Next Article in Journal
Effect of NaOH Concentration on Rapidly Quenched Cu–Al Alloy-Derived Cu Catalyst for CO2 Hydrogenation to CH3OH
Previous Article in Journal
Influence of Oxidation Temperature on the Regeneration of a Commercial Pt-Sn/Al2O3 Propane Dehydrogenation Catalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploring the Effect of the Solvothermal Time on the Structural Properties and Catalytic Activity of Cu-ZnO-ZrO2 Catalysts Synthesized by the Solvothermal Method for CO2 Hydrogenation to Methanol

School of Chemical and Environmental Engineering, Shanghai Institute of Technology, Shanghai 201418, China
*
Author to whom correspondence should be addressed.
Submission received: 27 May 2024 / Revised: 12 June 2024 / Accepted: 15 June 2024 / Published: 18 June 2024

Abstract

:
A series of Cu-ZnO-ZrO2 (CCZ) catalysts were prepared by the solvothermal method with different solvothermal times (1 h, 3 h, 6 h, and 12 h). The physicochemical properties of these catalysts and the catalytic performance for CO2 hydrogenation to methanol were studied. The highest methanol yield was achieved when the solvothermal time was 6 h (CCZ-6). Furthermore, we found that the copper surface area (SCu) increases and then decreases with an increase in the solvothermal time and that there is a strong correlation between the methanol yield and the SCu. This research highlights the crucial influence of the solvothermal time on the structure and catalytic behavior of Cu-ZnO-ZrO2 catalysts, providing a valuable reference for the development of efficient catalysts.

Graphical Abstract

1. Introduction

With the continuous growth of global energy demand and the increasing severity of environmental problems, the exploration of renewable energy sources and the reduction of greenhouse gas emissions have become important issues that need to be urgently addressed [1]. Among the greenhouse gases, the conversion and utilization of carbon dioxide (CO2) has received much attention [2,3]. Methanol (CH3OH) has gradually gained recognition as a clean and renewable fuel source by global industry due to its widespread availability, significant economic volume, and sustainable development of the whole industrial chain [4,5]. Therefore, among the many ways to convert CO2, the conversion of CO2 to methanol by the hydrogenation reaction is a prospective solution [6,7,8].
Copper-based catalysts have become a popular research topic due to their low cost and high catalytic activity for the CO2 hydrogenation to methanol reaction [9,10]. Among them, Cu-ZnO-ZrO2 ternary catalysts have attracted much attention due to their exceptional catalytic performance [11,12,13]. Witoon et al. [13] claimed that the doping of graphene oxide significantly improved the catalytic performance of CuO-ZnO-ZrO2 catalysts. Chang et al. [14] demonstrated that the ZnO/ZrO2 composition in Cu/ZnO/ZrO2 catalyst impacted CuZn alloy formation and surface basic sites, and the suitable ZnO/ZrO2 composition significantly enhanced the catalytic activity and the methanol yield produced by CO2 hydrogenation.
It has been shown that the preparation methods and conditions of catalysts have a significant effect on their structures and properties [15]. As we all know, the most commonly used method for the preparation of methanol synthesis catalysts is co-precipitation. The co-precipitation method involves the simultaneous precipitation of multiple components from a solution, leading to well-mixed and homogeneously distributed precursors [7,11]. However, the co-precipitation process is very sensitive to the conditions of the pH, temperature, and stirring speed of the solution, and a little carelessness may lead to the non-uniform particle size distribution of the precipitates, which may in turn affect the performance of the catalyst. The solvothermal method, as an alternative effective synthesis method, can prepare nanomaterials with high dispersibility and excellent catalytic properties under relatively mild conditions [16,17]. Specifically, the time of the solvothermal treatment, as a key parameter in the solvothermal method, plays a crucial role in regulating the physicochemical properties of the catalyst [18,19]. A previous study [16] has shown that the solvothermal temperature has a significant effect on the structure of Cu-ZnO-ZrO2 catalysts, and CCZ-180 with a solvothermal temperature of 180 °C had the best catalytic activity. However, the effect of the solvothermal time on the catalyst performance is equally important and needs to be further studied.
In this paper, a series of Cu-ZnO-ZrO2 catalysts were prepared by the solvothermal method under different solvothermal times (1 h, 3 h, 6 h, and 12 h). The catalysts were analyzed via characterizations such as XRD, SEM, EDX mapping, N2 adsorption–desorption, N2O chemisorption, XPS, H2-TPR, and CO2/H2-TPD. Furthermore, the relationships between the catalytic performance and the physicochemical properties were studied in detail.

