Next Article in Journal
Review on the Biomedical and Environmental Applications of Nonthermal Plasma
Previous Article in Journal
Real-Time UV/VIS Spectroscopy to Observe Photocatalytic Degradation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Perspective

Design of Cu/MoOx for CO2 Reduction via Reverse Water Gas Shift Reaction

1
Engineering Research Center for Waste Oil Recovery Technology and Equipment, School of Environment and Resources, Ministry of Education, Chongqing Technology and Business University, Chongqing 400067, China
2
The First Affiliated Hospital of Chongqing Medical, Pharmaceutical College, Chongqing 400060, China
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(4), 684; https://doi.org/10.3390/catal13040684
Submission received: 27 February 2023 / Revised: 24 March 2023 / Accepted: 28 March 2023 / Published: 31 March 2023

Abstract

:
CO2 reduction to CO as raw material for conversion to chemicals and gasoline fuels via the reverse water–gas shift (RWGS) reaction is generally acknowledged to be a promising strategy that makes the CO2 utilization process more economical and efficient. Cu-based catalysts are low-cost and have high catalytic performance but have insufficient stability due to hardening at high temperatures. In this work, a series of Cu-based catalysts supported by MoOx were synthesized for noble metal-free RWGS reactions, and the effects of MoOx support on catalyst performance were investigated. The results show that the introduction of MoOx can effectively improve the catalytic performance of RWGS reactions. The obtained Cu/MoOx (1:1) catalyst displays excellent activity with 35.85% CO2 conversion and 99% selectivity for CO at 400 °C. A combination of XRD, XPS, and HRTEM characterization results demonstrate that MoOx support enhances the metal-oxide interactions with Cu through electronic modification and geometric coverage, thus obtaining highly dispersed copper and more Cu-MoOx interfaces as well as more corresponding oxygen vacancies.

1. Introduction

In recent years, with the increase in CO2 emissions, the greenhouse effect has become increasingly serious, which has led to a series of natural disasters, causing widespread global attention to the field of CO2 capture, conversion, and utilization [1,2,3]. At present, carbon capture and storage (CCS) have been successfully used to reduce greenhouse gas emissions directly into the atmosphere by geological, mineralization, or oceanic means, but CO2, as a rich and cheap carbon source, has not been effectively utilized [4,5]. CO2 reduction to CO as raw material for conversion to chemicals and gasoline fuels via the reverse water–gas shift (RWGS) reaction is generally acknowledged to be a promising strategy that makes the CO2 utilization process more economical and efficient on a large industrial scale and realizes a good carbon cycle [6].
However, the high stability of CO2, the endothermic characteristics, thermodynamically favorable at high reaction temperatures, and the additional side reactions inhibit the industrialization of the RWGS reaction. Therefore, it is a great challenge to design high-activity and selectivity catalysts at a mild temperature range. At present, researchers and institutions mainly focus on the catalytic performance of metal catalysts [7,8], metal oxide catalysts [9,10,11], and transition metal carbides [12,13,14] for RWGS reactions. Su et al. systematically studied the CO2 hydrogenation on a Pt (111) surface and revealed that CO2 can be activated on a clean Pt (111) surface through the dissociation mechanism to form adsorbed CO and O at room temperature and at high temperatures [15]. Xu et al. reported that K doping changed the chemical states of Mo and Cu through interface interaction and electron transfer, resulting in improved adsorption and activation performances of CO2 on the K-promoted Cu/β-Mo2C catalysts [16]. Nickel-based catalysts can effectively catalyze the RWGS reaction, but their selectivity toward CO at low temperatures is poor. Researchers found that RWGS reactions and product selectivity can be adjusted by surface modification of Ni. The added MoOx and Ni have an interface effect, which improves the selectivity of CO through both geometric coverage and electronic modification [17].
Copper-based catalysts are very well-known for their high catalytic activity and low price, but the high temperature hardening and instability hinder their usefulness [18]. Cu catalysts often improve catalytic performance by combining with oxide materials, which will modify the surface of copper and promote the dispersion of copper particles, form strong interface interactions between Cu and oxides, and affect the chemical state of Cu through electronic transfer. The copper was deposited on silica to obtain highly dispersed nano-Cu/SiO2 catalysts. The catalyst has abundant acid sites and a specific surface area, which is beneficial for improved MTHF selectivity [19]. CeO2 and Fe2O3 are usually used as effective redox and textural promoters to form metal–metal oxide interfaces with Cu, which can boost oxygen vacancies and dispersion of the metal to facilitate the RWGS reaction [20,21,22]. Molybdenum oxide can be pre-reduced to obtain MoOx with oxygen defects on the surface and adjust its surface electronic structure. Using MoOx as the support for Cu catalysis, driven by the control of defect engineering, the Cu/MoOx catalyst surface will produce more surface metal atoms with low coordination, which will improve the binding ability of the material surface with other atoms, ions, or small molecules. Moreover, the introduction of MoOx will form new oxygen vacancies in Cu/MoOx materials through electronic modification, and oxygen vacancies are often related to greater RWGS reaction activity [23]. Therefore, understanding how the structure and properites of the Cu/MoOx catalysts affect the activity and selectivity of CO2 reduction helps to design more efficient catalysts for the RWGS reaction.
In this study, a series of Cu-based catalysts supported by MoOx were synthesized by the hydrothermal method for noble metal-free RWGS reactions. A combination of XRD, XPS, and HRTEM characterization results demonstrate that MoOx support enhances the metal-oxide interactions with Cu through electronic modification and geometric coverage, thus obtaining highly dispersed copper and more Cu-MoOx interfaces as well as more corresponding oxygen vacancies. The introduction of MoOx can effectively improve the catalytic performance of Cu/MoOx in the RWGS reaction.

