen. 2011, 400, 221–229. [Google Scholar] [CrossRef]
  42. Xiao, T.-C.; York, A.P.E.; Williams, V.C.; Al-Megren, H.; Hanif, A.; Zhou, X.-Y.; Green, M.L.H. Preparation of Molybdenum Carbides Using Butane and Their Catalytic Performance. Chem. Mater. 2000, 12, 3896–3905. [Google Scholar] [CrossRef]
  43. Pérez-Martínez, D.J.; Eloy, P.; Gaigneaux, E.M.; Giraldo, S.A.; Centeno, A. Study of the selectivity in FCC naphtha hydrotreating by modifying the acid–base balance of CoMo/γ-Al2O3 catalysts. Appl. Catal. A Gen. 2010, 390, 59–70. [Google Scholar] [CrossRef]
  44. Hao, Y.; Zhang, Y.; Chen, A.; Fang, W.; Yang, Y. Study on Methanethiol Synthesis from H2S-Rich Syngas Over K2MoO4 Catalyst Supported on Electrolessly Ni-Plated SiO2. Catal. Lett. 2009, 129, 486–492. [Google Scholar] [CrossRef]
  45. Blanchard, P.; Lamonier, C.; Griboval, A.; Payen, E. New insight in the preparation of alumina supported hydrotreatment oxidic precursors: A molecular approach. Appl. Catal. A Gen. 2007, 322, 33–45. [Google Scholar] [CrossRef]
  46. Le Bihan, L.; Blanchard, P.; Fournier, M.; Grimblot, J.; Payen, E. Raman spectroscopic evidence for the existence of 6-molybdoaluminate entities on an Mo/Al2O3 oxidic precursor. J. Chem. Soc. Faraday Trans. 1998, 94, 937–940. [Google Scholar] [CrossRef]
  47. Carrier, X.; Lambert, J.F.; Che, M. Ligand-Promoted Alumina Dissolution in the Preparation of MoOX/γ-Al2O3 Catalysts:  Evidence for the Formation and Deposition of an Anderson-type Alumino Heteropolymolybdate. J. Am. Chem. Soc. 1997, 119, 10137–10146. [Google Scholar] [CrossRef]
  48. Hussain, S.; Zaidi, S.A.; Vikraman, D.; Kim, H.-S.; Jung, J. Facile preparation of molybdenum carbide (Mo2C) nanoparticles and its effective utilization in electrochemical sensing of folic acid via imprinting. Biosens. Bioelectron. 2019, 140, 111330. [Google Scholar] [CrossRef]
  49. Huang, Y.; Gong, Q.; Song, X.; Feng, K.; Nie, K.; Zhao, F.; Wang, Y.; Zeng, M.; Zhong, J.; Li, Y. Mo2C Nanoparticles Dispersed on Hierarchical Carbon Microflowers for Efficient Electrocatalytic Hydrogen Evolution. ACS Nano 2016, 10, 11337–11343. [Google Scholar] [CrossRef]
  50. Pang, M.; Wang, X.; Xia, W.; Muhler, M.; Liang, C. Mo(VI)–Melamine Hybrid As Single-Source Precursor to Pure-Phase β-Mo2C for the Selective Hydrogenation of Naphthalene to Tetralin. Ind. Eng. Chem. Res. 2013, 52, 4564–4571. [Google Scholar] [CrossRef]
  51. Malaibari, Z.O.; Croiset, E.; Amin, A.; Epling, W. Effect of interactions between Ni and Mo on catalytic properties of a bimetallic Ni-Mo/Al2O3 propane reforming catalyst. Appl. Catal. A Gen. 2015, 490, 80–92. [Google Scholar] [CrossRef]
  52. Wang, G.; Schaidle, J.A.; Katz, M.B.; Li, Y.; Pan, X.; Thompson, L.T. Alumina supported Pt–Mo2C catalysts for the water–gas shift reaction. J. Catal. 2013, 304, 92–99. [Google Scholar] [CrossRef]
