Next Article in Journal
Overview of Research Status and Development Trends in Diesel Desulfurization Technology
Previous Article in Journal
Photocatalytic Cement Mortar with Durable Self-Cleaning Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation of Ternary CuZnM (M = Cr, Ce, Zr, Al) Catalysts in CO2 Hydrogenation for Methanol Synthesis

1
State Key Laboratory of Clean and Efficient Coal Utilization, College of Chemistry and Chemical Engineering, Taiyuan University of Technology, Taiyuan 030024, China
2
Faculty of Light Industry and Chemical Engineering, Dalian Polytechnic University, Dalian 116034, China
3
Shanxi Dongyi Coal Electric Aluminum Group Co., Ltd., Lvliang 033000, China
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(3), 250; https://doi.org/10.3390/catal15030250
Submission received: 7 February 2025 / Revised: 28 February 2025 / Accepted: 4 March 2025 / Published: 6 March 2025
(This article belongs to the Section Catalysis for Sustainable Energy)

Abstract

:
The hydrogenation of CO2 to methanol over Cu-based catalysts is one of the attractive routes to reduce greenhouse gas emissions and generate high-value-added chemicals. Industrial CuZnAl catalysts possess some shortcomings, but various promoters can enhance the activity and durability of Cu-based catalysts for CO2 hydrogenation. Herein, we systematically investigated the variations in the physicochemical properties of ternary CuZnM (M: Cr, Ce, Zr, Al) catalysts induced by different promoters, as well as their impact on CO2 hydrogenation to methanol. The results demonstrate that the catalytic activity followed the order of CZCr > CZCe > CZZr > CZAl, with CZCr exhibiting the highest stability among them. Combined with XRD, SEM, TEM, H2-TPR, TPD, XPS, Raman findings and the experimental results, the smaller Cu particle size was conducive to increasing the CH3OH yield, and the lower Cu+/Cu0 ratio exhibited better stability in CO2 hydrogenation to methanol reaction. This approach offers theoretical insights and practical solutions for the industrial application of ternary Cu-based catalysts.

Graphical Abstract

1. Introduction

The rapid consumption of fossil fuels currently releases a large amount of the greenhouse gas carbon dioxide (CO2) [1]. The resource utilization of carbon dioxide has always been the focus of research, particularly in the production of high-value chemicals such as methanol [2,3] and methane [4,5], which are also premium renewable energy sources. Developing a methanol production pathway from CO2, where methanol can complement and partially act as a substitute for petroleum-based fuels, may address energy scarcity and improve resource efficiency [6]. The catalytic coupling of CO2 and green hydrogen to produce methanol can not only reduce greenhouse gas emissions but also provide a sustainable energy path [7].
At present, the efficiency of CO2 hydrogenation to methanol predominantly relies on the catalyst activity, and the development of a highly efficient catalyst has been the focus of researchers, including Cu-based catalysts, precious metal catalysts, other metal oxide catalysts, etc. [8]. Among them, the ternary Cu-based catalyst has been widely used because of its excellent catalytic performance. Such catalysts are usually composed of copper, zinc and other metallic elements and can efficiently promote the activation and hydrogenation of CO2. Copper is often considered to be the main active phase, with Zn as a carrier or structural barrier and a third element as an accelerator [9]. In the past, CuZnAl catalysts have gained prominence in methanol synthesis due to their cost efficiency and exceptional catalytic performance, with a substantial body of literature dedicated to exploring doped Cu-based catalysts enhanced by various promoters [10,11,12]. However, some Cu-based catalysts can still face aggregation or re-oxidation during catalysis, leading to a decrease in active sites and consequently impacting their catalytic performance and long-term stability [13]. In order to surpass the aforementioned challenges in CO2 hydrogenation for methanol production, an effective tactic to enhance the efficacy of Cu/Zn-based catalysts involves the incorporation of a promoter, that is, the third element. The function of the third element in the catalyst is multifaceted, including but not limited to enhancing catalyst dispersion, strengthening metal–support interactions, modulating the electronic configuration and fine-tuning the reaction pathway. Some transition metal oxides such as Fe2O3, Ga2O3, ZrO2, CeO2, In2O3, Cr2O3 and Al2O3 are often selected as promoters [14,15]. Singh et al. revealed the accelerating effect of the CeO2 precursor in Cu/ZnO catalysts and established a relationship between the density of the strongly alkaline sites and the active H adsorption sites [16]. Zhan et al. developed a range of Cu/Zn/Zr catalysts with varying Zr/(Cu+Zn) ratios via a precipitation method and evaluated their performance in methanol synthesis from a CO2/H2 mixture, highlighting the reaction’s structural sensitivity [17]. Additionally, Chang et al. optimized the structure and surface properties of a Cu/ZnO/ZrO2 catalyst by adjusting the ratio of ZnO and ZrO2, which significantly improved the efficiency and selectivity of CO2 hydrogenation to methanol [15]. Xiong et al. synthesized Cu/ZnCr catalysts via the incipient wetness impregnation technique and investigated the impact of the Zn/Cr ratio on methanol selectivity; the catalyst with an optimal Zn/Cr ratio exhibited a methanol space–time yield of 4.7 mmol/gcat./h (XCO2 = 25.1%; SMeOH = 31.1%) at 300 °C and 2 MPa [18]. Hengne et al. synthesized copper (Cu) nanocomposite catalysts doped with gallium (Ga) and aluminum (Al) through a synchronous co-precipitation, assessing their efficacy in CO2 hydrogenation to methanol. The incorporation of Ga was found to strengthen the catalyst’s weak base sites to a greater extent than the addition of Al [19].
Nevertheless, these advancements have yet to be comprehensively investigated to elucidate the promotional mechanisms underlying ternary Cu-based catalyst systems and to attain enhanced methanol yields alongside extended operational longevity in industrial settings. For example, the incorporation of ZrO2 could stabilize the intermediates and reduce the energy barrier pathway by promoting the formation and conversion of formic acid intermediates [20]. Sun et al. demonstrated that Erbium (Er) doping could refine the ZnO particle size, consequently enlarging the Cu-ZnO interfacial area and fostering a robust electron metal–support interaction (EMSI) between Cu and ZnO. This enhancement was believed to expedite the activation of CO2 adsorbates and reaction intermediates [21].
To overcome the constraints associated with Cu-based catalysts, an in-depth study was undertaken to examine the effects of Al, Ce, Zr and Cr additives on the physicochemical properties and catalytic performance in the hydrogenation of CO2 to methanol. These four elements were the most frequently used in Cu-based catalysts, but they were not systematically studied as ternary Cu-based catalyst systems to improve catalytic efficiency more comprehensively. This study systematically assessed the efficacy of ternary Cu-based catalysts for CO2 hydrogenation to methanol and scrutinized the impact of various dopants on the catalysts’ physicochemical features using a suite of analytical methods, including XRD, SEM, TEM, H2-TPR, XPS, TPD and Raman spectroscopy. The findings revealed that the CuZnCr catalyst exhibited superior activity and stability in comparison with other catalysts, which was ascribed to the electronic interaction from Cu and Cr. Meanwhile, Ce, Zr and Al also played different roles in the system, such as particle size, reduction degree, adsorption capacity, etc. It might provide theoretical guidance for designing Cu-based catalysts with higher energy conversion efficiency in industrial applications.

2. Results

2.1. Crystal Structure Determined by XRD

In order to study the crystalline structure of ternary CZM catalysts and the effect of grain size of Cu species, XRD patterns of various catalysts are shown in Figure 1. All calcined catalysts exhibited characteristic CuO diffraction peaks at 2θ = 36.5° and 38.8°, as illustrated in Figure 1a, corresponding to the (002) and (111) crystal planes within the layered structure. In comparison, the CuO diffraction peaks of CZAl catalysts were sharper, indicating higher crystallinity and larger grain size. This suggested that the doping of different elements could greatly affect the grain sizes of copper species. According to Scherrer’s equation, the calculated mean CuO grain size was less than 20 nm. On the other hand, ZnO diffraction peaks located at ca. 32.2°, 38.8° and 66.8° were detected in all cases. However, there was no third element detected because of their high dispersion or amorphous property, and Raman spectroscopy (Figure S1) also failed to detect the vibration peak of the third element. Detailed descriptions of the Raman spectra can be found in the Supporting Information [22]. Their presence was demonstrated by EDS images, including Al, Ce, Zr and Cr (Figure S2). To further clarify the presence and content of each component in the catalyst sample, we characterized the sample using an X-ray fluorescence spectroscopy (XRF) technique. As shown in Table 1, all target elements were effectively detected in the samples, and their composition was in excellent agreement with the theoretical design values. This indicates that the promoter elements were successfully incorporated into the catalyst system, and the metal nitrates were almost completely precipitated. The XRD spectrum of the ternary CZM catalysts in the reduced state is shown in Figure 1b. It can be seen that the distinct diffraction peaks of reduced-state Cu2O and Cu appeared at 36.4° and 42.7°, respectively. From the diffraction peak patterns, the states and crystallinity of the copper species were different, which indicated that the CZM catalysts were in different reduced states influenced by different additives. Furthermore, no distinct CuO diffraction peaks were detected in either the CZCe or CZCr catalysts, indicating that they predominantly exist as Cu2O and Cu, suggesting a highly reduced state.