2. Results and Discussion

2.1. Catalyst Characterization

XRD patterns of Cu/MoOx catalysts pre-reduced at 450 °C for 2 h with different atomic ratios of Cu/Mo (1:2, 1:1, and 2:1) are displayed in Figure 1. Two diffraction peaks at approximately 43.3° and 50.5° are observed for all samples, which are attributed to the (111) and (200) planes of the cubic Cu phase (JCPDS No. 04-0836). At the same time, mixed phases of monoclinic MoO2 (JCPDS No. 32-0671) and hexagonal MoO3 (JCPDS No. 21-0569) can also be obtained in Cu/MoOx with different atomic ratios of Cu/Mo. When Cu/Mo is 2:1, there is (NH4)2Mo4O3 (2θ = 15.1°, 30.2°) as an impurity phase. Due to the high concentration of Cu2+ in the system consumed H+, the reaction between H+ and (Mo7O24) 6− is incomplete, and the intermediate state (NH4)2Mo4O3 appears [24].
Typical SEM micrographs of Cu/MoOx samples with different atomic ratios of Cu/Mo are shown in Figure 2. The SEM images clearly show the difference in the microstructure of the Cu/MoOx powder with different atomic ratios. The Cu/MoOx powders with a Cu/Mo ratio of 1:1 (Figure 2b,e) show micro/nano structures formed by loose sheets and particles. The particle size distribution is from 0.2 μm to 2.5 μm with many holes and caves. This micro/nano porous structure provides a larger specific surface area and more active sites, which is beneficial for the transportation of reactants and products in the reaction process. When Cu/Mo ratios are 1:2 and 2:1, the two samples appear to be composed of more widely distributed and irregular particles with smooth surfaces, as shown in Figure 2a,d,c,f.
Figure 3a is the TEM image of Cu/MoOx (1:1). Several sheet-like morphologies in Cu/MoOx ranging from several tens to hundreds of nanometers of particulates exhibit close contact, which enhances the interaction between Cu and MoOx. To analyze the crystalline structural details of the nanoparticles, HRTEM is measured and shown in Figure 3b. Cu/MoOx catalyst shows three distinct lattice fringes of 0.210 nm, 0.343 nm, and 0.347 nm, which correspond to characteristic Cu (111), MoO2 (−111), and MoO3 (210) planes, respectively. Moreover, it can be seen that the Cu/MoOx catalyst contains abundant Cu-MoOx interfaces, which is attributed to the close contact between the small size of metallic Cu and MoOx nanoparticles.
The interaction between Cu and MoOx was investigated by using H2−TPR. As shown in Figure 4, a reduction peak for MoOx is observed at 620 °C, which is related to the reduction of MoO3 to MoO2. The effect of copper addition to MoOx is largely beneficial for overall reducibility. Cu/MoOx exhibits significantly larger peaks and is shifted to lower temperatures compared to the MoOx. It can be attributed to the copper-support interaction at the interface of copper and MoOx, which has a beneficial effect on the redox property of Cu/MoOx precursors [25]. For Cu/MoOx precursors with different atomic ratios of Cu/Mo, the first reduction peak shifts to a higher temperature compared with pure Cu. This trend indicates that the MoOx stabilizes the Cu species [17]. As can be seen in Figure 4, the Cu/MoOx (1:1) shows the lowest reduction temperatures. The two main reduction peaks are located at 330 °C and 380 °C, indicating that it is more easily reduced. However, with increasing Cu proportion (Cu/Mo ratio of 2:1), the reduction peak gradually moves towards higher temperatures. The excessive CuO can induce aggregation and overlap of Cu species on the surface of the precursor in the process of calcination and obstruct the pores of MoOx, which blocks the entry of H2 [26]. This is consistent with the results of SEM. As a result, the reducibility of the corresponding precursor is gradually weakened. It is worth noting that the reduction peak temperature of the catalyst (Cu:Mo = 1:2) moves higher and the peak area is relatively smaller, indicating that Mo species cannot be reduced completely to the metallic state and a lower Mo/Cu ratio corresponds to a higher reduction degree of the Mo species [27].
X-ray photoelectron spectroscopy (XPS) was utilized to study the surface element composition and the existing state of surface elements. The XPS spectrum of Cu/MoOx is shown in Figure 5 and Table 1. The metal content is close to the theoretical ratios. It is feasible to synthesize Cu-based catalysts supported on MoOx in different proportions by the hydrothermal method. Further, the N2 adsorption-desorption isotherms analysis shows that the specific surface area is 146.0, 156.7, and 117.5 m2 g−1 for Cu/MoOx catalysts with Cu/Mo ratios of 1:2, 1:1, and 2:1, respectively. The materials are mesoporous, with a variation in BJH pore diameter of 9–17 nm.
As can be seen in Figure 5a, the Cu, Mo, and O are identified for all the Cu/MoOx catalysts without impurities on the surface of the compositions. The standard peak of C 1s is at 284.78 eV as the reference. Figure 5b is the O1s XPS spectra of the studied Cu/MoOx catalysts. The XPS of O1s can be deconvoluted into three major peaks. The peak at 529.5–530.5 eV corresponds to lattice oxygen (OLatt) enclosed by either copper or molybdenum; a higher binding energy peak at 531.0–532.0 eV is related to the surface oxygen (OSur) atom or OH, COOH group attached, whereas the peak at 533.5 eV may be due to the presence of adsorbed H2O in the Cu/MoOx nanocomposite [28,29]. Compared with the Cu/MoOx sample with a Cu/Mo ratio of 2:1, the OLatt XPS peaks of two other catalysts with a Cu/Mo ratio of 1:1 and 1:2 obviously shift to the high binding energy direction. This may be due to the increased molybdenum content significantly increasing the content of Mo4+ and Mo+6 species in Cu/MoOx catalysts, and the Mo4+ species changing the chemical environment of OLatt species [30]. The transfer of electrons from the relatively electron-rich d band of Cu to MoOx species is realized through the strong electronic interaction between Cu and MoOx. Compared with the catalysts with a Cu/Mo ratio of 2:1 and 1:2, the catalyst with a Cu/Mo ratio of 1:1 has a larger OLatt XPS peak area and a smaller OSur XPS peak area. The catalyst with a Cu/Mo ratio of 1:1 has a lower OSur/Olatt value, indicating that there are more oxygen vacancies in the sample. The oxygen vacancies are considered to be the main active sites for CO2 catalytic hydrogenation, which may lead to superior catalytic performances of Cu/MoOx (1:1) catalysts in CO2 hydrogenation reduction reactions [31].
The XPS spectra of Mo in Cu/MoOx are shown in Figure 6. Two distinct peaks centered at 232.00 eV and 235.60 eV are consistent with Mo4+ 3d5/2 and Mo6+ 3d5/2, and according to the previous references and the standard XPS spectrum for MoOx (x = 2~3) (Mo4+ 3d5/2 at 232.4 eV, Mo6+ 3d5/2 at 236.0 eV), the oxidation state of Mo is Mo (IV) and (VI) [32,33,34]. The results are consistent with the XRD measurement: Cu/MoOx (x = 2~3) is a functional heterostructure material. Increasing the Mo/Cu ratio results in a shift of peaks to a higher value, and we speculate that the electron transfer from Cu to MoOx species occurred due to strong electronic interactions between Cu and MoOx species.

