Next Article in Journal
Exploring the Methane to Methanol Oxidation over Iron and Copper Sites in Metal–Organic Frameworks
Next Article in Special Issue
Thermodynamic Insights into Sustainable Aviation Fuel Synthesis via CO/CO2 Hydrogenation
Previous Article in Journal
A Comprehensive Overview on Biochar-Based Materials for Catalytic Applications
Previous Article in Special Issue
Metal–Organic Frameworks for Electrocatalytic CO2 Reduction into Formic Acid
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Insight into Synergistic Metal-Oxide Interaction in CO2 Hydrogenation to Methanol over Cu/ZnO/ZrO2

1
Key Laboratory for Green Chemical Technology of Ministry of Education, Collaborative Innovation Center of Chemical Science and Engineering, National Industry-Education Integration Platform of Energy Storage, School of Chemical Engineering and Technology, Tianjin University, Tianjin 300350, China
2
State Key Laboratory of Solid Waste Reuse for Building Materials, Beijing Building Materials Academy of Science Research, Beijing 100041, China
3
Joint School of National University of Singapore and Tianjin University, International Campus of Tianjin University, Binhai New City, Fuzhou 350207, China
*
Author to whom correspondence should be addressed.
The authors contribute equally to this study.
Catalysts 2023, 13(10), 1337; https://doi.org/10.3390/catal13101337
Submission received: 13 September 2023 / Revised: 28 September 2023 / Accepted: 29 September 2023 / Published: 30 September 2023
(This article belongs to the Special Issue Catalysis for Selective Hydrogenation of CO and CO2)

Abstract

:
The metal-oxide interaction is of significance to the construction of active sites for Cu-catalyzed CO2 hydrogenation to methanol. This study examines the effect of ZnO and ZrO2 composition on the Cu/ZnO/ZrO2 catalyst structure and surface properties to further tune the catalytic activity for methanol synthesis. The ZnO/ZrO2 ratio can impact the CuZn alloy formation from strong Cu-ZnO interactions and the surface basic sites for CO2 adsorption at the Cu-ZrO2 interface. The proportional correlation of the CuZn alloy content and CO2 desorption amount with the space-time yield (STY) of methanol reveals a synergistic interaction between Cu and oxides (ZnO and ZrO2) that enhances methanol synthesis. The optimized Cu/ZnO/ZrO2 catalyst exhibits higher STY relative to the traditional Cu/ZnO/Al2O3 catalyst. The obtained results presented herein can provide insight into the catalyst design for methanol synthesis from CO2.

1. Introduction

The production process of methanol, a high-value-added chemical, has historically received widespread attention. As early as 1923, BASF Company in Germany took the lead in developing high-pressure methanol synthesis technology from syngas [1]. Afterward, ICI discovered a low-pressure process for the synthesis of CH3OH using Cu-based catalysts [2]. Recently, Bai et al. [3] proposed that CO2 from the atmosphere and H2 from renewable energy sources can be used for CO2 hydrogenation to methanol, which can realize chemical utilization of CO2 and storage of renewable energy.
CO2 hydrogenation to methanol is an exothermic reaction, with high-pressure and low-temperature conditions being thermodynamically beneficial for it. However, CO2 molecules are relatively stable, which makes CO2-to-methanol conversion difficult at low temperatures. At higher temperatures, CO2 is prone to reverse water-gas shift (RWGS), which is an endothermic reaction that generates by-product CO and reduces the methanol selectivity.
CO2 + 3H2 → CH3OH + H2O, ∆H298K = −49.5 kJ/mol
CO2 + H2 → CO + H2O, ∆H298K = 41.2 kJ/mol
CO2 hydrogenation to methanol catalysts can be roughly divided into Cu-based [4,5,6,7,8,9,10,11,12,13,14], precious metal (Au [15,16], Ag [17], Pt [18], etc.), bimetallic (Cu-Pd [19], Zn-Pd [20], Co-Pd [21], etc.), and oxide (In2O3) catalysts [22,23,24,25]. The Cu/ZnO/Al2O3 catalyst developed by ICI can also be applied to the CO2 hydrogenation to methanol. However, the catalytic performance bears room for improvement, and the presence of water in the product can promote the sintering of Cu and ZnO, leading to catalyst deactivation [26]. It is of importance to develop highly active and stable Cu-based catalysts for methanol synthesis from CO2.
The existing literature endeavors to design efficient active sites for Cu-based catalysts. However, there is no consensus on the active site structure due to the catalyst complexity. A synergistic effect between Cu and oxide (e.g., ZnO, ZrO2, and CeO2) at the interface has been proposed as the active site [7]. Another view is that CuZn alloy is an active site; its formation may promote the partial reduction of ZnO to Znq+ or Zn0, thereby modifying the Cu surface [9,11,27]. The dissociated H atom on Cu0 might overflow to the CuZn alloy, and then react with the adsorbed species CO2* by ZnO near the CuZn alloy, so as to improve the catalyst performance in the CO2 hydrogenation to methanol [28]. Compared to syngas, CO2 hydrogenation to methanol produces water, which requires a weakly hydrophilic material to timely desorb the generated water. In previous work [29], a highly dispersed Cu−ZrO2 interface can effectively improve the conversion rate and methanol selectivity. Meanwhile, compared to Al2O3, ZrO2 shows weaker hydrophilicity. In this work, we propose the design of a tertiary Cu-ZnO−ZrO2 catalyst with the construction of Cu-ZnO and Cu−ZrO2 interactions as active centers, which potentially promotes CO2 hydrogenation to methanol.
In this study, Cu/ZnO, Cu/ZrO2, and Cu/ZnO/ZrO2 with a series of mass ratios of ZnO/ZrO2 were constructed with the structure and surface properties studied using a range of characterizations [e.g., XRD (powder X-ray diffraction), TEM (transmission electron microscopy), XPS (X-ray photoelectron spectra), and CO2-TPD (CO2 temperature-programmed desorption)]. The optimized Cu/ZnO/ZrO2 catalyst outperforms the conventional Cu/ZnO/Al2O3 catalyst in terms of activity for CO2 hydrogenation to methanol. The presence of CuZn alloy and Cu−ZrO2 interaction enhances CO2 conversion and methanol formation rates.