2. Results and Discussion

2.1. Characterization

2.1.1. XRD

Figure 1 shows the XRD patterns of the calcined and reduced Cu-ZnO-ZrO2 catalysts. From Figure 1a, the diffraction peaks at 2θ of 35.5°, 38.7°, 48.8°, 61.6°, 66.3°, 68.1°, and 75.0°, respectively, can be attributed to the presence of the CuO phase (PDF #48-1548), whereas the diffraction peaks appearing at 2θ = 30.3° correspond to tetragonal ZrO2 (t-ZrO2) (PDF #050-1089). No diffraction peak of ZnO was found in any of the four catalysts, indicating that ZnO was highly dispersed or existed in an amorphous state in all four catalysts [20,21].
As shown in Figure 1b, after reduction at 300 °C, several peaks were observed at 2θ of 43.3°, 50.4°, and 74.1°, representing the crystalline surfaces of Cu(111), Cu(200), and Cu(220), respectively [22]. However, no peaks of CuO were found, suggesting that the CuO in these samples had been reduced [23]. The sizes of the metallic Cu and CuO crystals were determined by Scherer’s formula and the results are summarized in Table 1. It can be observed that the sizes of both CuO and metallic Cu first decrease and then increase with the extension of the solvothermal time. Especially, CZZ-6 exhibited the smallest size of copper species, which improved the dispersion and increased the surface area, as evidenced by the maximum SBET and SCu of CZZ-6. On the other hand, the sizes of metallic Cu are noticeably larger than those of CuO for the same catalysts, and the most plausible reason for this phenomenon is that the reduction of CuO to metallic Cu with H2 is an exothermic process (CuO + H2 → Cu + H2O, ΔH = −129.2 kJ/mol), and the heat released makes the metallic Cu particles aggregate and become larger. This phenomenon has also been reported in previous studies [17,24].

2.1.2. SEM and EDX Mapping

Figure 2 shows SEM images of the representative CZZ-6 sample. It can be seen that CZZ-6 is a built-up structure consisting of many particles that form a loose microstructure. An EDX mapping analysis was performed to investigate the distribution of the elements. It was confirmed that all the Cu, Zn, and Zr atoms in CCZ-6 are well-dispersed on the catalyst surface.

2.1.3. N2 Adsorption–Desorption

The N2 adsorption–desorption isotherms and pore size distribution curves for the different catalysts are shown in Figure 3. Clearly, all the samples showed type IV isotherms with H3-type hysteresis loops (Figure 3a), indicating that all the catalysts were mesoporous materials [25,26]. Furthermore, the hysteresis return line of CZZ-6 was larger than that of the other catalysts, indicating that it had the largest mesopore volume. As can be seen in Figure 3b, all four samples exhibited wide pore distributions, except for the concentration of pores at 1.9 nm, where the pore sizes gradually increased in the order of CZZ-3 < CZZ-6 < CZZ-1 < CZZ-12. These results indicate that the CZZ-3 catalyst had the narrowest pore distribution among the four catalysts.
The effects of the solvothermal time on the pore structure and specific surface area (SBET) of the Cu-ZnO-ZrO2 catalysts are listed in Table 1. As the solvothermal time increases, the SBET and pore volume of the catalyst initially increase and then decrease. This result is consistent with the order of CuO particle size measured by XRD (Table 1). The CZZ-6 catalyst, with a solvothermal time of 6 h, possesses the largest SBET and pore volume. A larger pore volume and SBET could enhance the adsorption and desorption of the reaction gases, thereby improving the catalytic performance [27].