2.2. Catalytic Performance

The performance of the Cu/MoOx catalyst was investigated by the RWGS reaction with an H2/CO2 ratio of 4:1 from 200 °C to 400 °C at 1 bar. As depicted in Figure 7, the CO2 conversion of three samples increases with the reaction temperature rising, evidencing an endothermic process of the RWGS reaction. Below 280 °C, the catalysts (Cu:Mo = 2:1) have a higher CO2 conversion because Cu0 is the main active site for RWGS. The catalyst with a Cu/Mo ratio of 1:1 displays the best catalytic activity, showing 35.85% CO2 conversion at 400 °C. Obviously, the support from MoOx can affect the catalytic activity. Metal-oxide interactions have been considered to play a crucial role in the reactivity of supported catalysts [35,36].
In this study, the interaction between Cu and MoOx should also be responsible for the improved activity because it will promote the high dispersion of Cu and produce more Cu-MoOx interfaces as well as more corresponding oxygen vacancies [37]. As stated above, TEM images of Cu/MoOx (1:1) show affluent Cu-MoOx interfaces. Because MoOx modifies the surface of Cu particles through both geometric coverage and electronic modification, XPS results also show that the Cu/MoOx (1:1) catalyst has more oxygen vacancies. Highly dispersed Cu and more oxygen vacancies lead to large active surfaces and more active points for CO2 catalytic hydrogenation. In terms of CO selectivity, the Cu/MoOx (1:2) catalyst shows high selectivity for CO at low temperatures. No impurity products are found in the system, and CO selectivity is up to 100%. In contrast, Cu/MoOx (2:1) shows poor CO selectivity of 97.4%. These results suggest that the reduced MoOx is capable of dissociating CO2, and the low activity is possibly related to the low ability of MoOx for H2 activation and hydrogenation, resulting in a change in the selectivity of the RWGS [17].
According to the literature, there are two reaction mechanisms for the RWGS reaction that have been widely accepted, namely, the redox mechanism and the decomposition intermediate species (carbonate, formate, carbonyl, etc.) mechanism. Based on the previous research, we speculate that the reverse water gas reaction may experience an associative mechanism. The reaction process is as follows: (I) H2 is activated and adsorbed on the surface of Cu (111). (II) CO2 reacts with the oxygen vacancies on Cu/MoOx. (III) H2 dissociates and spills over the H atoms to the intermediate, leading to the formed formats (HCOO*) of intermediate species. (Ⅳ) The formats (HCOO*) dissociate to CO, adsorbed OH, and linear CO (L-CO), which adsorb on Cu. (Ⅴ) The linear CO (L-CO) desorbs to CO, and the formed OH reacts with the H atom to form H2O.
Cu + H2 → 2Cu − H
OVs + CO2+ e → CO2
CO2 + H→HCOO*
2HCOO* →CO + 2OH* + L-CO
L − CO→CO, OH* + H→H2O
A comparison of Cu/MoOx (1:1) with the recently reported catalysts for RGWS is displayed in Table 2. In terms of CO2 conversion, our catalyst shows comparatively better activity than the recently reported catalyst under the same reduction reaction conditions. Although CO selectivity is not the highest, Cu/MoOx (1:1) delivers a competitive selectivity of 99%.