2. Results and Discussion

2.1. Characterization

A series of Cu/ZnO/ZrO2 catalysts with different ZnO/ZrO2 mass ratios were synthesized via oxalate coprecipitation. The catalysts were named based on element content. For example, the ZnO/ZrO2 ratio is 2:5 in the catalyst, which is named C3Z2Z5. The element contents measured using ICP-OES (inductively coupled plasma optical emission spectroscopy) are close to the target values [Table 1, including specific surface area (SBET), specific surface area of Cu (SCu), Cu dispersion (DCu), Cu diameters (dCu), and other physicochemical properties] indicative of no significant loss of Cu during the catalyst preparation. In the adsorption and desorption isotherms (Figure S1A), all the catalysts show typical Type IV hysteresis loops, indicating that the prepared catalysts are mesoporous. The pore size distribution in Figure S1B also confirms the mesoporous structure. As the ZnO/ZrO2 ratio increases, the specific surface area shows a decreasing trend, which is not conducive to the dispersion of Cu (Table 1). This indicates that ZrO2 plays an important role in maintaining the specific surface area of the Cu/ZnO/ZrO2 catalyst. In the Zr-containing catalysts, the change in ZnO/ZrO2 ratio does not have a significant impact on the pore volume. Compared with Cu/ZrO2 and Cu/ZnO/ZrO2, the pore volume of Cu/ZnO is significantly reduced. The pore size shows an increasing trend with the increase of the ZnO/ZrO2 ratio.
The crystalline structures of all the catalysts after calcination and reduction were analyzed by XRD (Figure 1). The monoclinic CuO (2θ = 35.5° and 38.7°, PDF#48-1548) was observed in all the calcined samples in Figure 1A, while ZnO (2θ = 31.8°, 34.4° and 38.7°, PDF#79-0206) [30,31] can only be detected when the Zn content exceeds 20 wt%. The ZrO2 in all samples was a-ZrO2, because of the low calcination temperature, and no characteristic peak of ZrO2 was observed. The monoclinic CuO phase was reduced to a cubic metallic Cu0 phase (2θ = 43.3°, 50.4°, and 74.1°, PDF#4-836) and Cu2O (2θ = 37.1°, PDF#65-3288) via H2 reduction at 300 °C in Figure 1B [32,33]. As the ZnO/ZrO2 ratio increased, the Cu reflection signal intensified, indicating that metal Cu continued to aggregate. In Figure 1C, it is worth noting that the Cu characteristic diffraction peak (2θ = 43.3°) moved to lower degrees as the ZnO content increased. This indicates that during the catalyst reduction activation process, a fraction of ZnO is reduced to Zn0 and interacts with Cu to form a CuZn alloy [34,35].
The TEM image (Figure S2) shows that the reduced catalyst exhibits irregular morphology, with a size ranging approximately from 400–600 nm. In Cu/ZrO2 and Cu/ZnO/ZrO2, the black area weakly increased with the ZnO content increased, due to Cu aggregation. However, when the ZrO2 content decreased from 10% (C3Z6Z1) to 0% (C3Z7Z0), the size of Cu particles increased significantly, which is consistent with the XRD analysis. The size of Cu particles in Cu/ZnO is significantly larger than that in Cu/ZrO2, which suggests that ZrO2 has a stronger dispersion effect on Cu than ZnO [36]. Regular hemispherical or spherical Cu nanoparticles can be found in the HR-TEM images (Figure 2). In the C3Z1Z6 sample, Cu particles expose the (100) and (111) crystal planes corresponding to the lattice spacing = 0.182 and 0.210 nm, and ZnO particles expose the (101) crystal planes corresponding to the lattice spacing = 0.248 nm [37]. As the ZnO/ZrO2 mass ratio increased, Cu (111) crystal plane lattice spacing expanded to 0.216–0.219 nm. The expansion of lattice stripes indicates that during the reduction process, ZnO is reduced to Zn0 and combines with the metallic Cu to form a CuZn alloy, which is consistent with the XRD result [35,38]. No lattice fringe of ZrO2 was found in the HR-TEM image, which is related to the amorphous morphology of ZrO2. EDS-mapping (Energy Dispersive Spectrometer mapping) was used to determine the distribution of Cu, Zn, and Zr in the catalyst. In Figure S3, as the ZnO/ZrO2 ratio increased, the metallic Cu0 particles were gradually aggregated.
N2O-oxidation followed by H2 titration was used to calculate the SCu, DCu, and dCu (Table 1). As the ZnO/ZrO2 ratio increases, metallic Cu0 particles gradually aggregate, reasonably agreeing with the TEM and XRD analyses. Due to the presence of CuZn alloy in the sample, the Cu particle size calculated from the N2O titration is too large, so the results are considered principally for qualitative analysis.
The H2-TPR (H2-temperature-programmed reduction) analysis (Figure 3) for the calcined catalysts generated overlapped hydrogen consumption peaks with resultant asymmetric profiles in the 100–230 °C temperature range. In the catalysts with the ZnO/ZrO2 ratio between 0:7 and 5:2, the three peaks (denoted as α, β, and γ) suggest the presence of different CuO species. Within this range, as the ZnO/ZrO2 ratio increased, α peak intensity gradually decreased, indicating a decrease in content of smaller Cu particles. At the same time, β and γ peaks gradually became dominant. The reduction temperature first decreases and then increases as the ZnO/ZrO2 ratio increases, and the C3Z2Z5 catalyst shows the lowest reduction temperature. This indicates that when the ZnO/ZrO2 ratio is 2:5, the interaction between Cu and oxide support is strong, promoting the reduction of Cu species [39]. Only β and γ peaks were found in the C3Z6Z1 and C3Z7Z0 catalyst, which demonstrate the Cu species in the sample are mostly large CuO particles or exist in the form of bulk CuO. The theoretical H2 reduction consumption was calculated and compared with the actual H2 reduction consumption (Table 2). The ratio was recorded as RH2/Cu, with RH2/Cu > 1, indicating that ZnO was partially reduced to Zn0.
The chemical states of Cu, Zn, and Zr on the catalyst surface can be analyzed by XPS (Figure 4). In the Cu 2p XPS spectra of the reduced catalyst (Figure 4A), the signals at binding energy (BE) = 932.6 eV and 952.4 eV belong to the characteristic peaks of Cu 2p3/2 and Cu 2p1/2. No Cu2+ satellite peak near 943.0 eV was observed, indicating that the surface CuO species were completely reduced [40,41]. The Cu LMM Auger region was used to better distinguish between Cu0 and Cu+. However, there is a certain overlap in the characteristic peak positions of Cu and Zn, so only samples with a ZnO/ZrO2 ratio less than 2:5 were analyzed by Cu LMM. In Figure 4C, the characteristic peaks of Cu0 or Cu+ appeared at binding energies of 569.9 eV and 573.9 eV, but the content of Cu+ was relatively low. After reduction, most of the Cu on the catalyst surface existed in the form of Cu0.
The characteristic peaks of Zr 3d5/2 and Zr 3d3/2 at BE = 181.8 eV and 184.2 eV were observed in the reduced catalysts (Figure 4B). At the same time, the reduction of the catalysts did not cause a significant shift in the position of the Zr 3d characteristic peak, indicating that the surface ZrO2 was basically not reduced to metal Zr. It can be inferred that the excessive consumption of H2 in H2-TPR is mainly related to the reduction of ZnO. The valence state distribution of Zn in the catalysts after reduction was analyzed by Zn LMM (Figure 4D). In C3Z2Z5, C3Z5Z2, and C3Z7Z0 catalysts, the characteristic peaks of Zn appeared at Kinetic energy = 988.0 eV, 991.2 eV, and 992.4 eV, respectively, corresponding to Zn2+, Znq+, and Zn0. The presence of Zn0 further clarifies that some ZnO was reduced to metal Zn during the reduction process. As the ZnO/ZrO2 ratio increases, the content of Zn0 shows a trend of first increasing and then decreasing.
In the process of CO2 hydrogenation to methanol, the capacity of the catalyst adsorption/activation of CO2 and H2 plays an important role. In Cu/ZrO2 and Cu/ZnO/ZrO2, the difference in H2 adsorption capacity (based on per gram catalyst, Table 1) is not significant, which means that the ZnO/ZrO2 ratio has a small impact on H2 adsorption. In the Cu/ZnO catalyst (C3Z7Z0), the adsorption capacity of H2 is significantly lower, which is related to the larger size of Cu particles on ZnO. The larger Cu particle size reduces the adsorption sites of H2, causing a decrease in the adsorption capacity of H2. In the H2-TPD (H2 chemisorption) profile (Figure 5A) for the catalysts containing Zr, a main desorption peak of 380–420 °C can be observed, which is the desorption peak of H species bound by Cu-H bonds on metal Cu [42]. As the ZnO/ZrO2 ratio increases, the peak temperature firstly shifts towards low temperature and then remains basically unchanged, indicating that the formation of CuZn alloy will reduce the desorption temperature of H2. In Cu/ZnO (C3Z7Z0), the desorption temperature further decreases. In the range of the ZnO/ZrO2 ratio between 1:6 and 5:2, the high temperature peak of 500–600 °C in H2-TPD did not appear in the H2-MS spectra (Figure 5B), which may be related to the unknown species released from the catalyst decomposition under high-temperature conditions (higher than the catalyst calcination temperature).
A total of three main desorption signals were detected in the CO2-TPD analysis (Figure 6A), including the low-temperature (80–150 °C), medium-temperature (350–450 °C), and high-temperature (450–620 °C) ranges. The difference in ZnO/ZrO2 ratio affects the CO2 desorption amount and temperature within different temperature ranges. With the increase of the ZnO/ZrO2 ratio, the desorption amount in the low-temperature range decreases to a certain extent, but the desorption temperature does not significantly increase. In the middle-temperature range, the CO2 desorption amount is relatively low in each catalyst and the change is not significant, but the desorption temperature gradually increases. The CO2 desorption amount in the high-temperature zone reaches a maximum value on C3Z2Z5 and then decreases. The desorption temperature is basically the same in the samples containing Zn but is higher than that in C3Z0Z7. The CO2 desorption amount (Table 1) depends on the ZnO and ZrO2 composition. The value firstly increases and then decreases with increasing ZnO/ZrO2 ratio and the highest amount (0.311 mmol/gcat) is recorded at the ratio of 2:5.
The CO2 desorption peaks correspond to three basic sites on the catalyst surface, namely weak, moderate, and strong basic sites. At weak basic sites, CO2 adsorbs as bicarbonate and desorbs at low temperatures [6,39]. The moderate basic sites are generally metal-O pairs, which adsorb CO2 and form carbonate species [43]. At medium temperature, the carbonate intermediate is prone to RWGS, generating CO [44]. The species generated by CO2 adsorption on strongly basic sites are not yet clear, but it can be determined that the presence of strongly basic sites is related to the Cu-ZrO2 interface [29,45,46,47]. CO2 hydrogenation to methanol is generally operated at 200–300 °C. Therefore, the moderate and strong basic sites are effective active sites for CO2 adsorption/activation and further conversion. The difference in ZnO/ZrO2 ratio affects the types and quantities of basic sites on the carrier surface and changes the desorption amount and temperature of CO2.