2.1.4. SCu

The SCu of Cu-based catalysts has an important effect on the catalytic performance [12,27,28], so the N2O chemisorption method was employed to determine the SCu of the different Cu-ZnO-ZrO2 catalysts. As shown in Table 1, with the extension of the solvothermal time, the SCu first increases and then decreases. This trend is consistent with the change in the SBET of the catalyst, with both reaching their maximum values in the CZZ-6 catalyst prepared with a solvothermal time of 6 h. The subsequent activity test results (Section 2.2) revealed that the larger the SCu, the better the catalytic activity that could be achieved. Especially, the CZZ-6 had the largest SCu, implying more highly dispersed copper particles [27,29], which was confirmed by its smallest copper particle size (dCu) in Table 1.

2.1.5. XPS

Figure 4 presents the XPS spectra of the copper species after calcination and in situ reduction. As depicted in Figure 4a, the peaks of the copper species appeared around 933 eV and 953 eV, with broad vibrational peaks appearing within 940–945 eV, indicating the presence of copper species in the form of Cu2+ [30]. Following in situ reduction at 300 °C (Figure 4b), the vibrational peaks disappeared and the signal peak of the copper species became narrower and shifted to lower binding energies at around 932 eV, suggesting the reduction of Cu2+ to Cu0 and/or Cu+ [31].
The Cu LMM XAES spectra of the Cu species were used to further distinguish between Cu0 and Cu+. From Figure 5, it can be observed that the CZZ-1 and CZZ-12 catalysts exhibited two overlapping peaks, indicating the coexistence of Cu+ and Cu0 species on the catalyst surface [32]. In contrast, only Cu0 species were observed at the binding energy of 919.0 eV for CZZ-3 and CZZ-6. On the other hand, the characteristic peaks of Cu+ were not observed in CZZ-1 and CZZ-12 according to the XRD analysis (Figure 1), possibly due to the amorphous nature of Cu2O or its high dispersion.
Figure 6 shows the O 1s XPS spectra of the Cu-ZnO-ZrO2 catalysts after in situ reduction. Three peaks were present on the spectra of all the catalysts: the α peak at 529 eV represented the lattice oxygen of the metal oxides, the β peak was the chemisorbed oxygen species, and the γ peak was the hydroxyl-like species [33]. Among them, the β and γ peaks were considered to be active sites and played an important role in CO2 hydrogenation to methanol [34,35]. From Table S1, the largest Aβ + Aγ value was found in CZZ-6, which provided the foundation for the high activity.
The binding energy values and surface contents of the catalyst surfaces before and after in situ reduction are illustrated in Table S2. From Table S2, it can be seen that the binding energy values of Zn 2p3/2 were in the range of 1021.3–1021.5 eV, and those of Zr 3d5/2 were in the range of 181.7–181.9 eV, and there was no significant change in the binding energy values of Zn 2p3/2 and Zr 3d5/2 before and after in situ reduction, which indicated that Zn and Zr existed in the catalyst in the form of oxides and would not be reduced at 300 °C under a hydrogen atmosphere. Moreover, with the extension of the solvothermal time, the copper content on the catalyst surface followed the order: CZZ-1 > CZZ-3 > CZZ-6 > CZZ-12, while the SCu order was: CZZ-12 < CZZ-1 < CZZ-3 < CZZ-6. Generally, the larger the copper content on the catalyst surface, the smaller the SCu. Clearly, in our study, there was no direct correlation between the copper content on the catalyst surface and the SCu. This discrepancy may be due to the aggregation of copper particles on the surface of the CZZ-12 with a prolonged solvothermal time, resulting in the smallest SCu.