2.2. Catalyst Surface Morphology and Particle Size

The morphological features of the ternary CZM catalysts were observed by SEM images as shown in Figure 2. The presence of amorphous flocculent structural species was detected in all samples, and was most pronounced in the CZZr catalyst. As seen in Figure 2A, a large number of cubic block particles can be observed in the CZAl catalyst, which may be related to Cu species [23,24]. There were many irregular spherical particles dominated in the CZCe catalyst, which was the common morphological structure in co-precipitation synthesis due to the existence of multiple crystal growth mechanisms, as shown in Figure 2B [25]. The presence of Zr caused the CZZr catalyst to exhibit more irregularities that the specific composition of Cu, Zn and Zr, which may lead to the development of amorphous structures [26]. Although the morphology of the CZCr catalyst in Figure 2D resembled that of the CZCe catalyst, the particles of the CZCr catalyst were notably smaller and more dispersed. The incorporation of Cr likely altered the lattice parameters of the catalyst, leading to a reduction in particle size and a more homogeneous distribution of Cu particles [27].
In order to determine the effect of each element on the particle size of the CZM catalysts, the particle size distribution statistics were carried out by TEM characterization, as shown in Figure 3. Firstly, it can be seen from the TEM images that the general morphology of each catalyst was similar to the results described by SEM images. For example, the CZAl catalyst in Figure 3a exhibits some cubic morphological structures, and the CZCr catalyst in Figure 3d shows smaller spherical particles. By counting the particle size of each catalyst, it was observed that the CZCr catalyst had a smaller particle size (11.1 nm) compared to the others. This suggested that the addition of Cr was more favorable for the formation of catalysts with smaller and more dispersed particles. The CZCe catalyst, also in granular form, had a particle size of about 14.3 nm and was exposed to the same CuO (110) crystal surface as the CZCr catalyst. However, the presence of CeO2 actually promoted the dispersion of Cu nanoparticles to form smaller particle sizes compared to the CZAl catalyst (19.3 nm) [28]. On the other hand, m-ZrO2 (111) was detected in the CZZr catalyst based on the lattice stripe spacing, which is one of the main reasons for the amorphous structure of the catalyst [29]. Table 2 shows the statistical particle sizes of all samples via XRD and TEM. It can be seen that the overall particle size calculated by the Scherrer formula was slightly smaller than that obtained by TEM image statistics, but the trend was consistent in that the particle size decreased as follows: CZCr < CZCe < CZZr < CZAl. Furthermore, the average particle size calculated from N2O chemisorption data aligns well with the results derived from XRD and TEM characterization, as summarized in Table 2.

2.3. Catalyst Reduction Behavior by H2-TPR

To study the reduction property of ternary CZM catalysts, TPR characterization tests were performed, as shown in Figure 4. All samples exhibited two reduction peaks at below 200 °C. This suggests that the process occurred predominantly with the reduction of Cu species, as the reduction of ZnO tended to occur at above 300 °C. It was reported that the reduction of CuO was easier than the reduction of Cu2O; the apparent activation energy for the reduction of CuO was about 14.5 kcal/mol, while the value was 27.4 kcal/mol for Cu2O reduction [30]. Specifically, all samples showed two major reduction peaks, labeled L and H. The L-peak was usually associated with the reduction process at lower temperatures, which may be related to the reduction of oxides on the catalyst surface or in smaller particles (dispersed CuOx species). The H-peak was related to the reduction process at higher temperatures, which may be related to larger particles (bulk CuOx species) or oxides that are difficult to reduce [31,32,33,34,35]. The CZCr catalyst showed an obvious reduction peak, indicating that the doping of Cr may promote the reduction of Cu oxide. On the contrary, the doping of Al introduced a complex reduction mechanism. In order to further understand the reducibility of the as-prepared catalyst, the fitted TPR curves were back-convolutionally resolved, as listed in Table 3. It is evident that varying promoters may induce differences in reduction temperatures, consequently influencing the extent of reduction in the ternary CZM catalysts. For instance, the two reduction peaks of the CZCr catalyst were similar compared to the CZAl catalyst, which was related to the interaction between the additives and the copper species. Meanwhile, an increase in the concentration of the high-temperature peak also indicated that more Cu2O was reduced to Cu.

2.4. Surface Electronic State of Elements on the Catalysts

In order to investigate the surface electronic properties of the ternary CZM catalysts, XPS characterization was carried out, as shown in Figure 5. The ternary catalysts before and after the reaction were first characterized for Cu 2p orbitals. It can be seen from Figure 5a that the electronic states of the surface Cu species were similar for each catalyst before the reaction, with Cu 2p1/2 located near 954 eV and Cu 2p2/3 located at near 934 eV. On the other hand, the satellite peaks were observed, indicating that each catalyst was in a highly oxidized state [36]. However, the reduced catalysts exhibited different surface electronic states, as shown in Figure 5b. The reduced CZAl and CZZr catalysts still retained the satellite peaks, but the 2p orbitals shifted to higher binding energies, indicating the gradual transformation of Cu2+ into Cu0 and Cu+. Meanwhile, the CZCe and CZCr catalysts showed strong 2p peaks and the satellite peaks disappeared, suggesting an increase in the content of Cu+ and Cu0 [37]. Considering the similar binding energies of Cu+ and Cu0, Cu LMM XAES spectroscopic characterization was chosen to further determine the relative contents of Cu+ and Cu0. Only the Cu2+ wave peak located at ca. 917.7 eV was detected in all cases, as presented in Figure 5c. However, the active components (Cu+ and Cu0) of the reduced ternary CZM catalysts changed significantly, as shown in Figure 5d. The contents and proportions of Cu+ and Cu0 in ternary CZM catalysts are listed in Table 4. It can be seen that a significant difference appeared in the distribution of active Cu species among the four catalysts, with the CZAl catalyst having a higher relative Cu+ content, followed by the CZCe and CZZr catalysts, and with the CZCr catalyst having a relative lower Cu+ content, even lower than that of Cu0. The Cu+/Cu0 ratios were as follows: CZAl (3.05), CZCe (2.01), CZZr (1.76) and CZCr (0.88). Hence, it was favorable to investigate the electronic states of each promoter (Figure 6).
In order to investigate the reasons leading to the different distributions of active Cu species (Cu+ and Cu0) in the ternary CZM catalysts, the third elemental orbitals were characterized and analyzed by XPS, and the results are shown in Figure 6, in which the upper curves are before reduction and the lower curves are after reduction. It is obvious that no significant change occurred for the Al 2p orbitals of the CZAl catalysts and the Zr 3d orbital profiles of the CZZr catalysts before and after reduction. In the Al 2p XPS spectra of the CZAl catalyst, the observed peaks are mainly related to the chemical state of the aluminum species. The peak with a binding energy of 74.4 eV corresponds to the presence of AlOOH species, and in addition, the shoulder peak at 77.0 eV is attributed to Cu 3p1/2, implying the presence of Cu2O [38,39,40]. This result depicts that the aluminum species in the catalysts are mainly in the form of aluminum hydroxide (AlOOH) and may have some association with Cu+. In the Zr 3d XPS profile of the CZZr catalyst, the binding energy ranges of Zr 3d3/2 and Zr 3d5/2 appear in the range of 180~182.5 eV and 182.5~186 eV, respectively, which suggests that zirconium elements were mainly present in the valence form of Zr4+ [41]. However, the binding energy of the reduced CZZr catalyst Zr 3d gradually shifts to a lower binding energy, indicating the existence of electron transfer from Cu to Zr and an enhanced interaction between Cu and ZrO2 [42]. In contrast, the Ce 3d orbitals of the CZCe catalysts and the Cr 2p orbitals of the CZCr catalysts changed significantly before and after reduction. Among them, a 10-peak model was fitted to the Ce 3d orbitals, which originated from different Ce oxidation states (Ce3+ and Ce4+) and their 4f electronic configurations [43,44]. From Figure 6b, it can see that the u’ and v’ peaks at 904.2 eV and 885.2 eV represent Ce3+ species, while the Ce4+ peaks are the u0 peaks at 900.5 eV and the v peaks at 882.0 eV. The Ce3+ content of CZCe catalyst after reduction obviously increased, and the conversion from Ce4+ to Ce3+ might have accelerated the generation of oxygen vacancy [45]. On the other hand, chromium oxides mainly existed as Cr(III) and Cr(VI) valence states according to the Cr 2p XPS spectra of the CuZnCr catalyst. The 2p1/2 and 2p3/2 binding energies of Cr(III) appear near 586.5 eV and 576.7 eV, while the 2p1/2 and 2p3/2 binding energies of Cr(VI) are present near 588.7 eV and 579.4 eV, respectively. Cr(VI) usually has a high electrophilicity and can interact with molecules containing lone pair electrons such as CO2 to promote the activation and hydrogenation of CO2 to methanol. Meanwhile, Cr(III) may participate in the reduction process of the catalyst by reacting with hydrogen to generate Cr(II), thus promoting hydrogen activation on the catalyst surface [46]. The presence of Cr(III) on the catalyst’s surface can also affect its acid–base properties and the distribution of active sites, which in turn affect the catalytic performance. It is evident that the peak corresponding to Cr(VI) diminished, while the peak associated with Cr(III) intensified following the reduction process. Studies have indicated that the synergistic interaction between Cr(III) and Cr(VI) species modulates the properties of the catalysts, with Cr(III) stabilizing the catalyst structure and Cr(VI) imparting robust oxidative capabilities [47].