2.2. Catalytic Test

Catalytic performance comparison of CuZn (C3Z7Z0), CuZr (C3Z0Z7), CuZnZr (C3Z2Z5), and commercial catalyst (CuZnAl) for the reaction is shown in Figure 7A. The CO2 conversion increases with the increasing temperature. The order of CO2 conversion follows CuZnAl > CuZnZr > CuZr > CuZn. Methanol selectivity is decreased with increasing temperature (Figure 7B). Except for CuZnAl, the methanol selectivity remains above 60% at 240 °C, indicating that the higher CO2 conversion over CuZnAl is due to the generation of more by-product CO. The methanol selectivity for the catalysts follows the order CuZr > CuZn ≈ CuZnZr > CuZnAl. STY of methanol was used to determine catalyst activity for methanol synthesis (Figure 7C). The STY of methanol rises with the increase in temperature (180–260 °C). CuZnZr shows the best catalytic performance in all four catalysts, which also ranks top in the existing literature (Table S1). In traditional CO2 hydrogenation to methanol catalysts, CuZn alloy [9,27,28] or Cu-ZnO interface [48] is generally considered as the active site. Here we can note that the CuZn sample displays the lowest Cu surface area with the smallest amount for reactant uptake, which can be considered to be responsible for the poor activity for methanol formation. Meanwhile, CuZr shows the highest dispersion and largest surface area of Cu and relatively high capacity for H2 and CO2 adsorption but displays a lower activity than CuZnZr (C3Z2Z5). This suggests the Cu surface area is not the exclusive factor in determining the catalyst activity for CO2 hydrogenation to methanol. Moreover, CuZr is more active than CuZn, indicating more effective active sites formed on CuZr. Based on the CO2-TPD results, the presence of the Cu-ZrO2 interface can effectively increase the CO2 adsorption capacity, thereby promoting the reaction. We consider that both ZnO and ZrO2 play an important role in working with Cu for the construction of the active sites over CuZnZr. As indicated by the XRD and TEM analyses, CuZn alloy is formed from strong Cu-ZnO interaction over C3Z2Z5. The C3Z2Z5 catalyst exhibits a higher capacity for CO2 adsorption relative to CuZn and CuZr, implying the involvement of the Cu-ZrO2 interface in catalyzing the reaction. It is possible that the cooperation between Cu-oxide interaction (e.g., CuZn alloy and Cu-ZrO2 interface) acts as the active sites, effectively improving carbon dioxide conversion rate and methanol selectivity. Since the composition of ZnO and ZrO2 impacts the Cu-oxide interaction, the following work further investigates the optimization of ZnO and ZrO2 contents and establishes the relationship between the catalyst structure and the activity.
The effect of varying ZnO/ZrO2 ratios on the catalytic performance in the reaction over Cu/ZnO/ZrO2 at 220 °C and 240 °C reveals the highest CO2 conversion is achieved at the ZnO/ZrO2 ratio of 2:5 to 3:4 (Figure 7D,E), suggesting an optimized content of ZnO and ZrO2 is required for efficient activation and conversion of CO2. The methanol selectivity does not vary dramatically with changes in the ZnO/ZrO2 ratio. Due to the larger change in CO2 conversion than in methanol selectivity, the overall trend of STY of methanol with the Zn/Zr ratio is similar to the trend of CO2 conversion (Figure 7F). Except for 220 °C, the STY of methanol of C3Z2Z5 is the highest at all other temperatures. The trend of activity variation is similar to the results of Zn LMM and CO2-TPD. The catalyst activity increased before the ZnO content increased to 20%, possibly due to the increase in CuZn alloy content. However, as the ZnO content increases, the ZrO2 content correspondingly decreases, and the Cu-ZrO2 interface gradually decreases, which may lead to a decrease in catalyst activity. In C3Z2Z5, the amount of CuZn alloy and Cu-ZrO2 interface reach the highest values, demonstrating the best catalytic activity.
The C3Z2Z5 catalyst was used to investigate the effect of reduction temperature on catalytic performance (Figure 8). When the reduction temperature exceeds 260 °C, there is no significant difference in CO2 conversion, methanol selectivity, and STY of methanol (Figure 8A–C), indicating that the catalyst had been completely reduced beyond 260 °C, which is consistent with the results of H2-TPR. Under the reaction conditions of 220, 240, and 260 °C, with the increase in reduction temperature, the conversion rate slightly increases and then remains unchanged, while the selectivity initially increases and then decreases (Figure 8D–F). When the reduction temperature is higher than 260 °C, the catalytic performance does not change significantly, indicating that a higher reduction temperature below 300 °C does not significantly alter the structure of the catalyst. The C3Z2Z5 catalyst was also selected to investigate the impact of gas hourly space velocity (GHSV, 6000, 12,000, 18,000, and 24,000 mL/gcat/h) on catalytic performance (Figure 9). Lower GHSV results in higher CO2 conversion but lower methanol selectivity, indicating that low GHSV is more conducive to the production of CO through the RWGS. The STY of methanol is also increased with the increase of GHSV, reaching a maximum value at 24,000 mL/gcat/h under the conditions used here.
In the long-term stability test for 110 h (P = 3 MPa, T = 240 °C, GHSV = 18,000 mL/gcat/h), the CO2 conversion and methanol selectivity over the C3Z2Z5 catalyst remain basically unchanged (Figure S4), and exhibit no significant deactivation. Comparing the XRD patterns of the catalyst after reduction and after 110 h reaction (Figure S4B), the intensities of the Cu peak (2θ = 43.3° and 50.4°) are similar, indicating that the C3Z2Z5 sample after 110 h reaction exhibits no severe sintering of Cu particles. TEM and EDS mapping of the spent catalyst (Figure S4C–F) verify that the long-term stability test does not cause aggregation of the Cu particles, indicating that the C3Z2Z5 catalyst exhibits good stability.
The adsorption capacity of catalysts on reactants will have a significant impact on the reaction performance. When the metal Cu content is fixed, except for the Cu/ZnO catalyst, there is not much difference in the H2 adsorption amount, but the change in the CO2 adsorption amount is more significant. The correlation analysis between the CO2 desorption amount and catalyst performance (Figure 10A) shows that the catalytic activity is directly proportional to the CO2 desorption amount, indicating that the ZnO/ZrO2 ratio affects the reaction performance by adjusting the CO2 adsorption sites. When the ZnO/ZrO2 ratio is 2:5, the number of strong basic sites on the catalyst surface is the highest, effectively improving the CO2 adsorption amount and enhancing the catalytic activity.
XRD, HR-TEM, and XPS analyses show that a fraction of ZnO is reduced to Zn0 and reacts with metallic Cu to form a CuZn alloy. The content of Zn0 in different samples can be calculated by Zn-LMM, which corresponds to the amount of CuZn alloy. By correlating the catalytic performance with the content of Zn0 (Figure 10B), C3Z2Z5 with the highest CuZn alloy content exhibits the highest activity, while the catalysts with lower Zn0 content also had lower catalytic performance, indicating that the presence of CuZn alloy effectively promotes the reaction. The change in the ZnO/ZrO2 ratio not only affects the amount of CuZn alloy but also affects the number of Cu-ZrO2 interfaces, thereby affecting the adsorption capacity of CO2. C3Z4Z3 and C3Z3Z4 have similar CuZn alloy content, but C3Z3Z4 exhibits better catalytic activity, which can be associated with an increase in the CO2 adsorption capacity. Thus, we consider the oxide (ZnO and ZrO2) composition mainly affects the catalytic performance in methanol synthesis by regulating the quantity of CuZn alloys and Cu-ZrO2 interaction, and furthers the ability for adsorption/activation of reactants. The synergistic sites associated with the CuZn alloy and Cu-ZrO2 interface in the optimized Cu/ZnO/ZrO2 catalyst (C3Z2Z5) reach the maximum, contributing to the enhanced production of methanol.