3. Experimental Section

3.1. Catalyst Preparation

The Cu/MoOx catalysts were synthesized by a hydrothermal method, followed by solid-phase sintering. Cu/MoOx precursors were prepared by dissolving stoichiometric amounts of Cu(NO3)2·3H2O and (NH4)6Mo7O24·4H2O in deionized water to form a solution. To assess the influences of catalyst components on the morphology, size, and catalytic performance, the molar ratios of Cu:Mo are 1:1, 1:2, and 2:1, respectively. The resulting solution was transferred into a 100 mL Teflon-lined stainless steel autoclave and sealed and heated at 180 °C for 6 h. The precursors were collected by filtration and dried at 80 ℃ in a vacuum oven. The obtained precursor was calcined at 450 °C for 2 h under flowing 4% H2/Ar mixture gas to make it pre-reduced and obtain a stable morphology.The preparation process of Cu/MoOx catalyst is shown in Figure 8.

3.2. Product Characterization

The X-ray diffraction (XRD, MO3xHF22, MacScience, Tokyo, Japan) data were obtained to determine material structural properties under the conditions of Cu K radiation with a speed of 1 °C min−1 and the scanning range (2θ) from 10.0° to 80.0°. Nitrogen adsorption was measured in a Tristar 3000 analyzer (Micrometrics, Ottawa, ON, Canada) at liquid nitrogen temperature. The samples were pretreated at 300 °C under a vacuum prior to measurements. The specific surface area of the catalysts was calculated using the Brunauer–Emmett–Teller (BET) method. X-ray photoelectron spectroscopy (XPS, Thermo VG Scientific Co., Ltd., Waltham, MA, USA) was conducted to analyze the valence states of elements in Cu/MoOx. The basic characteristics of the catalysts were studied with hydrogen temperature-programmed reduction (H2-TPR). In a typical experiment, 50 mg of samples was loaded in a fixed-bed quartz reactor, then heated to 300 °C at a rate of 10 °C min−1 in an Ar flow (25 mL min−1) for 1 h to clean the sample. The experiment was carried out in a 5% H2/Ar flow of 25.0 mL·min−1, with heating to 700 °C at a heating rate of 10 °C min−1. The hydrogen consumption of samples during the heating process was recorded by gas chromatography equipped with a TCD detector. Scanning electron microscopy (SEM, SIGMA 500, Zeiss, Oberkochen, Germany) and transmission electron microscopy (TEM, JEM 2100F, JEOL, Tokyo, Japan) were utilized to observe the morphology of catalysts.

3.3. Catalytic Evaluation

The catalytic performance of the Cu/MoOx catalyst was evaluated by the CO2 reverse water-gas shift (RWGS) reaction. The RWGS reaction was carried out in a quartz fixed-bed microreactor with an internal diameter of 6 mm under atmospheric pressure. Before the reaction, the catalysts were initially reduced by H2 50 mg as prepared samples were put into the quartz tube, which is placed in the tubular furnace, and then reduced at 400 °C for 2 h using pure H2 (15 mL min−1, atmospheric pressure). Next, the inlet flow was changed to 20 mL min−1 Ar for cooling down to 200 °C, and then the inlet flow was switched to a 50.0 mL min−1 CO2/H2/Ar mixture gas (10/40/50) for CO2 hydrogenation. The temperature range of the catalyst’s activity test is from 200 °C to 400 °C. The concentration of gas products was analyzed online by gas chromatography equipped with a thermal conductivity detector (TCD) and a flammable ionization detector (FID), and the chromatographic column model is TDX-01. The CO2 conversion (XCO2, %), selectivity to CO (Sco, %), and selectivity to CH4 (SCH4, %) were defined following Formulas (1)–(3):
CO 2   Conversion = [ CO 2 ] inlet [ CO 2 ] outlet [ CO 2 ] inlet × 100 %  
CO   Selectivity   = [ CO ] outlet [ CO ] outlet + [ CH 4 ] outlet × 100 %  
CH 4   Selectivity   = [ CH 4 ] outlet [ CO ] outlet + [ CH 4 ] outlet × 100 %  