2.5. Desorption Behavior of Reactants on Catalyst Surfaces

To assess the adsorption and activation capacity of the reactants CO2 and H2 on ternary CZM catalysts, TPD characterization tests were conducted, and the results are shown in Figure 7. CO2-TPD mainly detected two peaks with a low-temperature, weak alkaline adsorption peak near 200 °C and a high-temperature, strong alkaline adsorption peak at between 400 and 500 °C, respectively. The weak base sites were related to surface hydroxyl (-OH) species, while strong basic sites were attributed to low-coordination O2− or oxygen defects [48]. In particular, it was obviously noted that the CZCe catalyst exhibited a strong high-temperature desorption peak, which was related to the dense oxygen vacancy. When oxygen vacancies are present, hydrogen atoms may be more inclined to bind to them, forming unstable hydrogen–oxygen complexes whose desorption requires higher energy [49]. The desorption peak intensities of the other three catalysts were comparable; however, the desorption peak temperature of the CZCr catalyst was notably lower. This may mean that the advantages of Cr-doped catalysts in CO2 adsorption are not obvious. In contrast, for the desorption curve (right) of the other reactant, H2, the medium–low temperature desorption peak near 300 °C is almost the only peak that can be observed, which corresponded to the hydrogen adsorption on the surfaces of Cu+ and Cu0, or the interaction between the metal and the oxide. It can be seen that the desorption peak of the CZAl catalyst was small, indicating that the adsorption and dissociation ability of the CZAl catalyst was weak. It was reported that the Al3+ species replaced Zn in the ZnO lattice and promoted the formation of oxygen vacancy ( V O ¨ ) on the oxide surface [50]. It has been proved that V O ¨ has a positive effect on the adsorption of H2, just like CZCe catalyst, whose H2 desorption peak is not only lower in strength but also higher in energy barrier than other catalysts. The reason why the CZZr catalyst presented a large H2 desorption peak area might be that Zr has good H2 adsorption and dissociation ability [51]. Finally, the CZCr catalyst also maintained a low desorption energy barrier for H2, which indicated that Cr may affect the adsorption energy of hydrogen by changing the electronic properties of Cu in the catalyst.

2.6. Evaluation Performance of Catalyst

Figure 8 showed the performance evaluation results of the ternary CZM catalyst. Firstly, the catalyst activity for CO2 hydrogenation to methanol at different reaction temperatures was investigated. As depicted in Figure 8a, the CO2 conversion rates for all samples exhibit an upward trend with increasing temperature within the range of 200–260 °C. In addition, the CO2 conversions of all samples were above 10%, which fully reflected the high efficiency of the ternary CZM catalyst. When the full activation temperature is 240 °C, the CO2 conversion can reach about 20%. In contrast, the selectivity of CH3OH declined with increasing temperature due to the exothermic nature of the reaction and competition from the by-product CO, as shown in Figure 8b. At low temperatures (200 °C), the highest selectivity of CH3OH was about 65% over the CZCr catalyst. However, whether it was CO2 conversion or CH3OH selectivity, the activity order under this reaction condition decreased as follows: CZCr > CZCe > CZZr > CZAl catalysts. This finding suggests that Cr played a greater role in the ternary CZM catalyst for CO2 hydrogenation to methanol. Subsequently, the stability test of each catalyst was conducted for 100 h, and the results are shown in Figure 8c,d. It can be seen that although the activity of each catalyst in the reaction process still maintained the previous results, the stability change trend was different. After 100 h reaction, the decline rate of CO2 conversion and CH3OH selectivity was in the order of CZCr > CZZr > CZCe > CZAl. Additionally, the CH3OH yields and deactivation rates of all catalysts after 1 h and 100 h are presented in Figure 8e,f. As illustrated in Figure 8e, the CH3OH yield of the CZAl catalyst significantly declined from 130.68 mg·gcat.−1·h−1 to 82.40 mg·gcat.−1·h−1 after 100 h of reaction, with a deactivation rate of 36.9% (Figure 8f), indicating a relatively severe degradation. This is followed by the CZCe catalyst (21.1% deactivation), the CZZr catalyst (11.6% deactivation) and finally the CZCr catalyst deactivation rate, which was the lowest at only 3.5%. In summary, under the reaction conditions, the CO2 conversion, CH3OH selectivity and stability of the CZCr catalyst were superior to those of the other three catalysts. To more accurately evaluate the performance of CZM catalysts, their turnover frequencies (TOFs) were evaluated, as shown in Figure 8g. TOFs were calculated on the basis of Cu sites determined by the N2O, with conversions below 10.0% achieved by adjusting the GHSV. The results also demonstrate a consistent trend, with the TOF values decreasing in the order CZCr > CZCe > CZZr > CZAl, among which the CZCr catalyst exhibits the highest TOF value.

3. Discussion

The performances of CZM catalysts promoted by different third elements (Zr, Ce, Cr) in CO2 hydrogenation to methanol is better than those of the traditional CZAl catalyst. At the same time, the structure and electron promotion of the third element were observed during the evolution of the corresponding physicochemical properties of the catalyst. The correlation between the catalytic performance and the physicochemical properties of the ternary CZM catalysts was systematically investigated, as shown in Figure 9. The results showed that the CH3OH yield was inversely proportional to the particle size of the catalyst. With the increase in Cu species’ particle size, the CH3OH yield decreased from 194.89 mg·gcat.−1·h−1 to 130.68 mg·gcat.−1·h−1, as shown in Figure 9a. Thus, the addition of Cr could form smaller and dispersed particles, which increased the specific surface area of Cu species and improved the catalytic activity [52]. Compared with other catalysts, the particle size effect of the CZCr catalyst was obvious. This was also attributed to the more efficient electronic interaction between Cu and Cr when Cr was introduced as the third element, thereby enhancing the dispersion of Cu. As shown by the reduced catalyst, the electronic states of both Cu and Cr have undergone significant changes. The first was that the CZCr catalyst produces more Cu0 than the other catalysts. On the other hand, the catalyst stability was related to Cu species (Cu+/Cu0) in that the higher the Cu+/Cu0 ratio, the faster the deactivation rate of the CZM catalysts. This might be attributed to the REDOX reaction between Cu species and reactants during the reaction. As can be seen from Figure 9b, when the Cu+/Cu0 ratio (3.05) was high, the deactivation rate of the CZAl catalyst reached 36.9%. On the contrary, the deactivation rate of the CZCr catalyst (Cu+/Cu0 = 0.88) was only 3.5%. This finding revealed that the higher the initial reduction degree of the ternary CZM catalysts, the better the long-term stability. In addition, the apparent activation energy (Ea) of CZM (M = Al, Ce, Zr, Cr) catalysts were tested. Under the action of different third elements, the catalysts showed different intrinsic activities in the reaction. As shown in Figure 9c, compared with the traditional CZAl catalyst (Ea = 53.59 ± 6.45 kJ/mol), other catalysts showed a lower apparent activation energy, especially the CZCr catalyst (Ea = 33.00 ± 3.97 kJ/mol), which could more effectively reduce the activation energy barrier of the reaction and promote the reaction. The introduction of Cr not only optimizes the dispersity of Cu species, but also produces certain electronic interactions with Cu, resulting in a smaller particle size and Cu+/Cu0 ratio of the CZCr catalyst, which further promotes the adsorption and activation of CO2 and H2 (Figure 7), thus reducing the apparent activation energy of the reaction.
For the CO2 hydrogenation to methanol, the traditional CZAl catalyst has been the most commonly used in the industry. The role of CuZnO as the reaction substrate is indisputable; however, the influence of the third element remains to be thoroughly examined, particularly elements such as Ce, Zr and Cr, which have demonstrated superior promotional effects compared to Al. This phenomenon can also be attributed to the intrinsic role of the third element as a promoter, which significantly influences the physicochemical properties of Cu-based catalysts, encompassing particle size, dispersion, surface electronic state and adsorption activation capacity, among others. More critically, it effectively lowers the activation energy of the reaction.