3. Materials and Methods

3.1. Catalyst Preparation

A series of Cu/ZnO/ZrO2 catalysts with varying Cu contents were prepared by coprecipitation. The mixed nitrates (including copper nitrate, zinc nitrate, and zirconium nitrate) in ethanol were precipitated at 70 °C using excess oxalic acid under stirring for 1 h, then aged for 4 h at room temperature. The precipitate was separated by centrifugation, washed three times with ethanol, and dried at 80 °C for 12 h. The dried samples were calcined at 450 °C for 4 h. After calcination, the samples were subjected to compression, crushing, and sieving treatment and 40–60 mesh particles were selected for the catalysts test. The target mass fraction of Cu was fixed at 30 wt%, and a total of 8 samples were prepared by sequentially adjusting the mass ratio of ZnO/ZrO2. The calcined catalysts were named based on element content.

3.2. Catalyst Characterization

Nitrogen physisorption was tested using a Micromeritics ASAP 2460 (Micromeritics instrument Ltd., Atlanta, America) system for analysis of the specific surface area (SBET) and pore structure (pore volume and pore size) according to the standard Braeuer–Emmett–Teller (BET) method and Barrett–Joyner–Halenda desorption branch. All the samples were outgassed under vacuum at 300 °C for 4 h before the nitrogen physisorption test.
XRD patterns of both calcined and reduced samples were carried out by using a Rigaku D/MAX 2500 X-ray diffractometer (Rigaku Corporation, Tokyo, Japan) using Cu Kα radiation (λ = 1.542 Å), and the results were analyzed using JADE6 (V6.5.26@07/02/05). All samples were measured over the range of 20° ≤ 2θ ≤ 80° with a step size of 0.02° at ambient temperature.
TEM and HAADF-STEM (high-angle annular dark-field scanning transmission electron microscopy) were used to observe the morphology of all the samples, which were tested by using a JEM-2100F electron microscope equipped with an EDX spectrometer (JEOL, Tokyo, Japan), and the obtained images were analyzed with a Gatan Digital Micrograph. All samples were sonicated in an ethanol solution and then dropped onto a holey carbon/Mo grid before measurements.
ICP-OES was used to measure the actual content of Cu, Zn, and Zr by using Vista-MPX (Agilent, Palo Alto, America). A certain mass of catalyst was dissolved in hydrofluoric acid, and the resultant solution was diluted using boric acid. The diluted solution was injected into the instrument through a peristaltic pump, and the average value of each element content was measured three times to reduce the error. The characteristic wavelengths selected for Cu, Zn, and Zr elements are 327.395 nm, 213.857 nm, and 349.619 nm.
XPS were obtained to analyze the valence states of surface elements by using a Thermo K-Alpha+ system (Thermo Fisher Scientific, Waltham, America), and Al Kα (E = 1486.6 eV) was applied for exciting the photoelectron spectra under ultra-high vacuum (6.67 × 10−7 Pa). All the binding energy was calibrated by C1s (284.6 eV), and the results were analyzed using Avantage software (V5.9921).
H2-TPR, H2-TPD, CO2-TPD, and N2O titration were conducted using a Micromeritics Chemisorb 2920 system (Micromeritics instrument Ltd., Atlanta, America). Subsequently, 50 mg of catalyst was loaded into a quartz U-tube and pretreated in Ar at 200 °C for 1 h to remove adsorbed water. After the pretreatment, samples were cooled to 30 °C, subjected to 10% H2/Ar (10 mL/min), and heated to 300 °C at 10 °C/min. For H2 chemisorption, the reduced samples were cooled to 200 °C, purged in Ar for 30 min, and subjected to 10% H2/Ar (10 mL/min) for pulse titration until saturation. Following H2 chemisorption, the samples were cooled to 30 °C and heated from 30 °C to 800 °C at a rate of 10 °C/min in Ar for H2-TPD analysis. For the CO2-TPD analysis, the reduced samples were purged in He at 300 °C for 30 min, cooled to 30 °C, subjected to 10% CO2/He (30 mL/min) for 1 h for CO2 adsorption, and heated to 800 °C at 10 °C/min in He. N2O titration was used to measure the exposed metal Cu surface area. The samples were reduced as mentioned above, purged in Ar at 300 °C for 30 min, cooled to 30 °C, and subjected to N2O titration at 30 °C for 30 min. The samples were purged in Ar again to remove the physisorbed N2O and heated to 300 °C, subjected to 10% H2/Ar (10 mL/min) for H2 titration until saturated. All the signals were recorded by a TCD (thermal conductivity detector). The DCu, SCu, and dCu were calculated according to the following equations:
N Cu 0 = 2 n H 2 m cat
D Cu = N Cu 0 M Cu W Cu
S Cu 0 = N Cu 0 N A S D Cu
d Cu = 6 M Cu D Cu ρ σ N A
where MCu is the atomic weight of copper, WCu is the Cu content determined by ICP-OES, SDCu is the copper surface density (1.47 × 1019 atoms/m2), ρ is the copper metal density (8.94 g/cm3), and σ is the area occupied by a surface copper atom (6.85 Å2/atom).
For ex situ characterizations (including nitrogen physisorption, XRD, TEM, and XPS), the calcined catalyst should be reduced at 300 °C for 3 h.