4. Conclusions

A series of Cu-based catalysts supported on MoOx were synthesized by the hydrothermal method for noble metal-free RWGS reactions. MoOx support enhances the metal-oxide interactions with Cu through electronic modification and geometric coverage. As a result, MoOx support improves the high dispersion of Cu and produces more Cu-MoOx interfaces as well as more corresponding oxygen vacancies. The Cu/MoOx (1:1) displays excellent catalytic performance with 35.85% CO2 conversion and 99% selectivity for CO at 400 °C. The Cu/MoOx as a non-noble metal catalyst shows great potential in the RWGS reaction used as an industrially developed process at a medium-low temperature.

Author Contributions

Data curation, Y.G.; formal analysis, Y.G. and B.Z.; investigation, Y.G. and K.X.; resources, Y.G. and B.Z.; supervision, B.Z.; writing—original draft, Y.G.; writing—review and editing, Y.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Science and Technology Project from the Chongqing Education Commission (Grant No. KJQN2019008832), the Natural Science Foundation Project of Chongqing (Grant No. cstc2020jcyj-msxmX0345 and CSTB2022NSCQ-MSX1230), and the Research Platform Open Project of the Chongqing Technology and Business University (No. KFJJ2019086).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, W.H.; Himeda, Y.; Muckerman, J.T.; Manbeck, G.F.; Fujita, E. CO2 Hydrogenation to Formate and Methanol as an Alternative to Photo- and Electrochemical CO2 Reduction. Chem. Rev. 2015, 115, 12936–12973. [Google Scholar] [CrossRef] [PubMed]
  2. Fu, J.; Jiang, K.; Qiu, X.; Yu, J.; Liu, M. Product selectivity of photocatalytic CO2 reduction reactions. Mater. Today 2020, 32, 222–243. [Google Scholar] [CrossRef]
  3. Wang, X.X.; Duan, Y.H.; Zhang, J.F.; Tan, Y.S. Catalytic Conversion of CO2 into High Value-Added Hydrocarbons over Tandem Catalyst. J. Fuel Chem. Technol. 2022, 50, 538–563. [Google Scholar] [CrossRef]
  4. Chen, J.; Duan, L.; Sun, Z. Review on the development of sorbents for calcium looping. Energy Fuels 2020, 34, 7806–7836. [Google Scholar] [CrossRef]
  5. Leeson, D.; Mac Dowell, N.; Shah, N.; Petit, C.; Fennell, P. A Techno-economic analysis and systematic review of carbon capture and storage (CCS) applied to the iron and steel, cement, oil refining and pulp and paper industries, as well as other high purity sources. Int. J. Greenh. Gas Control 2017, 61, 71–84. [Google Scholar] [CrossRef]
  6. Oshima, K.; Shinagawa, T.; Nogami, Y. Low temperature catalytic reverse water gas shift reaction assisted by an electric field. Catal. Today 2014, 232, 27–32. [Google Scholar] [CrossRef]
  7. Rodriguez, J.A.; Ma, S.; Liu, P.; Hrbek, J.; Evans, J.; Perez, M. Activity of CeOx and TiOx Nanoparticles Grown on Au(111) in the Water-Gas Shift Reaction. Science 2007, 318, 1757–1760. [Google Scholar] [CrossRef] [PubMed]
  8. Chen, X.; Su, X.; Su, H.; Liu, X.; Miao, S.; Zhao, Y.; Sun, K.; Huang, Y.; Zhang, T. Theoretical insights and the corresponding construction of supported metal catalysts for highly selective CO2-to-CO conversion. ACS Catal. 2017, 7, 4613–4620. [Google Scholar] [CrossRef]
  9. Park, S.W.; Joo, O.S.; Jung, K.D.; Kim, H.; Han, S.H. Development of ZnO/Al2O3 catalyst for reverse-water-gas-shift reaction of CAMERE (carbon dioxide hydrogenation to form methanol via a reverse-water-gas-shift reaction) process. Appl. Catal. A 2001, 211, 81–90. [Google Scholar] [CrossRef]
  10. Zonetti, P.C.; Letichevsky, S.; Gaspar, A.B.; Sousa-Aguiar, E.F.; Appel, L.G. The NixCe0.75Zr0.25-xO2 solid solution and the RWGS. Appl. Catal. A 2014, 475, 48–54. [Google Scholar] [CrossRef]