2.1.6. H2-TPR

The reduction properties of the Cu-ZnO-ZrO2 catalysts were studied by H2-TPR. As shown in Figure 7, all the catalysts showed a broad reduction peak between 150 and 300 °C. As mentioned earlier, ZnO and ZrO2 were not reduced at 300 °C, and therefore, the peak was attributed to the reduction of the CuOx species [7].
The small shoulder peak (α peak) on the low-temperature side can be clearly observed in the CZZ-1, and this result indicated that there were two different sizes of CuO particles in the CZZ-1. Unlike CZZ-1, the reduction peaks of the CZZ-3, CZZ-6, and CZZ-12 catalysts were relatively symmetrical, indicating that the CuO crystal sizes on these three catalysts were more uniform than those on CZZ-1. Furthermore, CZZ-6 was observed to have the lowest reduction temperature (202 °C) for CuOx species. Low-temperature reduction usually indicates strong interactions between the metal and the oxide in the catalyst, which can enhance the dispersion of the metal species and make the reduction process easier to carry out [36,37,38]. Table 2 displays the quantitative data of the hydrogen consumption of the CuO-ZnO-ZrO2 catalysts. It is observed that CZZ-1 and CZZ-3 have similar H2 consumption, while CZZ-6 and CZZ-12 have more H2 consumption, which suggests the presence of more reducible CuOx species in CZZ-6 and CZZ-12. Furthermore, the H2 consumption of all the catalysts was lower than the theoretical value required for the complete reduction of copper oxide to the metal Cu (the degree of the reduction was less than 1), further confirming that ZnO and ZrO2 were not reduced. It has been well-documented that Cu-based catalysts with high activity for CO2 hydrogenation to methanol have high reducibility [39,40]. CZZ-6 was expected to exhibit the highest activity due to its optimal reducibility among the four catalysts, as confirmed by the results of the activity tests (Section 2.2).

2.1.7. CO2-TPD

The adsorption capacity of the catalysts for CO2 was determined through CO2-TPD. Figure 8 illustrates the CO2-TPD profiles of the pre-reduced Cu-ZnO-ZrO2 catalysts. It is observed that all the catalysts exhibited three CO2 desorption peaks, which were deconvolved Gaussian peaks designated as α, β, and γ peaks, respectively. The low-temperature α peak corresponds to the desorption of CO2 from weak basic sites associated with hydroxyl functional groups. The β peak represents the desorption of CO2 from medium basic sites formed by the combination of metal and oxygen. The γ peak corresponds to the desorption of CO2 from strong basic sites associated with low-coordination oxygen anions [16]. The CO2 desorption data for different Cu-ZnO-ZrO2 catalysts are presented in Table 3. It can be observed that the CZZ-3 and CZZ-6 catalysts exhibit larger CO2 desorption (Aα + Aβ + Aγ values), but the strength of the basic sites of the CZZ-6 is relatively weak compared to CZZ-3. The total desorption of CO2 is crucial for the production of methanol by CO2 hydrogenation [25], which will be discussed in Section 2.3.

2.1.8. H2-TPD

Figure 9 displays the H2-TPD profiles of the pre-reduced Cu-ZnO-ZrO2 catalysts. The desorption peaks on the catalysts represent the H species adsorbed on the surface of the active component Cu [15,27]. It is generally observed that the low-temperature α peak represents the desorption of undissociated molecular hydrogen (H2) weakly adsorbed on the catalyst surface, while the high-temperature β peak represents the desorption of dissociated atomic hydrogen (H) [41]. The quantitative data for the temperature and peak area of the H2 desorption peaks are listed in Table 4. Clearly, the temperatures of the α and β desorption peaks show irregular changes with the prolongation of the solvothermal time. However, the lowest desorption peak temperature of CZZ-6 indicates the weakest adsorption strength of H2. On the other hand, with the extension of the solvothermal time, the peak areas of the α and β peaks first increased and then decreased, with the largest peak area observed at 6 h of solvothermal time. The maximum peak area of CZZ-6 indicates its highest H2 adsorption capacity.