3.3. Catalyst Testing

The catalytic performance was evaluated using a fixed-bed continuous-flow stainless-steel reactor. The catalyst (100 mg, 40−60 mesh) was reduced in H2 (10 mL/min) for 3 h before the reaction and cooled down to 180 °C, which was monitored by a thermocouple in the catalyst bed. Afterward, the hydrogen gas was turned off and the feed gases (CO2/H2 = 3/1) and N2 (internal standard) were introduced, increasing the pressure to 3 MPa. Next, the reactor temperature was raised to the test temperature (180–300 °C). The composition of the outlet gas was analyzed online using a gas chromatographic system (Shimadzu, GC-2014) equipped with an FID (flame ionization detector, for methanol) and a TCD (thermal conductivity detector, for N2, CO2, CO, and CH4). The XCO2 (conversion of CO2), SCH3OH (methanol selectivity), and STYCH3OH (space-time yield of methanol) were calculated with the following equations:
X CO 2 = F ( CO ) out + F ( CH 3 OH ) out F ( CO 2 ) out + F ( CO ) out + F ( CH 3 OH ) out × 100 %
S CH 3 OH = F ( C H 3 OH ) out F ( CO ) out + F ( C H 3 OH ) out × 100 %
STY CH 3 OH = F ( CO 2 ) in · X CO 2 · S CH 3 OH · M CH 3 OH 22.4 · m
where F is the volumetric flow rate, MCH3OH is the molecule weight of methanol, and mcat refers to the mass of the catalyst. The deviation of the carbon balance is within 5%.

4. Conclusions

This study has established the effect of ZnO and ZrO2 contents on the Cu/ZnO/ZrO2 catalyst structure/surface characteristics and catalytic performance in the CO2 hydrogenation to methanol. CuZn alloy was formed in the catalyst containing ZnO, and the ZnO/ZrO2 ratio can affect the content of CuZn alloy. A correlation between the amount of CuZn alloy and the catalytic activity reveals the participation of CuZn alloy in the reaction. The ZnO/ZrO2 ratio also affects the amount of strong basic sites on the catalyst surface. With an increase in strong basic sites related to the Cu-ZrO2 interfaces, the amount of CO2 desorption significantly increases. The increase in both CuZn alloy content and Cu-ZrO2 interfaces promotes the process of CO2 hydrogenation to methanol. The optimized catalyst C3Z2Z5 shows the highest STY of methanol (610.8 gCH3OH/kgcat/h) with no significant decrease in activity during long-term stability testing.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13101337/s1, Figure S1: N2 adsorption-desorption isotherms (A) and pore size distributions (B); Figure S2: TEM images of reduced catalysts; Figure S3: EDS-mapping images of reduced catalysts; Figure S4: C3Z2Z5 (A) 110 h stability test, (B) XRD pattern after reduction and stability test, TEM image (C) after reduction and (D) stability test, EDS-mapping images (E) after reduction and (F) stability test; Table S1: Catalytic performance of Cu-based catalysts for CO2 hydrogenation to methanol [49,50,51,52,53,54,55,56,57,58,59].

Author Contributions

Conceptualization and methodology, X.C., X.Z., M.L. and X.M.; investigation, X.C., X.Z., X.H., J.C., Z.H., Z.Z., H.Z. and P.G.; data curation, X.C., X.Z. and M.L.; writing—original draft, X.C. and X.Z.; writing—review and editing, X.C., X.Z. and M.L.; supervision, M.L., J.L. (Jing Lv), and X.M.; project administration, J.L. (Jing Li), F.L. and X.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (22178265, U21B2096, 21938008) and Tianjin Natural Science and Technology Project (21JCYBJC00400).

Data Availability Statement

The original data are available from X.C., X.Z. and M.L.

Acknowledgments

The authors acknowledge the financial support by National Natural Science Foundation of China (22178265, U21B2096, 21938008), Tianjin Natural Science and Technology Project (21JCYBJC00400) and Beijing Building Materials Academy of Science Research.