  11. Silva-Calpa, L.D.R.; Zonetti, P.C.; Rodrigues, C.P.; Alves, O.C.; Appel, L.G.; Avillez, R.R.de. The ZnxZr1-xO2-y solid solution on m-ZrO2: Creating O vacancies and improving the m-ZrO2 redox properties. J. Mol. Catal. A Chem. 2016, 425, 166–173. [Google Scholar] [CrossRef]
  12. Rodriguez, J.A.; Liu, P.; Stacchiola, D.J.; Senanayake, S.D.; White, M.G.; Chen, J.G. Hydrogenation of CO2 to Methanol: Importance of Metal-Oxide and Metal-Carbide Interfaces in the Activation of CO2. ACS Catal. 2015, 5, 6696–6706. [Google Scholar] [CrossRef]
  13. Andrés, A.G.B.; Octavio, J.F.; Dario, J.S.; Karim, S.; Marcelo, S.N. Porous MoxCy/SiO2 Material for CO2 Hydrogenation. Top. Catal. 2019, 62, 1026–1034. [Google Scholar]
  14. Zhang, X.; Zhu, X.; Lin, L.; Yao, S.; Zhang, M.; Liu, X.; Wang, X.; Li, Y.; Shi, C.; Ma, D. Highly Dispersed Copper over β-Mo2C as an Efficient and Stable Catalyst for the Reverse Water Gas Shift (RWGS) Reaction. ACS Catal. 2017, 7, 912–918. [Google Scholar] [CrossRef]
  15. Su, H.; Ye, Y.; Lee, K.J.; Zeng, J.; Mun, B.S.; Crumlin, E.J. Probing the surface chemistry for reverse water gas shift reaction on Pt (1 1 1) using ambient pressure X-ray photoelectron spectroscopy. J. Catal. 2020, 391, 123–131. [Google Scholar] [CrossRef]
  16. Xu, J.J.; Gong, X.X.; Hu, R.R.; Liu, Z.W.; Liu, Z.T. Highly active K-promoted Cu/β-Mo2C catalysts for reverse water gas shift reaction: Effect of potassium. Mol. Catal. 2021, 516, 111954. [Google Scholar] [CrossRef]
  17. Zhang, R.Y.; Wei, A.L.; Zhu, M.; Wu, X.X.; Wang, H.; Zhu, X.L.; Ge, Q.F. Tuning reverse water gas shift and methanation reactions during CO2 reduction on Ni catalysts via surface modification by MoOx. J. CO2 Util. 2021, 52, 101678. [Google Scholar] [CrossRef]
  18. Noelia, L.G.; Jannier, C.; Jose, M.P. Graphene-TLL-Cu2ONPs Hybrid as Highly Effificient Catalyst for Degradation of Organic Compounds. Nanomaterials 2023, 13, 449. [Google Scholar]
  19. Ramyakrishna, P.; Prathap, C.; Rajendiran, R.; Rajender, B.; Ravi, B.; Putrakumar, B.; Vijayanand, P.; Ahmed, B.R.; Aboubakr, M.A.; Noora, A.Q. Vapour-Phase Selective Hydrogenation of γ-Valerolactone to 2-Methyltetrahydrofuran Biofuel over Silica-Supported Copper Catalysts. Nanomaterials 2022, 12, 3414. [Google Scholar]
  20. Gonzalez-Castano, M.; Navarro de Miguel, J.C.; Sinha, F.; Ghomsi, W.S.; Klepel, H. Cu supported Fe-SiO2 nanocomposites for reverse water gas shift reaction. J. CO2 Util. 2021, 46, 101493. [Google Scholar] [CrossRef]
  21. Chen, L.; Wu, D.; Wang, C.; Ji, M.; Wu, Z. Study on Cu-Fe/CeO2 bimetallic catalyst for reverse water gas shift reaction. J. Environ. Chem. Eng. 2021, 9, 105183. [Google Scholar] [CrossRef]
  22. Yang, L.; Pastor-Perez, L.; Villora-Pico, J.J.; Sepúlveda-Escribano, A.; Tian, F.; Zhu, M. Highly Active and Selective Multicomponent Fe-Cu/CeO2-Al2O3Catalysts for CO2 Upgrading via RWGS: Impact of Fe/Cu Ratio. ACS Sustain. Chem. Eng. 2021, 9, 101493. [Google Scholar] [CrossRef]
  23. Zhang, B.W.; Cao, L.C.; Tang, C.; Tan, C.H.; Cheng, N.Y.; Lai, W.H.; Wang, Y.X.; Cheng, Z.X.; Dong, J.C.; Kong, Y.; et al. Atomically Dispersed Dual-Site Cathode with a Record High Sulfur Mass Loading for High-Performance Room-Temperature Sodium–Sulfur Batteries. Adv. Mater. 2023, 35, 2206828. [Google Scholar] [CrossRef] [PubMed]
  24. Song, J.M.; Ni, X.M.; Zhang, D.E. Fabrication and photoluminescence properties of hexagonal MoO3 rods. Solid State Sci. 2006, 8, 1164–1167. [Google Scholar] [CrossRef]
  25. Qi, T.Q.J.; Li, W.Z.; Li, H.; Jia, K.; Chen, S.Y.; Zhang, Y.C. Yttria-doped Cu/ZnO catalyst with excellent performance for CO2 hydrogenation to methanol. Mol. Catal. 2021, 509, 111641. [Google Scholar] [CrossRef]