4. Materials and Methods

4.1. Catalyst Preparation

In this work, a range of ternary Cu-based catalysts, for which the mass ratio was 6:3:1, were prepared by conventional co-precipitation method. Firstly, copper nitrate (Cu(NO3)2·3H2O), zinc nitrate (Zn(NO3)2·6H2O) and auxiliary metal nitrate (among them, Cu(NO3)2·3H2O (99%, AR) and Zr(NO3)3·5H2O) were purchased from Sinopharm Chemical Reagent Co., Ltd., Shanghai, China; Zn(NO3)2·6H2O (99%, AR) was purchased from Xilong Scientific Co., Ltd., Shantou, China; Al(NO3)3·9H2O (99%, AR) was purchased from Tianjin Kemiou Chemical Reagent Co., Ltd., Tianjin, China; and Ce(NO3)3·6H2O and Cr(NO3)3·6H2O (99%, AR) were purchased from Shanghai Aladdin Biochemical Technology Co., Ltd., Shanghai, China). They were dissolved in deionized water to prepare a nitrate mixed solution with a metal ion concentration of 1.0 mol/L. In addition, a certain amount of Na2CO3 (99.8%, AR) was purchased from Tianjin Kemiou Chemical Reagent Co., Ltd., Tianjin, China, and was dissolved in distilled water, and the concentration of 1.0 mol/L solution was also prepared. Under stirring at 70 °C, the metal salt solution and precipitant solution were simultaneously added into 200 mL of distilled water. During this period, the drop rate of sodium carbonate was adjusted to control the pH of the suspension to 7.0 ± 0.2. After the drip addition, the mixture was stirred and aged for 1 h, then filtered, washed, dried overnight at 80 °C to obtain the catalyst precursor, and then calcined at 350 °C for 4 h to achieve the catalyst parent, which was recorded as CZM (M = Al, Ce, Zr, Cr).

4.2. Catalyst Characterization

X-ray fluorescence spectroscopy (XRF) characterization was performed on a Zetium instrument produced by Malvern Panalytical B.V., Almelo, The Netherlands, to determine the elemental composition of the catalysts. The method employed in this study was the pelletizing method. During the pelletizing process, boric acid (H3BO3) was used as a binder to improve the formability, uniformity and stability of the samples, while also reducing matrix effects and measurement errors.
The 1H NMR spectrum of methanol was measured using an AVANCE NEO 400 MHz instrument manufactured by Bruker, Berlin, Germany, with deuterated chloroform (CDCl3) as the solvent. The required magnetic field strength was 400 MHz.
XRD measurements were carried out using a TD-M001 diffractometer produced by Bruker, Berlin, Germany, for copper radiation. XRD patterns were collected by performing a stepwise scan at a rate of 8° per minute, covering a 2θ range extending from 10° to 80°. The mean crystallite sizes of Cu nanoparticles were determined using the Scherrer equation as follows.
d Cu = k λ β c o s θ
The specified parameters include the full width at half maximum (β), the diffraction angle (θ), the Scherrer constant (k = 0.89) and the X-ray wavelength (λ = 0.154 nm), in that order.
The dispersion and morphology of Cu were analyzed using a Regulus-8100 scanning electron microscope from Hitachi, Tokyo, Japan, which was furnished with an energy-dispersive X-ray spectrometer. A 40 mg sample was attached to a sample tray and characterized under accelerating voltage conditions.
The micro-morphology and structure of the catalyst were characterized by transmission electron microscopy (TEM). The experimental equipment was a field-emission transmission electron microscope (TEM) of the JEM-2100F model, produced by JEOL Ltd., Tokyo, Japan, with an acceleration voltage of 200 kV. A specific quantity of catalyst powder was incorporated into an ethanolic solution. After ultrasonic treatment for 40 min, the ensuing suspension was spread onto a carbon-supported copper mesh and left to dry in the air.
H2-TPR was performed using a FINESORB-3010 instrument produced by Beijing Jingwei Gaobo Science and Technology Co., Ltd., Beijing, China. A 40 mg portion of the sample was introduced into a U-shaped quartz reactor, which was subsequently flushed with argon (Ar) at a temperature of 300 °C for 1 h. Once the system reached ambient temperature, the gas stream was replaced with a 10 vol% H2/Ar flow (30 mL/min). Subsequently, the temperature was incrementally increased to 600 °C at a gradient of 10 °C/min.
CO2-TPD analysis was performed on an instrument of the Micromeritics AutoChem 2920 model, produced by Micromeritics Instrument Corporation, Norcross, GA, USA. Initially, a 60 mg catalyst was loaded into a U-shaped quartz tube and pretreated at 50 °C under a continuous argon (Ar) flow. After removing surface impurities, the catalyst was reduced by injecting 10% H2 at 250 °C. The gas supply was subsequently switched to pure CO2 and maintained under these conditions for 1 h. Following this period, the system was purged with argon (Ar) for 30 min to eliminate any residual gases. Ultimately, the catalyst was subjected to a temperature ramp from 50 °C to 800 °C, conducted at a rate of 10 °C/min. The H2-TPD analysis was performed following the same methodology as the CO2-TPD procedure, with the exception that a gas mixture containing 10 vol% H2 in Ar was utilized.
XPS and XAES analyses were conducted on a state-of-the-art ESCALab 220i-XL VG instrument, produced by Thermo Fisher Scientific, Waltham, MA, USA. Excitations of the energy spectra were achieved with Al Kα radiation, which had a characteristic wavelength of 1486.6 eV. For the sake of enhancing the accuracy of the measurements, the binding energy (BE) readings were carefully aligned with reference to the C1s peak, set at 284.6 eV, with a precision of ±0.1 eV. The positions of photoelectron peaks are reported in terms of binding energy (BE), while the Auger peaks are referenced to the kinetic energy (KE). In order to prevent severe oxidation of copper-based catalysts, our XPS tests were performed in a low-temperature and dry environment.
Room-temperature Raman spectroscopy was conducted using a high-resolution LabRAM HR system from HYJ, Cannes, France, equipped with a CCD detector and a spectral resolution of 1 cm−1. The system utilized a 532 nm laser as the excitation source, coupled with 1800 grooves per millimeter grating in a backscattering configuration. The laser power was set to 6 mW, with an exposure time of 60 s. The wavenumber range spanned from 200 to 800 cm−1.
The dispersity of Cu species was determined using N2O chemisorption on a Micromeritics AutoChem 2920 system. A 40 mg sample was heated from 50 °C to 500 °C (10 °C/min) in a flow of 10% H2/Ar fluid (30 mL/min), and we recorded the H2 consumption as X. After reduction, the sample was cooled down to 50 °C and re-oxidized at 50 °C with N2O (30 mL/min) for 30 min. Subsequently, the sample was flushed with a pure Ar fluid for 30 min to remove the physically adsorbed N2O. The catalyst was then reduced again according to the above procedure and the H2 consumption was recorded as Y. The dispersion of Cu species and the size of exposed Cu particles were calculated using the following formulas.
CuO + H2 → Cu + H2O; H2 consumption = X
N2O + 2Cu → N2 + Cu2O
Cu2O + H2 → 2Cu + H2O; H2 consumption = Y
D Cu ( % ) = 2 Y X × 100 %
S Cu ( % ) = 2 × Y × N A X × M Cu × 1 . 4 × 10 19 = 1353 Y X   ( m 2 / g )
d Cu = 104 . 25 D Cu × 100 %
where D Cu represents the dispersion of Cu species, S Cu denotes the surface area of Cu species, M Cu is the molar mass of Cu, N A stands for Avogadro’s constant (6.022 × 1023 mol−1) and 1.4 × 1019 corresponds to the number of Cu atoms per unit area (/m2).