2.2. Catalytic Activity

Figure 10 presents the activity evaluation results of the Cu-ZnO-ZrO2 catalysts prepared with different solvothermal times in the CO2 hydrogenation to methanol. Under the current reaction conditions, CO and methanol were detected as carbon-containing reaction products. However, with the extension of the solvothermal time, the CO2 conversion and methanol yield first increased and then decreased. CZZ-1 and CZZ-12 had low CO2 conversion despite the high methanol selectivity, resulting in low methanol yields. The CO2 conversion of CZZ-6 was 15.6%, with methanol selectivity of 46.1%, and the methanol yield reached a maximum of 7.2%. This indicates that the catalyst prepared with 6 h of solvothermal time is most suitable for methanol production.
To provide a comprehensive context for our catalytic activity results, we compared our findings with similar Cu-Zn-Zr catalysts reported in the literature for CO2 hydrogenation to methanol. The comparison is summarized in Table S3 in the Supplementary Materials. It is well known that the performance of Cu-Zn-Zr catalysts for CO2 hydrogenation heavily depends on factors such as the composition, pressure, temperature, H2/CO2 ratio, and feed gas space velocity [27,31]. Nevertheless, under similar reaction conditions, our CCZ-6 catalyst demonstrated comparable or even superior methanol yields. Our study indicates that optimizing the reaction time during solvothermal synthesis can effectively enhance the catalytic performance for CO2 hydrogenation to methanol.

2.3. Structure–Activity Relationship Analysis

As mentioned in relation to CO2-TPD, the amount of CO2 adsorbed by the different catalysts was observed to be clearly different. The adsorption of CO2 is a pre-condition for conducting CO2 hydrogenation [25,27]. The total CO2 adsorption is an important index in the evaluation of high-performance catalysts. Therefore, the effect of the total amount of desorbed CO2 on the methanol yield was investigated (Figure 11a). However, the relationship was not completely linear (R2 = 0.79), which indicated that the amount of CO2 absorbed was not directly responsible for the methanol yield.
On the other hand, the SCu stands as a crucial parameter for copper-based catalysts in methanol synthesis. Researchers have extensively explored the relationship between the catalytic activity of copper-based catalysts and the SCu [12,42,43]. However, there are conflicting views about their relationship. Natesakhawa et al. [42] reported a linear correlation between the methanol yield and the SCu. Our previous study [27] also demonstrated a strong link between the SCu and CO2 conversion over Cu-Ce1-xZrxO2 catalysts in CO2 hydrogenation to methanol. Conversely, Zhang et al. [43] and Hou et al. [44] found no connection between the methanol yield and the SCu. In this study, the solvothermal time significantly impacted the SCu of the Cu-ZnO-ZrO2 catalysts (Table 1). Figure 11b shows the relationship between the SCu and the methanol yield. It can be seen that there is a strong linear correlation between the methanol yield and the SCu (R2 = 0.95), indicating that the SCu of the Cu-ZnO-ZrO2 catalyst is a determining factor in methanol production. From the above analyses, it is clear that a suitable solvothermal time can promote the uniform growth and high dispersion of Cu particles, and thus increasing the SCu. This can be confirmed by the results of the XRD, SEM, EDX mapping, N2O chemisorption, and N2 adsorption–desorption, as described above.
It is well known that the solvothermal time greatly affects the properties of materials synthesized by the solvothermal method [18,19,45]. During solvothermal synthesis, the reaction time determines the growth rate of the catalyst particles and the interaction between different components. An optimal reaction time can lead to the formation of highly dispersed and well-crystallized catalysts with a large specific surface area, enhancing their catalytic performance. However, an insufficient or excessive reaction time can result in catalysts with poor activity due to incomplete crystallization or excessive particle growth. In our case, as the solvothermal time increased from 1 h to 6 h, the size of the formed CuO particles decreased, as observed from the XRD findings (Figure 1). As a result, the catalysts exhibited a larger SBET, as indicated by the N2 adsorption–desorption (Figure 3 and Table 1), and the SCu increased noticeably (Table 1). As the solvothermal time further increased to 12 h, nevertheless, small particles of CuO began to aggregate together and larger particles were formed, as evidenced by the data in Table 1. Consequently, the SCu was distinctly decreased. Evidently, the CZZ-6 catalyst possesses the largest SCu, which exhibits a maximal capacity for dissociative adsorption of H2 (Table 4) and thus the highest CO2 conversion and methanol yield (Figure 10) among the investigated CZZ-X catalysts. Figure 12 illustrates the effect of the solvothermal time for CO2 hydrogenation to methanol.