  53. Tian, R.; Lu, J.; Xu, Z.; Zhang, W.; Liu, J.; Wang, L.; Xie, Y.; Zhao, Y.; Cao, X.; Luo, Y. Unraveling the Synergistic Reaction and the Deactivation Mechanism for the Catalytic Degradation of Double Components of Sulfur-Containing VOCs over ZSM-5-Based Materials. Environ. Sci. Technol. 2023, 57, 1443–1455. [Google Scholar] [CrossRef]
  54. Lu, J.; Tian, R.; Zhang, W.; Zhang, Y.; Yang, Y.; Xu, Z.; He, D.; Ai, T.; Luo, Y. An ultra-long stability of lanthanum (La) modified molecular sieve for catalytic degradation of typical sulfur-containing VOCs in a near-real environment. Appl. Catal. B Environ. 2023, 339, 123114. [Google Scholar] [CrossRef]
  55. Fang, J.; Lu, J.; Xu, Z.; Feng, S.; Li, Y.; He, B.; Luo, M.; Wang, H.; Luo, Y. Modulation the metal-support interactions of potassium molybdenum-based catalysts for tuned catalytic performance of synthesizing CH3SH. Sep. Purif. Technol. 2023, 316, 123815. [Google Scholar] [CrossRef]
Figure 1. K-Mo/Al2O3 catalysts prepared with different carbonization temperatures were used to synthesize CH3SH from CO/H2/H2S; (a) K-Mo/Al2O3-600 °C; (b) K-Mo/Al2O3-500 °C; (c) K-Mo/Al2O3-400 °C. Rate of CO consumption (A); Rate of CH3SH formation (B); Distribution of CO conversion and main product selectivity (C); Curves of H2-TPR (D).
Figure 1. K-Mo/Al2O3 catalysts prepared with different carbonization temperatures were used to synthesize CH3SH from CO/H2/H2S; (a) K-Mo/Al2O3-600 °C; (b) K-Mo/Al2O3-500 °C; (c) K-Mo/Al2O3-400 °C. Rate of CO consumption (A); Rate of CH3SH formation (B); Distribution of CO conversion and main product selectivity (C); Curves of H2-TPR (D).
Nanomaterials 13 02602 g001
Figure 2. Catalytic activity of K-Mo2C/Al2O3 catalysts before and after passivation, (a) before passivation, (b) after passivation. Rate of CO consumption (A); Rate of CH3SH formation (B); Distribution diagram of unpassivated catalyst products (C); Distribution diagram of unpassivated catalyst products (D).
Figure 2. Catalytic activity of K-Mo2C/Al2O3 catalysts before and after passivation, (a) before passivation, (b) after passivation. Rate of CO consumption (A); Rate of CH3SH formation (B); Distribution diagram of unpassivated catalyst products (C); Distribution diagram of unpassivated catalyst products (D).
Nanomaterials 13 02602 g002
Figure 3. Effect of different carbonization atmospheres on the catalytic performance of K-Mo2C/Al2O3 catalysts, (a) CH4/H2, (b) C2H6/H2, (c) C3H8/H2. Rate of CO consumption (A); Rate of CH3SH formation (B); Distribution diagram of carbonization catalyst products in CH4/H2 atmosphere (C); Distribution diagram of carbonization catalyst products in C2H6/H2 atmosphere (D); Distribution diagram of carbonization catalyst products in C3H8/H2 atmosphere (E).
Figure 3. Effect of different carbonization atmospheres on the catalytic performance of K-Mo2C/Al2O3 catalysts, (a) CH4/H2, (b) C2H6/H2, (c) C3H8/H2. Rate of CO consumption (A); Rate of CH3SH formation (B); Distribution diagram of carbonization catalyst products in CH4/H2 atmosphere (C); Distribution diagram of carbonization catalyst products in C2H6/H2 atmosphere (D); Distribution diagram of carbonization catalyst products in C3H8/H2 atmosphere (E).