4.3. Catalyst Evaluation

The catalytic performance was evaluated using a continuous fixed-bed reactor. A 0.5 g sample of catalyst, mesh size 20–40, was initially charged into a stainless steel reaction tube with dimensions of a 9 mm diameter and a 1450 mm length. The catalyst bed was then uniformly surrounded by quartz sand of equivalent particle size, ensuring a consistent environment above and below the bed. Before the reaction, the catalyst was reduced to 4 h under atmospheric pressure in 10% H2/N2 in 60 mL/min at 250 °C. Following the initial setup, a gas mixture with a composition of 72 vol% H2, 24 vol% CO2 and 4 vol% Ar was introduced into the reaction tube. The pressure was rapidly accumulated to 3 MPa, followed by heating the reactor between 200 and 260 °C to study the temperature’s effect. CO and CO2 concentrations were determined using GC (Gas Chromatography, Manufacturer: PANNA, Model: A91, Manufacturer: Changzhou Panna Instrument Co., Ltd., Changzhou, China) with a TCD and two columns: a 5A molecular sieve and a GDX-104 packed column. Methanol was detected using an FID (Flame Ionization Detector) equipped with a Porapak-Q column and further characterized by 1H NMR. The test was conducted using deuterated chloroform (CDCl3) as the solvent, and the results are shown in Figure S3. The spectrum clearly displays the characteristic signals of methanol, including two main peaks: Peak A corresponds to the methyl hydrogen (-CH3) with a chemical shift range of 3.3–3.5 ppm and Peak B corresponds to the hydroxyl hydrogen (-OH) with a chemical shift range of 2.0–4.0 ppm, exhibiting a broad peak shape. The integration ratio of these signals is approximately 3:1, consistent with the expected structure of methanol. Ar was included in the feedstock as a reference for analytical purposes. The CO2 conversion rates and selectivity of the products for each catalyst were recorded. The calculation methods were as follows:
X C O 2 ( % ) = ( 1 A o u t ( C O 2 ) / A o u t ( A r ) A i n ( C O 2 ) / A i n ( A r ) ) × 100 %
where A i n ( C O 2 ) and A i n ( A r ) are the pre-reaction peak areas of CO2 and Ar; A o u t ( C O 2 ) and A o u t ( A r ) are the post-reaction peak areas of CO2 and Ar.
S ( C O ) ( % ) = A o u t ( C O ) / A o u t ( A r ) × f C O / A r ( A i n ( C O 2 ) / A i n ( A r ) A o u t ( C O 2 ) / A o u t ( A r ) ) × f C O 2 / A r × 100 %
S ( C H 3 O H ) ( % ) = ( A C H 3 O H × f C H 3 O H ( A i n ( C O 2 ) / A i n ( A r ) A o u t ( C O 2 ) / A o u t ( A r ) ) × f C O 2 / A r ) × 100 %
where A o u t ( C O ) and A C H 3 O H are the peak areas of CO and CH3OH, respectively. f C O / A r , f C O 2 / A r and f C H 3 O H , respectively, correspond to the correction factors of CO, CO2 and CH3OH.
The formula for calculating the turnover frequency ( T O F ) of methanol is as follows:
T O F   ( h 1 ) = F C O 2 , i n × X C O 2 × S C H 3 O H × M C u W C u × D C u
where F C O 2 , i n (mol h−1) is the molar flow rate of CO2 in pristine mixed gas; M C u is the molecular weight of copper, 63.546 g mol−1; and W C u and D C u (%) are the catalyst weight and the actual loading of copper, respectively.

5. Conclusions

In summary, the impact of various promoters (Al, Ce, Zr, Cr) on the catalytic performance was systematically investigated. The results indicate that the optimal CZCr catalyst exhibited a higher CO2 conversion of 21.1%, CH3OH selectivity of 56.4% and superior stability with only 3.57% deactivation over 100 h, under conditions of 3 MPa and 240 °C. The characterization results showed that the physical and chemical properties of the catalyst greatly changed under the action of different accelerators. The third element affected the dispersion and interaction degree of Cu species in different degrees, resulting in different particle sizes and Cu+/Cu0 ratios. The smaller particle size led to higher methanol yield, while the lower Cu+/Cu0 ratio resulted in a lower deactivation rate. As a whole, the introduction of high-quality third elements (such as Cr) can reduce the apparent activation energy of the ternary CZM catalysts for the hydrogenation of CO2 to methanol. The preliminary research also provides an important way to improve the energy conversion efficiency reasonably and develop the catalyst for industrial application.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal15030250/s1, Figure S1: Raman results of ternary CZM catalysts. Figure S2: The corresponding EDS elemental mappings of Cu, Zn, O, Al, Ce, Zr, Cr and C from (a) CZAl catalyst; (b) CZCe catalyst; (c) CZZr catalyst; (d) CZCr catalyst, respectively. Figure S3: The ¹H NMR spectrum of methanol. Refs. [53,54,55,56,57,58,59,60] are cited in the Supplementary Materials.

Author Contributions

Writing—original draft, validation, data curation and formal analysis, J.W.; writing—review and editing, supervision and software, P.S.; methodology, conceptualization and software, J.H.; resources and investigation, Y.T.; validation and data curation, Z.Y.; formal analysis and validation, P.L.; investigation, validation and data curation, H.B.; resources and investigation, Y.L.; investigation, conceptualization and supervision, W.C.; conceptualization and data curation, G.L.; data curation, Z.Z.; writing—review and editing, conceptualization, data curation, supervision and funding acquisition, C.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Key R & D Program of Shanxi Province (202202090301013), the National Natural Science Foundation of China (No. 21676176 and 22309130) and the Natural Science Foundation of Shanxi Province (20210302123143).

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Material. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