  26. Jiang, L.; Zhu, H.W.; Razzaq, R.; Zhu, M.L.; Li, C.S.; Li, Z.X. Effect of zirconium addition on the structure and properties of CuO/ CeO2 catalysts for high-temperature wateregas shift in an IGCC system. Int. J. Hydrog. Energy 2012, 37, 15914–15924. [Google Scholar] [CrossRef]
  27. Liu, Y.J.; Li, Z.F.; Xu, H.B.; Han, Y.Y. Reverse wateregas shift reaction over ceria nanocube synthesized by hydrothermal method. Catal. Commun. 2016, 76, 1–6. [Google Scholar] [CrossRef]
  28. Gao, D.; Yang, G.; Li, J.; Zhang, J.; Xue, D. Room-Temperature Ferromagnetism of Flowerlike CuO Nanostructures. J. Phys. Chem. C 2010, 114, 18347–18351. [Google Scholar] [CrossRef]
  29. Hoch, L.B.; Wood, T.E.; O’Brien, P.G.; Liao, K.; Reyes, C.A.M.; Ozin, G.A. Photomethanation of Gaseous CO2 over Ru/Silicon Nanowire Catalysts with Visible and Near-Infrared Photons. Adv. Sci. 2014, 1, 1400001. [Google Scholar]
  30. Xiong, K.; Zhou, G.L.; Zhang, H.D.; Shen, Y.; Zhang, X.M.; Zhang, Y.H.; Li, J.L. Bridging Mo2C-C and highly dispersed copper by incorporating N-functional groups to greatly enhance catalytic activity and durability for the carbon dioxide hydrogenation. J. Mater. Chem. A 2018, 6, 15510–15516. [Google Scholar] [CrossRef]
  31. Wang, L.C.; TahvildarKhazaneh, M.; Widmann, D.; Behm, R.J. TAP reactor studies of the oxidizing capability of CO2 on a Au/CeO2 catalyst-A first step toward identifying a redox mechanism in the Reverse Water-Gas Shift reaction. J. Catal. 2013, 302, 20–30. [Google Scholar] [CrossRef]
  32. Midya, A.; Ghorai, A.; Mukherjee, S.; Maiti, R.; Ray, S.K. Hydrothermal growth of few layer 2H-MoS2 for heterojunction photodetector and visible light induced photocatalytic applications. J. Mater. Chem. A 2016, 4, 4534–4543. [Google Scholar] [CrossRef]
  33. Pan, X.N.; Xi, B.J.; Lu, H.B.; Zhang, Z.C.; An, X.G.; Liu, J.; Feng, J.K. Molybdenum Oxynitride Atomic Nanoclusters Bonded in Nanosheets of N-Doped Carbon Hierarchical Microspheres for Effcient Sodium Storage. Nano-Micro Lett. 2022, 10, 154–169. [Google Scholar]
  34. Zhang, S.; Zeng, Y.W.; Wang, Z.T.; Zhao, J.; Dong, G.D. Glycerol-controlled synthesis of MoS2 hierarchical architectures with well-tailored subunits and enhanced electrochemical performance for lithium ion batteries. Chem. Eng. J 2018, 334, 487–496. [Google Scholar] [CrossRef]
  35. Gao, P.; Li, F.; Zhan, H.; Zhao, N.; Xiao, F.; Wei, W.; Zhong, L.; Wang, H.; Sun, Y. 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]
  36. Wang, L.; Fang, W.; Xiao, F.S. NbOPO4 supported Rh nanoparticles with strong metal-support interactions for selective CO2 hydrogenation. ChemSusChem 2020, 13, 6300–6306. [Google Scholar]
  37. Zhou, Z.; Kong, Y.; Tan, H.; Huang, Q.W.; Wang, C.; Pei, Z.X.; Wang, H.Z.; Liu, Y.Y.; Wang, Y.H.; Li, S.; et al. Cation-Vacancy-Enriched Nickel Phosphide for Efficient Electrosynthesis of Hydrogen Peroxides. Adv. Mater. 2022, 34, 2106541. [Google Scholar] [CrossRef]
  38. Bobadilla, L.F.; Santos, J.L.; Ivanova, S.; Odriozola, J.A.; Urakawa, A. Unravelling the Role of Oxygen Vacancies in the Mechanism of the Reverse Water–Gas Shift Reaction by Operando DRIFTS and Ultraviolet–Visible Spectroscopy. ACS Catal. 2018, 8, 7455–7467. [Google Scholar] [CrossRef]
  39. Liu, Y.X.; Li, L.W.; Zhang, R.Y.; Guo, Y.H.; Wang, H.; Qingfeng Ge, Q.F.; Zhu, X.L. Synergetic enhancement of activity and selectivity for reverse water gas shift reaction on Pt-Re/SiO2 catalysts. J. CO2 Util. 2022, 63, 02128. [Google Scholar] [CrossRef]
  40. Da, B.; Zhou, G.; Ge, S.; Xie, H.; Jiao, Z.; Zhang, G. CO2 reverse water-gas shift reaction on mesoporous M-CeO2 catalysts. Can. J. Chem. Eng. 2017, 95, 634–642. [Google Scholar] [CrossRef]