Nanomaterials 13 02602 g003
Figure 4. N2 adsorption–desorption isotherms of oxidation K-Mo/Al2O3 and Mo2C/Al2O3 samples treated in different carbonization atmospheres, (a) oxidized, (b) CH4/H2, (c) C2H6/H2, (d) C3H8/H2.
Figure 4. N2 adsorption–desorption isotherms of oxidation K-Mo/Al2O3 and Mo2C/Al2O3 samples treated in different carbonization atmospheres, (a) oxidized, (b) CH4/H2, (c) C2H6/H2, (d) C3H8/H2.
Nanomaterials 13 02602 g004
Figure 5. XRD (A) and Raman (B) patterns of K-Mo2C/Al2O3 samples under different carbonation atmospheres, (a) CH4/H2, (b) C2H6/H2, (c) C3H8/H2.
Figure 5. XRD (A) and Raman (B) patterns of K-Mo2C/Al2O3 samples under different carbonation atmospheres, (a) CH4/H2, (b) C2H6/H2, (c) C3H8/H2.
Nanomaterials 13 02602 g005
Figure 6. The H2-TPR spectra of K-Mo2C/Al2O3 samples in different carbonation atmospheres, (a) CH4/H2, (b) C2H6/H2, (c) C3H8/H2.
Figure 6. The H2-TPR spectra of K-Mo2C/Al2O3 samples in different carbonation atmospheres, (a) CH4/H2, (b) C2H6/H2, (c) C3H8/H2.
Nanomaterials 13 02602 g006
Figure 7. CO-TPD (A), H2-TPD (B) and H2S-TPD (C) spectra of (a) CH4-K-Mo2C/Al2O3, (b) C2H6-K-Mo2C/Al2O3, (c) C3H8-K-Mo2C/Al2O3 samples.
Figure 7. CO-TPD (A), H2-TPD (B) and H2S-TPD (C) spectra of (a) CH4-K-Mo2C/Al2O3, (b) C2H6-K-Mo2C/Al2O3, (c) C3H8-K-Mo2C/Al2O3 samples.
Nanomaterials 13 02602 g007
Table 1. Textural characteristics of oxidation K-Mo/Al2O3 and Mo2C/Al2O3 samples treated in different carbonization atmospheres.
Table 1. Textural characteristics of oxidation K-Mo/Al2O3 and Mo2C/Al2O3 samples treated in different carbonization atmospheres.
SampleSurface Area (m2/g)Pore Volume (cc/g)Pore Diameter Dv (d) (nm)
K-Mo/Al2O3116.10.2697.853
CH4-K-Mo2C/Al2O3103.90.3219.618
C2H6-K-Mo2C/Al2O3101.30.2908.598
C3H8-K-Mo2C/Al2O3113.40.39011.159
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zheng, X.; Ai, T.; Hu, Y.; Xu, Z.; Li, Y.; Jiang, H.; Luo, Y. Influence of Carbonization Conditions on Structural and Surface Properties of K-Doped Mo2C Catalysts for the Synthesis of Methyl Mercaptan from CO/H2/H2S. Nanomaterials 2023, 13, 2602. https://doi.org/10.3390/nano13182602

AMA Style

Zheng X, Ai T, Hu Y, Xu Z, Li Y, Jiang H, Luo Y. Influence of Carbonization Conditions on Structural and Surface Properties of K-Doped Mo2C Catalysts for the Synthesis of Methyl Mercaptan from CO/H2/H2S. Nanomaterials. 2023; 13(18):2602. https://doi.org/10.3390/nano13182602

Chicago/Turabian Style

Zheng, Xiangqian, Tianhao Ai, Yuhong Hu, Zhizhi Xu, Yubei Li, Huan Jiang, and Yongming Luo. 2023. "Influence of Carbonization Conditions on Structural and Surface Properties of K-Doped Mo2C Catalysts for the Synthesis of Methyl Mercaptan from CO/H2/H2S" Nanomaterials 13, no. 18: 2602. https://doi.org/10.3390/nano13182602

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