Authors Guanjing Lin and Zhiqiang Zhai were employed by the company Shanxi Dongyi Coal Electric Aluminum Group Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Zhou, Y.; Liu, F.; Geng, S.; Yao, M.; Ma, J.; Cao, J. Tuning the content of S vacancies in MoS2 by Cu doping for enhancing catalytic hydrogenation of CO2 to methanol. Mol. Catal. 2023, 547, 113288. [Google Scholar] [CrossRef]
  2. Liu, L.; Wang, C.; Xue, F.; Li, J.; Zhang, H.; Lu, S.; Su, X.; Cao, B.; Huo, W.; Fang, T. DFT investigation of CO2 hydrogenation to methanol over Ir-doped Cu surface. Mol. Catal. 2022, 528, 112460. [Google Scholar] [CrossRef]
  3. Kapiamba, K.F.; Otor, H.O.; Viamajala, S.; Alba-Rubio, A.C. Inverse oxide/metal catalysts for CO2 hydrogenation to methanol. Energy Fuels 2022, 36, 11691–11711. [Google Scholar] [CrossRef]
  4. Roy, S.; Mondal, D.K.; Chatterjee, S.; Chowdhury, A.; Khan, T.S.; Haider, M.A.; Mandal, S.; Chandra, D.; Hara, M.; Bhaumik, A. Selective CO2 reduction to methane catalyzed by mesoporous Ru-Fe3O4/CeOx-SiO2 in a fixed bed flow reactor. Mol. Catal. 2022, 528, 112486. [Google Scholar] [CrossRef]
  5. Lykaki, M.; Stefa, S.; Varvoutis, G.; Binas, V.D.; Marnellos, G.E.; Konsolakis, M. Comparative assessment of first-row 3d transition metals (Ti-Zn) supported on CeO2 nanorods for CO2 hydrogenation. Catalysts 2024, 14, 611. [Google Scholar] [CrossRef]
  6. Murthy, P.S.; Liang, W.; Jiang, Y.; Huang, J. Cu-based nanocatalysts for CO2 hydrogenation to methanol. Energy Fuels 2021, 35, 8558–8584. [Google Scholar] [CrossRef]
  7. Kumar, A.; Daw, P.; Milstein, D. Homogeneous catalysis for sustainable energy: Hydrogen and methanol economies, fuels from biomass, and related topics. Chem. Rev. 2021, 122, 385–441. [Google Scholar] [CrossRef]
  8. Zhong, J.; Yang, X.; Wu, Z.; Liang, B.; Huang, Y.; Zhang, T. State of the art and perspectives in heterogeneous catalysis of CO2 hydrogenation to methanol. Chem. Soc. Rev. 2020, 49, 1385–1413. [Google Scholar] [CrossRef]
  9. Li, J.; Wang, L.; Hui, X.; Zhang, C.; Cao, Y.; Xu, S.; He, P.; Li, H. Effective hydrogenation of carbonates to produce methanol over a ternary Cu/Zn/Al catalyst. RSC Adv. 2020, 10, 13083–13094. [Google Scholar] [CrossRef]
  10. Carvalho, D.F.; Almeida, G.C.; Monteiro, R.S.; Mota, C.J.A. Hydrogenation of CO2 to methanol and dimethyl ether over a bifunctional Cu·ZnO catalyst impregnated on modified γ-alumina. Energy Fuels 2020, 34, 7269–7274. [Google Scholar]
  11. Cheng, Z.; Wang, M.; Jiang, C.; Zhang, L.; He, X.; Sun, X.; Lan, G.; Qiu, Y.; Li, Y. Tuning lattice strain of copper particles in Cu/ZnO/Al2O3 catalysts for methanol steam reforming. Energy Fuels 2024, 38, 15611–15621. [Google Scholar] [CrossRef]
  12. Al Abdulghani, A.J.; Turizo-Pinilla, E.E.; Fabregas-Angulo, M.J.; Hagmann, R.H.; Ibrahim, F.; Jansen, J.H.; Agbi, T.O.; Bhat, S.; Sepúlveda-Pagán, M.; Kraimer, M.O.; et al. Realizing synergy between Cu, Ga, and Zr for selective CO2 hydrogenation to methanol. Appl. Catal. B 2024, 355, 124198. [Google Scholar] [CrossRef]
  13. Liang, B.; Ma, J.; Su, X.; Yang, C.; Duan, H.; Zhou, H.; Deng, S.; Li, L.; Huang, Y. Investigation on deactivation of Cu/ZnO/Al2O3 catalyst for CO2 hydrogenation to methanol. Ind. Eng. Chem. Res. 2019, 58, 9030–9037. [Google Scholar] [CrossRef]
  14. Santana, C.S.; Rasteiro, L.F.; Marcos, F.C.F.; Assaf, E.M.; Gomes, J.F.; Assaf, J.M. Influence of Al, Cr, Ga, or Zr as promoters on the performance of Cu/ZnO catalyst for CO2 hydrogenation to methanol. Mol. Catal. 2022, 528, 112512. [Google Scholar] [CrossRef]
  15. Chang, X.; Zi, X.; Li, J.; Liu, F.; Han, X.; Chen, J.; Hao, Z.; Zhang, H.; Zhang, Z.; Gao, P.; et al. An insight into synergistic metal-oxide interaction in CO2 hydrogenation to methanol over Cu/ZnO/ZrO2. Catalysts 2023, 13, 1337. [Google Scholar] [CrossRef]
  16. Singh, R.; Kundu, K.; Pant, K.K. CO2 hydrogenation to methanol over Cu-ZnO-CeO2 catalyst: Reaction structure-activity relationship, optimizing Ce and Zn ratio, and kinetic study. Chem. Eng. J. 2024, 479, 147783. [Google Scholar] [CrossRef]
  17. Zhan, H.; Shi, X.; Tang, B.; Wang, G.; Ma, B.; Liu, W. The performance of Cu/Zn/Zr catalysts of different Zr/(Cu+Zn) ratio for CO2 hydrogenation to methanol. Catal. Commun. 2021, 149, 106264. [Google Scholar] [CrossRef]
  18. Xiong, S.; Lian, Y.; Xie, H.; Liu, B. Hydrogenation of CO2 to methanol over Cu/ZnCr catalyst. Fuel 2019, 256, 115975. [Google Scholar] [CrossRef]
  19. Hengne, A.M.; Yuan, D.J.; Date, N.S.; Saih, Y.; Kamble, S.P.; Rode, C.V.; Huang, K.-W. Preparation and activity of copper-gallium nanocomposite catalysts for carbon dioxide hydrogenation to methanol. Ind. Eng. Chem. Res. 2019, 58, 21331–21340. [Google Scholar] [CrossRef]
  20. 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]
  21. Huang, C.J.; Zhang, S.N.; Wang, W.B.; Zhou, H.Z.; Shao, Z.L.; Xia, L.; Wang, H.; Sun, Y.H. Modulation of electronic metal-support interaction between Cu and ZnO by Er for effective low temperature CO2 hydrogenation to methanol. ACS Catal. 2024, 14, 1324–1335. [Google Scholar] [CrossRef]
  22. Phongamwong, T.; Chantaprasertporn, U.; Witoon, T.; Numpilai, T.; Poo-arporn, Y.; Limphirat, W.; Donphai, W.; Dittanet, P.; Chareonpanich, M.; Limtrakul, J. CO2 hydrogenation to methanol over CuO-ZnO-ZrO2-SiO2 catalysts: Effects of SiO2 contents. Chem. Eng. J. 2017, 316, 692–703. [Google Scholar] [CrossRef]
  23. Lei, T.; Dang, Q.; Wu, T.; Wu, Y.; Yu, J. The role of ZnO nanowires to improve methanol selectivity on Cu2O nanocubes during CO2 electroreduction. Appl. Catal. A 2023, 666, 119394. [Google Scholar] [CrossRef]
  24. Kordus, D.; Jelic, J.; Luna, M.L.; Divins, N.J.; Timoshenko, J.; Chee, S.W.; Rettenmaier, C.; Kröhnert, J.; Kühl, S.; Trunschke, A.; et al. Shape-dependent CO2 hydrogenation to methanol over Cu2O nanocubes supported on ZnO. J. Am. Chem. Soc. 2023, 145, 3016–3030. [Google Scholar] [CrossRef]
  25. LaGrow, A.P.; Besenhard, M.O.; Hodzic, A.; Sergides, A.; Bogart, L.K.; Gavriilidis, A.; Thanh, N.T.K. Unravelling the growth mechanism of the co-precipitation of iron oxide nanoparticles with the aid of synchrotron X-Ray diffraction in solution. Nanoscale 2019, 11, 6620–6628. [Google Scholar] [CrossRef]
  26. Ban, H.; Li, C.; Asami, K.; Fujimoto, K. Influence of rare-earth elements (La, Nd and Pr) on the performance of Cu/Zn/Zr catalyst for CH3OH synthesis from CO2. Catal. Commun. 2014, 54, 50–54. [Google Scholar] [CrossRef]
  27. Liu, Q.; Zhong, Z.; Gu, F.; Wang, X.; Lu, X.; Li, H.; Xu, G.; Su, F. CO methanation on ordered mesoporous Ni-Cr-Al catalysts: Effects of the catalyst structure and Cr promoter on the catalytic properties. J. Catal. 2016, 337, 221–232. [Google Scholar] [CrossRef]
  28. Si, C.; Ban, H.; Chen, K.; Wang, X.; Cao, R.; Yi, Q.; Qin, Z.; Shi, L.; Li, Z.; Cai, W.; et al. Insight into the positive effect of Cu0/Cu+ ratio on the stability of Cu-ZnO-CeO2 catalyst for syngas hydrogenation. Appl. Catal. A 2020, 594, 117466. [Google Scholar] [CrossRef]
  29. Li, Z.; Men, Y.; Liu, S.; Wang, J.; Qin, K.; Tian, D.; Shi, T.; Zhang, L.; An, W. Boosting CO2 hydrogenation efficiency for methanol synthesis over Pd/In2O3/ZrO2 catalysts by crystalline phase effect. Appl. Surf. Sci. 2022, 603, 154420. [Google Scholar] [CrossRef]
  30. Kim, J.Y.; Rodriguez, J.A.; Hanson, J.C.; Frenkel, A.I.; Lee, P.L. Reduction of CuO and Cu2O with H2: H embedding and kinetic effects in the formation of suboxides. J. Am. Chem. Soc. 2003, 125, 10684–10692. [Google Scholar] [CrossRef]
  31. Qi, T.Q.; Zhao, Y.M.; Chen, S.Y.; Li, W.Z.; Guo, X.W.; Zhang, Y.C.; Song, C.S. Bimetallic metal organic framework-templated synthesis of a Cu-ZnO/Al2O3 catalyst with superior methanol selectivity for CO2 hydrogenation. Mol. Catal. 2021, 514, 111870. [Google Scholar] [CrossRef]
  32. Vergara, T.; Gómez, D.; de Oliveira Campos, B.L.; Delgado, K.H.; Concepción, P.; Jiménez, R.; Karelovic, A. Combined role of Ce promotion and TiO2 support improves CO2 hydrogenation to methanol on Cu catalysts: Interplay between structure and kinetics. J. Catal. 2023, 426, 200–213. [Google Scholar] [CrossRef]
  33. Zhu, J.D.; Ciolca, D.; Liu, L.; Parastaev, A.; Kosinov, N.; Hensen, E.J.M. Flame synthesis of Cu/ZnO-CeO2 catalysts: Synergistic metal-support interactions promote CH3OH selectivity in CO2 Hydrogenation. ACS Catal. 2021, 11, 4880–4892. [Google Scholar] [CrossRef] [PubMed]
  34. Guo, T.; Guo, Q.; Li, S.Z.; Hu, Y.K.; Yun, S.; Qian, Y.H. Effect of surface basicity over the supported Cu-ZnO catalysts on hydrogenation of CO2 to methanol. J. Catal. 2022, 407, 312–321. [Google Scholar] [CrossRef]
  35. Li, S.Z.; Guo, L.M.; Ishihara, T. Hydrogenation of CO2 to methanol over Cu/AlCeO catalyst. Catal. Today 2020, 339, 352–361. [Google Scholar] [CrossRef]
  36. Shi, P.; Han, J.; Yan, Z.; Luo, P.; Wang, J.; Ban, H.; Zhang, X.; Li, C. Insight into the correlation between Cu species and methanol selectivity from CO2 hydrogenation over Cu-based catalyst. Mol. Catal. 2024, 563, 114278. [Google Scholar] [CrossRef]
  37. Zhao, Y.; Li, S.; Wang, Y.; Shan, B.; Zhang, J.; Wang, S.; Ma, X. Efficient tuning of surface copper species of Cu/SiO2 catalyst for hydrogenation of dimethyl oxalate to ethylene glycol. Chem. Eng. J. 2017, 313, 759–768. [Google Scholar] [CrossRef]
  38. Yilmaz, M.; Handoko, A.D.; Parkin, I.P.; Sankar, G. Probing the electronic and geometric structures of photoactive electrodeposited Cu2O films by X-ray absorption spectroscopy. J. Catal. 2020, 389, 483–491. [Google Scholar] [CrossRef]
  39. Zhang, K.; Liu, X.; Zhang, L.; Qing, S.J.; Zhang, C.S.; Liu, Y.J.; Gao, Z.X. Cu-Zn-Al spinel catalyst for hydrogen production from methanol steam reforming. J. Fuel Chem. Technol. 2022, 50, 494–502. [Google Scholar]
  40. Liu, J.; Ren, H.; Fan, J.C.; Hang, W. Effects of heat treatment methods on the structure of Cu-Zn-Al catalyst prepared by liquid phase method and the catalytic performance of CO hydrogenation. J. Mol. Catal. 2019, 33, 166–173. [Google Scholar]