  41. Tarifa, P.; Gonzalez-Castano, M.; Cazana, F.; Arellano-García, A.M.H. Development of one-pot Cu/cellulose derived carbon catalysts for RWGS reaction. Fuel 2022, 319, 123707. [Google Scholar] [CrossRef]
  42. Yang, L.; Pastor-Pérez, L.; Villora-Pico, J.J.; Gu, S.; Sepúlveda-Escribano, A.; Reina, T.R. CO2 valorisation via reverse water-gas shift reaction using promoted Fe/CeO2-Al2O3 catalysts: Showcasing the potential of advanced catalysts to explore new processes design. Appl. Catal. A-Gen. 2020, 593, 117442. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of Cu/MoOx samples pre-reduced at 450 °C for 2 h.
Figure 1. XRD patterns of Cu/MoOx samples pre-reduced at 450 °C for 2 h.
Catalysts 13 00684 g001
Figure 2. SEM images of Cu/MoOx samples with different atomic ratios of Cu/Mo (a,d) 1:2, (b,e) 1:1, and (c,f) 2:1.
Figure 2. SEM images of Cu/MoOx samples with different atomic ratios of Cu/Mo (a,d) 1:2, (b,e) 1:1, and (c,f) 2:1.
Catalysts 13 00684 g002
Figure 3. The TEM (a) and HRTEM (b) images of Cu/MoOx (1:1) catalyst.
Figure 3. The TEM (a) and HRTEM (b) images of Cu/MoOx (1:1) catalyst.
Catalysts 13 00684 g003
Figure 4. H2−TPR profiles of pre-reduced Cu/MoOx and MoOx at 450 °C under the flow of 4% H2/Ar mixture gas for 2 h.
Figure 4. H2−TPR profiles of pre-reduced Cu/MoOx and MoOx at 450 °C under the flow of 4% H2/Ar mixture gas for 2 h.
Catalysts 13 00684 g004
Figure 5. (a) Survey X-ray photoelectron spectra for Cu/MoOx as well as the core level of (b) O1s spectra.
Figure 5. (a) Survey X-ray photoelectron spectra for Cu/MoOx as well as the core level of (b) O1s spectra.
Catalysts 13 00684 g005
Figure 6. The deconvoluted peaks of Mo3d in Cu/MoOx samples.
Figure 6. The deconvoluted peaks of Mo3d in Cu/MoOx samples.
Catalysts 13 00684 g006
Figure 7. CO2 conversion, CO selectivity, and CH4 selectivity for Cu/MoOx catalysts in the RWGS process as a function of temperature.
Figure 7. CO2 conversion, CO selectivity, and CH4 selectivity for Cu/MoOx catalysts in the RWGS process as a function of temperature.
Catalysts 13 00684 g007
Figure 8. The pattern of Cu/MoOx catalyst preparation process.
Figure 8. The pattern of Cu/MoOx catalyst preparation process.
Catalysts 13 00684 g008
Table 1. Atomic composition by XPS, surface area, and pore size distribution.
Table 1. Atomic composition by XPS, surface area, and pore size distribution.
CatalystsSurface Composition by XPS (at%)Physicochemical Properties
CuMoOCSBET (m2 g−1)XPd (nm)
Cu/MoOx (1:2)7.1712.0648.9931.77146.011.30
Cu/MoOx (1:1)6.285.8336.451.48156.79.32
Cu/MoOx (2:1)4.282.9931.9160.81117.517.02
Table 2. Cu/MoOx (1:2) catalyst performance comparison with the recent reports in the literature.
Table 2. Cu/MoOx (1:2) catalyst performance comparison with the recent reports in the literature.
CatalystsH2:CO2Temperature (°C)CO2 ConversionCO SelectivityRef.
Cu/MoOx (1:2)440035.9%99.0%This work
1K-Cu/β-Mo2C440024.8%99.5%[16]
Mo/β-Mo2C44003.8%96.5%[16]
Au/Al2O3440011.0%100%[38]
Pt/SiO2440012.1%100%[39]
Pt-O.5Re/SiO2440024.3 %97.3%[39]
Cu/CeO2440031%100%[40]
Cu-Ce/CDC440024.2%98.4%[3]
Ni-1Mo440023.6%93.8%[41]
FeCu/CeAl440023.6%96.7%[42]
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

Gao, Y.; Xiong, K.; Zhu, B. Design of Cu/MoOx for CO2 Reduction via Reverse Water Gas Shift Reaction. Catalysts 2023, 13, 684. https://doi.org/10.3390/catal13040684

AMA Style

Gao Y, Xiong K, Zhu B. Design of Cu/MoOx for CO2 Reduction via Reverse Water Gas Shift Reaction. Catalysts. 2023; 13(4):684. https://doi.org/10.3390/catal13040684

Chicago/Turabian Style

Gao, Yuan, Kun Xiong, and Bingfeng Zhu. 2023. "Design of Cu/MoOx for CO2 Reduction via Reverse Water Gas Shift Reaction" Catalysts 13, no. 4: 684. https://doi.org/10.3390/catal13040684

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