Conflicts of Interest

The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Alwin, M.; Mathias, P. Synthetic Manufacture of Methanol. U.S. Patent No. 1,569,775, 12 January 1926. [Google Scholar]
  2. Cifre, P.G.; Badr, O. Renewable hydrogen utilisation for the production of methanol. Energy Convers. Manag. 2007, 48, 519–527. [Google Scholar] [CrossRef]
  3. Shih, C.F.; Zhang, T.; Li, J.; Bai, C. Powering the future with liquid sunshine. Joule 2018, 2, 1925–1949. [Google Scholar] [CrossRef]
  4. Wang, Z.-Q.; Xu, Z.-N.; Peng, S.-Y.; Zhang, M.-J.; Lu, G.; Chen, Q.-S.; Chen, Y.; Guo, G.-C. High-performance and long-lived Cu/SiO2 nanocatalyst for CO2 hydrogenation. ACS Catal. 2015, 5, 4255–4259. [Google Scholar] [CrossRef]
  5. Arena, F.; Italiano, G.; Barbera, K.; Bordiga, S.; Bonura, G.; Spadaro, L.; Frusteri, F. Solid-state interactions, adsorption sites and functionality of Cu-ZnO/ZrO2 catalysts in the CO2 hydrogenation to CH3OH. Appl. Catal. A Gen. 2008, 350, 16–23. [Google Scholar] [CrossRef]
  6. Dong, X.; Li, F.; Zhao, N.; Xiao, F.; Wang, J.; Tan, Y. CO2 hydrogenation to methanol over Cu/ZnO/ZrO2 catalysts prepared by precipitation-reduction method. Appl. Catal. B Environ. 2016, 191, 8–17. [Google Scholar] [CrossRef]
  7. Nakamura, J.; Uchijima, T.; Kanai, Y.; Fujitani, T. The role of ZnO in Cu/ZnO methanol synthesis catalysts. Catal. Today 1996, 28, 223–230. [Google Scholar] [CrossRef]
  8. Kattel, S.; Ramírez, P.J.; Chen, J.G.; Rodriguez, J.A.; Liu, P. Active sites for CO2 hydrogenation to methanol on Cu/ZnO catalysts. Science 2017, 355, 1296–1299. [Google Scholar] [CrossRef]
  9. Fujitani, T.; Nakamura, J. The chemical modification seen in the Cu/ZnO methanol synthesis catalysts. Appl. Catal. A Gen. 2000, 191, 111–129. [Google Scholar] [CrossRef]
  10. Lunkenbein, T.; Schumann, J.; Behrens, M.; Schlögl, R.; Willinger, M.G. Formation of a ZnO overlayer in industrial Cu/ZnO/Al2O3 catalysts induced by strong metal–support interactions. Angew. Chem. 2015, 127, 4627–4631. [Google Scholar] [CrossRef]
  11. Kuld, S.; Thorhauge, M.; Falsig, H.; Elkjær, C.F.; Helveg, S.; Chorkendorff, I.; Sehested, J. Quantifying the promotion of Cu catalysts by ZnO for methanol synthesis. Science 2016, 352, 969–974. [Google Scholar] [CrossRef]
  12. Behrens, M.; Studt, F.; Kasatkin, I.; Kühl, S.; Hävecker, M.; Abild-Pedersen, F.; Zander, S.; Girgsdies, F.; Kurr, P.; Kniep, B.-L. The active site of methanol synthesis over Cu/ZnO/Al2O3 industrial catalysts. Science 2012, 336, 893–897. [Google Scholar] [CrossRef]
  13. Bansode, A.; Tidona, B.; von Rohr, P.R.; Urakawa, A. Impact of K and Ba promoters on CO2 hydrogenation over Cu/Al2O3 catalysts at high pressure. Catal. Sci. Technol. 2013, 3, 767–778. [Google Scholar] [CrossRef]
  14. Noh, G.; Lam, E.; Alfke, J.L.; Larmier, K.; Searles, K.; Wolf, P.; Copéret, C. Selective hydrogenation of CO2 to CH3OH on supported Cu nanoparticles promoted by isolated TiIV surface sites on SiO2. ChemSusChem 2019, 12, 968–972. [Google Scholar] [CrossRef]
  15. Hartadi, Y.; Widmann, D.; Behm, R.J. CO2 hydrogenation to methanol on supported Au catalysts under moderate reaction conditions: Support and particle size effects. ChemSusChem 2015, 8, 456–465. [Google Scholar] [CrossRef]
  16. Yang, X.; Kattel, S.; Senanayake, S.D.; Boscoboinik, J.A.; Nie, X.; Graciani, J.; Rodriguez, J.A.; Liu, P.; Stacchiola, D.J.; Chen, J.G. Low pressure CO2 hydrogenation to methanol over gold nanoparticles activated on a CeO x/TiO2 interface. J. Am. Chem. Soc. 2015, 137, 10104–10107. [Google Scholar] [CrossRef]
  17. Köppel, R.A.; Stöcker, C.; Baiker, A. Copper-and silver–zirconia aerogels: Preparation, structural properties and catalytic behavior in methanol synthesis from carbon dioxide. J. Catal. 1998, 179, 515–527. [Google Scholar] [CrossRef]
  18. Collins, S.E.; Baltanas, M.A.; Bonivardi, A.L. An infrared study of the intermediates of methanol synthesis from carbon dioxide over Pd/β-Ga2O3. J. Catal. 2004, 226, 410–421. [Google Scholar] [CrossRef]
  19. Jiang, X.; Koizumi, N.; Guo, X.; Song, C. Bimetallic Pd–Cu catalysts for selective CO2 hydrogenation to methanol. Appl. Catal. B Environ. 2015, 170, 173–185. [Google Scholar] [CrossRef]
  20. Bahruji, H.; Bowker, M.; Hutchings, G.; Dimitratos, N.; Wells, P.; Gibson, E.; Jones, W.; Brookes, C.; Morgan, D.; Lalev, G. Pd/ZnO catalysts for direct CO2 hydrogenation to methanol. J. Catal. 2016, 343, 133–146. [Google Scholar] [CrossRef]
  21. Bai, S.; Shao, Q.; Feng, Y.; Bu, L.; Huang, X. Highly efficient carbon dioxide hydrogenation to methanol catalyzed by zigzag platinum–cobalt nanowires. Small 2017, 13, 1604311. [Google Scholar] [CrossRef]
  22. Martin, O.; Martín, A.J.; Mondelli, C.; Mitchell, S.; Segawa, T.F.; Hauert, R.; Drouilly, C.; Curulla-Ferré, D.; Pérez-Ramírez, J. Indium oxide as a superior catalyst for methanol synthesis by CO2 hydrogenation. Angew. Chem. 2016, 128, 6369–6373. [Google Scholar] [CrossRef]
  23. Chen, T.-Y.; Cao, C.; Chen, T.-B.; Ding, X.; Huang, H.; Shen, L.; Cao, X.; Zhu, M.; Xu, J.; Gao, J. Unraveling highly tunable selectivity in CO2 hydrogenation over bimetallic In-Zr oxide catalysts. ACS Catal. 2019, 9, 8785–8797. [Google Scholar] [CrossRef]
  24. Men, Y.-L.; Liu, Y.; Wang, Q.; Luo, Z.-H.; Shao, S.; Li, Y.-B.; Pan, Y.-X. Highly dispersed Pt-based catalysts for selective CO2 hydrogenation to methanol at atmospheric pressure. Chem. Eng. Sci. 2019, 200, 167–175. [Google Scholar] [CrossRef]
  25. Shen, C.; Sun, K.; Zhang, Z.; Rui, N.; Jia, X.; Mei, D.; Liu, C.-J. Highly active Ir/In2O3 catalysts for selective hydrogenation of CO2 to methanol: Experimental and theoretical studies. ACS Catal. 2021, 11, 4036–4046. [Google Scholar] [CrossRef]
  26. 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]
  27. Kuld, S.; Conradsen, C.; Moses, P.G.; Chorkendorff, I.; Sehested, J. Quantification of zinc atoms in a surface alloy on copper in an industrial-type methanol synthesis catalyst. Angew. Chem. Int. Ed. 2014, 53, 5941–5945. [Google Scholar] [CrossRef]