  41. Gong, N.; Zhang, T.; Tan, M.; Wang, L.; Yang, J.; Tan, L.; Yang, G.; Wu, P.; Wu, Y.; Tan, Y. Realizing and revealing complex isobutyl alcohol production over a simple Cu-ZrO2 catalyst. ACS Catal. 2023, 13, 3563–3574. [Google Scholar] [CrossRef]
  42. Marcos, F.C.F.; Alvim, R.S.; Lin, L.; Betancourt, L.E.; Petrolini, D.D.; Senanayake, S.D.; Alves, R.M.B.; Assaf, J.M.; Rodriguez, J.A.; Giudici, R.; et al. The role of copper crystallization and segregation toward enhanced methanol synthesis via CO2 hydrogenation over CuZrO2 catalysts: A combined experimental and computational study. Chem. Eng. J. 2023, 452, 139519. [Google Scholar] [CrossRef]
  43. Zhang, F.; Gutiérrez, R.A.; Lustemberg, P.G.; Liu, Z.; Rui, N.; Wu, T.; Ramírez, P.J.; Xu, W.; Idriss, H.; Ganduglia-Pirovano, M.V.; et al. Metal-support interactions and C1 chemistry: Transforming Pt-CeO2 into a highly active and stable catalyst for the conversion of carbon dioxide and methane. ACS Catal. 2021, 11, 1613–1623. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, J.; Shen, S.; Wu, P.; Guo, L. Nitrogen-doped CeOx nanoparticles modified graphitic carbon nitride for enhanced photocatalytic hydrogen production. Green Chem. 2015, 17, 509–517. [Google Scholar] [CrossRef]
  45. Xiao, Y.; Tan, S.; Wang, D.; Wu, J.; Jia, T.; Liu, Q.; Qi, Y.; Qi, X.; He, P.; Zhou, M. CeO2/BiOIO3 heterojunction with oxygen vacancies and Ce4+/Ce3+ redox centers synergistically enhanced photocatalytic removal heavy metal. Appl. Surf. Sci. 2020, 530, 147116. [Google Scholar] [CrossRef]
  46. Lin, S.D.; Hsiao, T.C.; Chen, L.-C. The steady state methanol decomposition reaction over Cu/Zn and Cu/Cr catalysts: Pretreatment, operando EXAFS, and activity study. Appl. Catal. A 2009, 360, 226–231. [Google Scholar] [CrossRef]
  47. Gholami, S.; Alavi, S.M.; Rezaei, M. CO2 methanation over nanocrystalline Ni catalysts supported on mechanochemically synthesized Cr2O3-M (M = Fe, Co, La, and Mn) carriers. Int. J. Hydrogen Energy 2021, 46, 35571–35584. [Google Scholar] [CrossRef]
  48. Zuo, J.; Na, W.; Zhang, P.; Yang, X.; Wen, J.; Zheng, M.; Wang, H. Enhanced activity of CexZr1-xO2 solid solutions supported Cu-based catalysts for hydrogenation of CO2 to methanol. Mol. Catal. 2022, 526, 112357. [Google Scholar] [CrossRef]
  49. Proaño, L.; Jones, C.W. CO2 hydrogenation to methanol over ceria-zirconia NiGa alloy catalysts. Appl. Catal. A 2024, 669, 119485. [Google Scholar] [CrossRef]
  50. Bronsato, B.J.; Souza, E.F.; Gonzalez, G.G.; Chagas, L.H.; Zonetti, P.C.; Mendoza, C.D.; Huaman, N.R.; de Avillez, R.R.; Appel, L.G. What role does Al3+ play in the methanol synthesis from CO2 hydrogenation using Cu/ZnO/Al catalysts. Fuel 2024, 375, 132533. [Google Scholar] [CrossRef]
  51. Wang, X.; Yao, Z.; Guo, X.; Yan, Z.; Ban, H.; Wang, P.; Yao, R.; Li, L.; Li, C. Modulating electronic interaction over Zr-ZnO catalysts to enhance CO2 hydrogenation to methanol. ACS Catal. 2023, 14, 508–521. [Google Scholar] [CrossRef]
  52. Ay, S.; Ozdemir, M.; Melikoglu, M. Effects of magnesium and chromium addition on stability, activity and structure of copper-based methanol synthesis catalysts. Int. J. Hydrogen Energy 2021, 46, 12857–12873. [Google Scholar] [CrossRef]
  53. Chen, H.; Cui, H.S.; Lv, Y.; Liu, P.L.; Hao, F.; Xiong, W.; Luo, H. CO2 hydrogenation to methanol over Cu/ZnO/ZrO2 catalysts: Effects of ZnO morphology and oxygen vacancy. Fuel 2022, 314, 123035. [Google Scholar] [CrossRef]
  54. Sun, Y.; Huang, C.L.; Chen, L.M.; Zhang, Y.J.; Fu, M.L.; Wu, J.L.; Ye, D.Q. Active site structure study of Cu/Plate ZnO model catalysts for CO2 hydrogenation to methanol under the real reaction conditions. J. CO2 Util. 2020, 37, 55–64. [Google Scholar] [CrossRef]
  55. Wang, J.; Xia, Y.; Dong, Y.; Chen, R.S.; Xiang, L.; Komarneni, S. Defect-rich ZnO nanosheets of high surface area as an efficient visible-light photocatalyst. Appl. Catal. B Environ. 2016, 192, 8–16. [Google Scholar] [CrossRef]
  56. Hu, Z.; Liu, X.F.; Meng, D.M.; Guo, Y.; Guo, Y.L.; Lu, G.Z. Effect of ceria crystal plane on the physicochemical and catalytic properties of Pd/ceria for CO and propane oxidation. ACS Catal. 2016, 6, 2265–2279. [Google Scholar] [CrossRef]
  57. Luo, P.C.; Shi, P.X.; Yan, Z.Q.; Han, J.H.; Wang, J.J.; Li, Y.C.; Ban, H.Y.; Cai, W.J.; Li, C.M. Ternary synergistic interaction of Cu-ZnO-ZrO2 promoting CO2 hydrogenation to methanol. Appl. Catal. A Gen. 2025, 689, 120006. [Google Scholar] [CrossRef]
  58. Jiang, F.; Wang, S.S.; Liu, B.; Liu, J.; Wang, L.; Xiao, Y.; Xu, Y.B.; Liu, X.H. Insights into the influence of CeO2 crystal facet on CO2 hydrogenation to methanol over Pd/CeO2 catalysts. ACS Catal. 2020, 10, 11493–11509. [Google Scholar] [CrossRef]
  59. Singh, R.; Pandey, V.; Pant, K.K. Promotional role of oxygen vacancy defects and Cu-Ce interfacial sites on the activity of Cu/CeO2 catalyst for CO2 hydrogenation to methanol. ChemCatChem 2022, 14, e202201053. [Google Scholar] [CrossRef]
  60. Qin, H.; Ye, Y.K.; Lin, G.L.; Zhang, J.Y.; Jia, W.Q.; Xia, W.; Jiao, L.F. Regulating the electrochemical microenvironment of Ni(OH)2 by Cr doping for highly efficient methanol electrooxidation. ACS Catal. 2024, 14, 16234–16244. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of CZM catalysts (a) after calcination and (b) after reduction.
Figure 1. XRD patterns of CZM catalysts (a) after calcination and (b) after reduction.
Catalysts 15 00250 g001
Figure 2. SEM images of calcined catalysts: (A) CZAl; (B) CZCe; (C) CZZr; (D) CZCr.
Figure 2. SEM images of calcined catalysts: (A) CZAl; (B) CZCe; (C) CZZr; (D) CZCr.
Catalysts 15 00250 g002
Figure 3. TEM images and particle size distribution statistics of ternary CZM catalysts: (a) CZAl; (b) CZCe; (c) CZZr; (d) CZCr. (The particle size count is 200 for each image.)
Figure 3. TEM images and particle size distribution statistics of ternary CZM catalysts: (a) CZAl; (b) CZCe; (c) CZZr; (d) CZCr. (The particle size count is 200 for each image.)
Catalysts 15 00250 g003
Figure 4. TPR profiles of the ternary CZM (M = Al, Ce, Zr, Cr) catalysts.
Figure 4. TPR profiles of the ternary CZM (M = Al, Ce, Zr, Cr) catalysts.
Catalysts 15 00250 g004
Figure 5. XPS results of ternary CZM catalysts: (a) Cu 2p spectra of catalysts after calcination; (b) Cu 2p spectra of catalysts after reduction; (c) Cu LMM spectra of catalysts after calcination; (d) Cu LMM spectra of catalysts after reduction.
Figure 5. XPS results of ternary CZM catalysts: (a) Cu 2p spectra of catalysts after calcination; (b) Cu 2p spectra of catalysts after reduction; (c) Cu LMM spectra of catalysts after calcination; (d) Cu LMM spectra of catalysts after reduction.
Catalysts 15 00250 g005
Figure 6. XPS characterization results of the third element before and after reduction: (a) Al 2p; (b) Ce 3d; (c) Zr 3d; (d) Cr 2p.
Figure 6. XPS characterization results of the third element before and after reduction: (a) Al 2p; (b) Ce 3d; (c) Zr 3d; (d) Cr 2p.
Catalysts 15 00250 g006
Figure 7. CO2-TPD (left) and H2-TPD (right) patterns of the ternary CZM catalysts.
Figure 7. CO2-TPD (left) and H2-TPD (right) patterns of the ternary CZM catalysts.
Catalysts 15 00250 g007
Figure 8. Performance evaluation results of ternary CZM catalysts: (a) CO2 conversion and (b) CH3OH selectivity under 3 MPa and 200–260 °C; (c,d) stability test for 100 h (reaction conditions: T = 240 °C, P = 3 MPa, GHSV= 3600 mL/gcat./h); (e) CH3OH yields for 1 h and 100 h; (f) inactivation rate. (g) TOF values of the CZM catalyst at conversions below 10% by adjusting GHSV.
Figure 8. Performance evaluation results of ternary CZM catalysts: (a) CO2 conversion and (b) CH3OH selectivity under 3 MPa and 200–260 °C; (c,d) stability test for 100 h (reaction conditions: T = 240 °C, P = 3 MPa, GHSV= 3600 mL/gcat./h); (e) CH3OH yields for 1 h and 100 h; (f) inactivation rate. (g) TOF values of the CZM catalyst at conversions below 10% by adjusting GHSV.
Catalysts 15 00250 g008aCatalysts 15 00250 g008b
Figure 9. Correlation between performance and physicochemical properties of ternary CZM catalyst: (a) CH3OH yield and Cu particle size (reaction condition: H2/CO2 = 3/1, P = 3 MPa, T = 240 °C, GHSV = 3600 mL/gcat./h); (b) deactivation and Cu+/Cu0 (reaction for 100 h); (c) Arrhenius plots of CZM catalysts (M = Al, Ce, Zr, Cr).
Figure 9. Correlation between performance and physicochemical properties of ternary CZM catalyst: (a) CH3OH yield and Cu particle size (reaction condition: H2/CO2 = 3/1, P = 3 MPa, T = 240 °C, GHSV = 3600 mL/gcat./h); (b) deactivation and Cu+/Cu0 (reaction for 100 h); (c) Arrhenius plots of CZM catalysts (M = Al, Ce, Zr, Cr).
Catalysts 15 00250 g009
Table 1. Mass percentage content (wt%) of each element in the CZM catalysts.
Table 1. Mass percentage content (wt%) of each element in the CZM catalysts.
CatalystElementMass Percentage Content (wt%)
CZAlCu60.761
Zn29.126
Al9.732
CZCeCu60.169
Zn29.267
Ce9.8
CZZrCu59.867
Zn28.475
Zr11.137
CZCrCu61.430
Zn28.194
Cr10.078
Table 2. Textural properties and particle size of CZM catalysts.
Table 2. Textural properties and particle size of CZM catalysts.
SampleDCu 1 (%)SCu 1 (m2/g)dCu 1/nmdCu 2/nmdCu 3/nm
CZAl6.3142.6916.5217 ± 0.219.3 ± 0.4
CZCe7.5250.8713.8614 ± 0.414.3 ± 0.5
CZZr6.0841.1517.1316 ± 0.217.4 ± 0.3
CZCr9.7866.1710.6511 ± 0.311.1 ± 0.3
1 Via N2O chemisorption. 2 Based on Scherrer equation by XRD. 3 The size of Cu particles counted by TEM.
Table 3. TPR peak position temperature and concentration of ternary CZM catalysts.
Table 3. TPR peak position temperature and concentration of ternary CZM catalysts.
SamplePeak LPeak H
T/°CConcentration/%T/°CConcentration/%
CZAl154.351.8180.848.2
CZCe150.957.5164.342.5
CZZr156.856.6175.143.4
CZCr160.057.9169.642.1
Table 4. Deconvolution results of the Cu LMM XAES of ternary CZM catalysts after reduction.
Table 4. Deconvolution results of the Cu LMM XAES of ternary CZM catalysts after reduction.
SampleCu+ (%)Cu0 (%)Cu+/Cu0
CZAl75.324.73.05
CZCe66.833.22.01
CZZr63.836.21.76
CZCr46.853.20.88
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