  28. Tisseraud, C.; Comminges, C.; Belin, T.; Ahouari, H.; Soualah, A.; Pouilloux, Y.; Le Valant, A. The Cu–ZnO synergy in methanol synthesis from CO2, Part 2: Origin of the methanol and CO selectivities explained by experimental studies and a sphere contact quantification model in randomly packed binary mixtures on Cu–ZnO coprecipitate catalysts. J. Catal. 2015, 330, 533–544. [Google Scholar] [CrossRef]
  29. Chang, X.; Han, X.; Pan, Y.; Hao, Z.; Chen, J.; Li, M.; Lv, J.; Ma, X. Insight into the Role of Cu–ZrO2 Interaction in Methanol Synthesis from CO2 Hydrogenation. Ind. Eng. Chem. Res. 2022, 61, 6872–6883. [Google Scholar] [CrossRef]
  30. Natesakhawat, S.; Lekse, J.W.; Baltrus, J.P.; Ohodnicki, P.R., Jr.; Howard, B.H.; Deng, X.; Matranga, C. Active sites and structure–activity relationships of copper-based catalysts for carbon dioxide hydrogenation to methanol. ACS Catal. 2012, 2, 1667–1676. [Google Scholar] [CrossRef]
  31. Cargnello, M.; Doan-Nguyen, V.V.; Gordon, T.R.; Diaz, R.E.; Stach, E.A.; Gorte, R.J.; Fornasiero, P.; Murray, C.B. Control of metal nanocrystal size reveals metal-support interface role for ceria catalysts. Science 2013, 341, 771–773. [Google Scholar] [CrossRef]
  32. Zhang, Z.; Jing, M.; Chen, H.; Okejiri, F.; Liu, J.; Leng, Y.; Liu, H.; Song, W.; Hou, Z.; Lu, X. Transfer hydrogenation of fatty acids on Cu/ZrO2: Demystifying the role of carrier structure and metal–support interface. ACS Catal. 2020, 10, 9098–9108. [Google Scholar] [CrossRef]
  33. Yu, J.; Yang, M.; Zhang, J.; Ge, Q.; Zimina, A.; Pruessmann, T.; Zheng, L.; Grunwaldt, J.-D.; Sun, J. Stabilizing Cu+ in Cu/SiO2 catalysts with a shattuckite-like structure boosts CO2 hydrogenation into methanol. ACS Catal. 2020, 10, 14694–14706. [Google Scholar] [CrossRef]
  34. Feng, Y.; Li, Z.; Liu, H.; Dong, C.; Wang, J.; Kulinich, S.A.; Du, X. Laser-prepared CuZn alloy catalyst for selective electrochemical reduction of CO2 to ethylene. Langmuir 2018, 34, 13544–13549. [Google Scholar] [CrossRef] [PubMed]
  35. Yuan, L.; Wang, C.; Cai, R.; Wang, Y.; Zhou, G. Spontaneous ZnO nanowire formation during oxidation of Cu-Zn alloy. J. Appl. Phys. 2013, 114, 023512. [Google Scholar] [CrossRef]
  36. Li, K.; Chen, J.G. CO2 hydrogenation to methanol over ZrO2-containing catalysts: Insights into ZrO2 induced synergy. ACS Catal. 2019, 9, 7840–7861. [Google Scholar] [CrossRef]
  37. Wang, Y.; Kattel, S.; Gao, W.; Li, K.; Liu, P.; Chen, J.G.; Wang, H. Exploring the ternary interactions in Cu–ZnO–ZrO2 catalysts for efficient CO2 hydrogenation to methanol. Nat. Commun. 2019, 10, 1166. [Google Scholar] [CrossRef]
  38. Frei, E.; Gaur, A.; Lichtenberg, H.; Zwiener, L.; Scherzer, M.; Girgsdies, F.; Lunkenbein, T.; Schlögl, R. Cu−Zn alloy formation as unfavored state for efficient methanol catalysts. ChemCatChem 2020, 12, 4029–4033. [Google Scholar] [CrossRef]
  39. Witoon, T.; Chalorngtham, J.; Dumrongbunditkul, P.; Chareonpanich, M.; Limtrakul, J. CO2 hydrogenation to methanol over Cu/ZrO2 catalysts: Effects of zirconia phases. Chem. Eng. J. 2016, 293, 327–336. [Google Scholar] [CrossRef]
  40. Liang, Z.-Q.; Zhuang, T.-T.; Seifitokaldani, A.; Li, J.; Huang, C.-W.; Tan, C.-S.; Li, Y.; De Luna, P.; Dinh, C.T.; Hu, Y. Copper-on-nitride enhances the stable electrosynthesis of multi-carbon products from CO2. Nat. Commun. 2018, 9, 3828. [Google Scholar] [CrossRef]
  41. Rather, R.A.; Singh, S.; Pal, B. A Cu+1/Cu0-TiO2 mesoporous nanocomposite exhibits improved H2 production from H2O under direct solar irradiation. J. Catal. 2017, 346, 1–9. [Google Scholar] [CrossRef]
  42. Wang, W.; Qu, Z.; Song, L.; Fu, Q. An investigation of Zr/Ce ratio influencing the catalytic performance of CuO/Ce1-xZrxO2 catalyst for CO2 hydrogenation to CH3OH. J. Energy Chem. 2020, 47, 18–28. [Google Scholar] [CrossRef]
  43. 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]
  44. Choi, E.; Song, K.; An, S.; Lee, K.; Youn, M.; Park, K.; Jeong, S.; Kim, H. Cu/ZnO/AlOOH catalyst for methanol synthesis through CO2 hydrogenation. Korean J. Chem. Eng. 2018, 35, 73–81. [Google Scholar] [CrossRef]
  45. Gao, P.; Li, F.; Zhao, N.; Xiao, F.; Wei, W.; Zhong, L.; Sun, Y. Influence of modifier (Mn, La, Ce, Zr and Y) on the performance of Cu/Zn/Al catalysts via hydrotalcite-like precursors for CO2 hydrogenation to methanol. Appl. Catal. A Gen. 2013, 468, 442–452. [Google Scholar] [CrossRef]
  46. Han, X.; Li, M.; Chang, X.; Hao, Z.; Chen, J.; Pan, Y.; Kawi, S.; Ma, X. Hollow structured Cu@ZrO2 derived from Zr-MOF for selective hydrogenation of CO2 to methanol. J. Energy Chem. 2022, 71, 277–287. [Google Scholar] [CrossRef]
  47. Zhao, H.B.; Yu, R.F.; Ma, S.C.; Xu, K.Z.; Chen, Y.; Jiang, K.; Fang, Y.; Zhu, C.X.; Liu, X.C.; Tang, Y.; et al. The role of Cu1-O3 species in single-atom Cu/ZrO2 catalyst for CO2 hydrogenation. Nat. Catal. 2022, 5, 818–831. [Google Scholar] [CrossRef]
  48. Choi, Y.; Futagami, K.; Fujitani, T.; Nakamura, J. The role of ZnO in Cu/ZnO methanol synthesis catalysts—Morphology effect or active site model? Appl. Catal. A Gen. 2001, 208, 163–167. [Google Scholar] [CrossRef]
  49. Lam, E.; Larmier, K.; Wolf, P.; Tada, S.; Safonova, O.V.; Coperet, C. Isolated Zr surface sites on silica promote hydrogenation of CO2 to CH3OH in supported Cu catalysts. J. Am. Chem. Soc. 2018, 140, 10530–10535. [Google Scholar] [CrossRef]
  50. Larmier, K.; Liao, W.C.; Tada, S.; Lam, E.; Verel, R.; Bansode, A.; Urakawa, A.; Comas-Vives, A.; Coperet, C. CO2-to-methanol hydrogenation on zirconia-supported copper nanoparticles: Reaction intermediates and the role of the metal-support interface. Angew. Chem. Int. Ed. 2017, 56, 2318–2323. [Google Scholar] [CrossRef]
  51. Tada, S.; Fujiwara, K.; Yamamura, T.; Nishijima, M.; Uchida, S.; Kikuchi, R. Flame spray pyrolysis makes highly loaded Cu nanoparticles on ZrO2 for CO2-to-methanol hydrogenation. Chem. Eng. J. 2020, 381, 122750. [Google Scholar] [CrossRef]
  52. Shohei, T.; Shingo, K.; Tetsuo, H.; Hiromu, K.; Akane, N.; Kenichi, K.; Takashi, T.; Ken-Ichi, S.; Shigeo, S. Design of interfacial sites between Cu and amorphous ZrO2 dedicated to CO2-to-methanol hydrogenation. ACS Catal. 2018, 8, 7809–7819. [Google Scholar]
  53. Amenomiya, Y. Methanol synthesis from CO2 + H2 II. Copper-based binary and ternary catalysts. Appl. Catal. 1987, 30, 57–68. [Google Scholar] [CrossRef]
  54. Haowang, Y.; Guigao, W.; Wang, H.; Zheng, Y.E.; Na, W.; Zhaili, K. Structure–activity relationships of Cu–ZrO2 catalysts for CO2 hydrogenation to methanol: Interaction effects and reaction mechanism. RSC Adv. 2017, 7, 8709–8717. [Google Scholar]
  55. Guo, X.; Mao, D.; Lu, G.; Wang, S.; Wu, G. Glycine–nitrate combustion synthesis of CuO–ZnO–ZrO2 catalysts for methanol synthesis from CO2 hydrogenation. J. Catal. 2010, 271, 178–185. [Google Scholar] [CrossRef]
  56. Sánchez-Contador, M.; Ateka, A.; Rodriguez-Vega, P.; Bilbao, J.; Aguayo, A.T. Optimization of the Zr content in the CuO-ZnO-ZrO2 /SAPO-11 catalyst for the selective hydrogenation of CO+CO2 mixtures in the direct synthesis of dimethyl ether. Ind. Eng. Chem. Res. 2018, 57, 1169–1178. [Google Scholar] [CrossRef]
  57. Arena, F.; Mezzatesta, G.; Zafarana, G.; Trunfio, G.; Frusteri, F.; Spadaro, L. Effects of oxide carriers on surface functionality and process performance of the Cu–ZnO system in the synthesis of methanol via CO2 hydrogenation. J. Catal. 2013, 300, 141–151. [Google Scholar] [CrossRef]
  58. Ramli, M.Z.; Syed-Hassan, S.S.A.; Hadi, A. Performance of Cu-Zn-Al-Zr catalyst prepared by ultrasonic spray precipitation technique in the synthesis of methanol via CO2 hydrogenation. Fuel Process. Technol. 2018, 169, 191–198. [Google Scholar] [CrossRef]
  59. Zhang, Y.; Zhong, L.; Wang, H.; Gao, P.; Li, X.; Xiao, S.; Ding, G.; Wei, W.; Sun, Y. Catalytic performance of spray-dried Cu/ZnO/Al2O3/ZrO2 catalysts for slurry methanol synthesis from CO2 hydrogenation. J. CO2 Util. 2016, 15, 72–82. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of calcined (A) and reduced (B,C) catalysts.
Figure 1. XRD patterns of calcined (A) and reduced (B,C) catalysts.
Catalysts 13 01337 g001
Figure 2. HR-TEM images of reduced catalysts.
Figure 2. HR-TEM images of reduced catalysts.
Catalysts 13 01337 g002
Figure 3. H2-TPR profiles for all the catalysts.
Figure 3. H2-TPR profiles for all the catalysts.
Catalysts 13 01337 g003
Figure 4. (A) Cu 2p XPS, (B) Zr 3d XPS, (C) Cu LMM, and (D) Zn LMM Auger spectra of reduced catalysts.
Figure 4. (A) Cu 2p XPS, (B) Zr 3d XPS, (C) Cu LMM, and (D) Zn LMM Auger spectra of reduced catalysts.
Catalysts 13 01337 g004
Figure 5. (A) The H2-TPD, (B) MS-H2 profiles.
Figure 5. (A) The H2-TPD, (B) MS-H2 profiles.
Catalysts 13 01337 g005
Figure 6. (A) The CO2-TPD, (B) MS-CO2 profiles.
Figure 6. (A) The CO2-TPD, (B) MS-CO2 profiles.
Catalysts 13 01337 g006
Figure 7. (A) CO2 conversion, (B) methanol selectivity, (C) STY of methanol of CuZn, CuZr, CuZnZr, and commercial CuZnAl catalysts. The variation of CO2 conversion, product selectivity in the reaction at (D) 220 °C and (E) 240 °C, and (F) STY of methanol for all catalysts. (P = 3 MPa, GHSV = 18,000 mL/gcat/h).
Figure 7. (A) CO2 conversion, (B) methanol selectivity, (C) STY of methanol of CuZn, CuZr, CuZnZr, and commercial CuZnAl catalysts. The variation of CO2 conversion, product selectivity in the reaction at (D) 220 °C and (E) 240 °C, and (F) STY of methanol for all catalysts. (P = 3 MPa, GHSV = 18,000 mL/gcat/h).
Catalysts 13 01337 g007
Figure 8. (A) CO2 conversion, (B) methanol selectivity, (C) STY of methanol of C3Z2Z5 at different reduction temperatures, the variation of CO2 conversion, product selectivity in the reaction at (D) 220 °C, (E) 240 °C, and (F) 260 °C at different reduction temperatures (P = 3 MPa, GHSV = 18,000 mL/gcat/h).
Figure 8. (A) CO2 conversion, (B) methanol selectivity, (C) STY of methanol of C3Z2Z5 at different reduction temperatures, the variation of CO2 conversion, product selectivity in the reaction at (D) 220 °C, (E) 240 °C, and (F) 260 °C at different reduction temperatures (P = 3 MPa, GHSV = 18,000 mL/gcat/h).
Catalysts 13 01337 g008
Figure 9. (A) CO2 conversion and methanol selectivity, (B) STY of methanol of C3Z2Z5 at different GHSV (P = 3 MPa, T = 240 °C).
Figure 9. (A) CO2 conversion and methanol selectivity, (B) STY of methanol of C3Z2Z5 at different GHSV (P = 3 MPa, T = 240 °C).
Catalysts 13 01337 g009
Figure 10. Relationship between STY of methanol (P = 3 MPa, T = 240 °C, GHSV = 18,000 mL/gcat/h) and (A) CO2 desorption amount, (B) Zn0 content.
Figure 10. Relationship between STY of methanol (P = 3 MPa, T = 240 °C, GHSV = 18,000 mL/gcat/h) and (A) CO2 desorption amount, (B) Zn0 content.
Catalysts 13 01337 g010
Table 1. Physicochemical properties of the catalysts.
Table 1. Physicochemical properties of the catalysts.
SamplesCu
(wt%) a
SBET
(m2/g)
Pore Volume
(cm3/g)
Pore Size
(mm)
DCu
(%) b
dCu
(nm) b
SCu
(m2/gcat) b
SCu
(m2/gCu) b
H2 Uptake
(mmol/gcat)
CO2 Desorption
(mmol/gcat)
C3Z0Z728.6295.50.4113.07.1114.613.1045.790.01470.244
C3Z1Z628.8994.40.4013.14.7421.98.8230.520.01470.278
C3Z2Z530.8088.70.4415.24.1225.28.1826.570.01480.311
C3Z3Z430.7380.50.4517.13.6028.97.1323.210.01390.250
C3Z4Z328.5471.50.5121.92.7338.05.0217.620.01410.203
C3Z5Z228.1971.00.4920.03.1233.35.6720.120.01550.180
C3Z6Z130.2270.70.5021.02.7238.35.2917.520.01610.117
C3Z7Z031.1330.70.2728.60.973.130.00920.061
a Derived from ICP-OES. b Derived from N2O titration.
Table 2. H2 consumption during the TPR analysis.
Table 2. H2 consumption during the TPR analysis.
SamplesTemperatureActual H2
Reduction Consumption
Theoretical H2
Reduction Consumption
RH2/Cu
°Cmmol/gmmol/g
C3Z0Z7184.64.48 4.20 1.07
C3Z1Z6172.94.49 4.24 1.06
C3Z2Z5162.4 4.86 4.50 1.08
C3Z3Z4179.94.99 4.48 1.11
C3Z4Z3182.1 5.35 4.18 1.28
C3Z5Z2163.25.34 4.14 1.29
C3Z6Z1163.15.50 4.42 1.25
C3Z7Z0164.6 5.89 4.54 1.30
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

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. https://doi.org/10.3390/catal13101337

AMA Style

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(10):1337. https://doi.org/10.3390/catal13101337

Chicago/Turabian Style

Chang, Xiao, Xiaohui Zi, Jing Li, Fengdong Liu, Xiaoyu Han, Jiyi Chen, Ziwen Hao, Heng Zhang, Zhenmei Zhang, Pengju Gao, and et al. 2023. "An Insight into Synergistic Metal-Oxide Interaction in CO2 Hydrogenation to Methanol over Cu/ZnO/ZrO2" Catalysts 13, no. 10: 1337. https://doi.org/10.3390/catal13101337

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