Wang, J.; Shi, P.; Han, J.; Tian, Y.; Yan, Z.; Luo, P.; Ban, H.; Li, Y.; Cai, W.; Lin, G.; et al. Investigation of Ternary CuZnM (M = Cr, Ce, Zr, Al) Catalysts in CO2 Hydrogenation for Methanol Synthesis. Catalysts 2025, 15, 250. https://doi.org/10.3390/catal15030250

AMA Style

Wang J, Shi P, Han J, Tian Y, Yan Z, Luo P, Ban H, Li Y, Cai W, Lin G, et al. Investigation of Ternary CuZnM (M = Cr, Ce, Zr, Al) Catalysts in CO2 Hydrogenation for Methanol Synthesis. Catalysts. 2025; 15(3):250. https://doi.org/10.3390/catal15030250

Chicago/Turabian Style

Wang, Jingjing, Peixiang Shi, Jiahao Han, Yuhao Tian, Zhiqiang Yan, Pengcheng Luo, Hongyan Ban, Yanchun Li, Weijie Cai, Guanjing Lin, and et al. 2025. "Investigation of Ternary CuZnM (M = Cr, Ce, Zr, Al) Catalysts in CO2 Hydrogenation for Methanol Synthesis" Catalysts 15, no. 3: 250. https://doi.org/10.3390/catal15030250

APA Style

Wang, J., Shi, P., Han, J., Tian, Y., Yan, Z., Luo, P., Ban, H., Li, Y., Cai, W., Lin, G., Zhai, Z., & Li, C. (2025). Investigation of Ternary CuZnM (M = Cr, Ce, Zr, Al) Catalysts in CO2 Hydrogenation for Methanol Synthesis. Catalysts, 15(3), 250. https://doi.org/10.3390/catal15030250

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