3. Materials and Methods

3.1. Catalyst Preparation

In this paper, Cu-ZnO-ZrO2 catalysts with the same molar ratio of Cu/Zn/Zr = 5:2:3 were prepared by the solvothermal method. The choice of this specific molar ratio was made to ensure comparability with our prior studies [16,17] and to build upon the successful results achieved earlier. Firstly, Cu(NO3)2·3H2O (AR), Zn(NO3)2·6H2O (AR), and Zr(NO3)4·5H2O (AR) were dissolved in 50 mL of ethylene glycol to prepare a solution with a total ion concentration of 1 mol/L. After stirring and dissolving, the solution was transferred to a reaction kettle and placed in an oven at 180 °C for a certain period (1 h, 3 h, 6 h, and 12 h). After the reaction, the kettle was cooled to room temperature, and the obtained precipitate was filtered, washed thoroughly with anhydrous ethanol and deionized water, and dried overnight in an oven at 120 °C. The yields of the solid products obtained at different solvothermal times were not noticeably different. Finally, the dried product was calcined at 450 °C for 3 h in a muffle furnace. Both the heating rate during the solvothermal process and the calcination process were set to 5 °C/min. The obtained catalysts were named as CZZ-X, where X represents the solvothermal synthesis time.

3.2. Characterization

X-ray diffraction analysis (XRD) of the catalysts was conducted using a Bruker D8 ADVANCE diffractometer (Karlsruhe, Germany) with Cu Kα radiation. The scanning range was set from 10° to 80° at a scan rate of 10°/min.
Scanning Electron Microscopy (SEM) and (Energy-dispersive X-ray spectroscopy) (EDX) mapping data of samples were gained on Hitachi SU8010 (Tokyo, Japan).
N2 adsorption–desorption experiments were performed using an ASAP 2020 HD88 instrument (Norcross, GA, USA). The specific surface area and pore volume of the samples were calculated by the Brunauer–Emmett–Teller (BET) and Barrett–Joyner–Halenda (BJH) models, respectively.
The copper surface area (SCu) was determined using two-step H2-TPR. Initially, 0.1 g of catalyst was purged with N2 at 300 °C for 60 min. The first reduction was then performed in a 10 vol% H2/N2 mixture to 300 °C. After the reduction, the sample was cooled to 60 °C, purged with N2 for 30 min, oxidized in 2 vol% N2O/He for 1 h to ensure the complete oxidation of surface metallic Cu to Cu2O, and then purged with N2 for another 30 min. Finally, the catalyst was again reduced to 500 °C in the 10 vol% H2/N2 mixture. The SCu was calculated using the formula:
S C u = 2 X × N 1.46 × 1 0 19 × W  
where X is the amount of H2 consumed during the second TPR process, N is Avogadro’s number (6.02 × 1023 atoms/mol), and W is the weight of the catalyst (g).
X-ray photoelectron spectroscopy (XPS) analysis was conducted using an ESCALAB 250Xi spectrometer from Thermo Scientific (Waltham, MA, USA), with Al Kα as the excitation source, and the binding energy of the measured elements was calibrated to C 1 s (284.6 eV).
The H2 temperature-programmed reduction (H2-TPR) was used to evaluate the reducibility of the catalysts. The catalyst (30 mg) was pretreated in N2 at 300 °C for 1 h to remove the surface-adsorbed water and impurities, then cooled to 50 °C. Subsequently, the sample was exposed to the 10 vol% H2/N2 mixture until the chromatographic baseline stabilized, followed by heating to 600 °C while monitoring the consumption of H2 with a thermal conductivity detector (TCD).
The temperature-programmed desorption of H2 and CO2 (H2/CO2-TPD) was used to characterize the adsorption property of the catalysts. Prior to CO2 adsorption, the catalyst was reduced at 300 °C for 1 h in a 10 vol% H2/N2 mixture. The sample was then cooled to 50 °C, saturated with 10 vol% CO2/N2 for 30 min, purged with He for 30 min to remove physically adsorbed CO2 molecules, and subjected to CO2-TPD under He flow from 50 °C to 600 °C, with CO2 signal detection using a mass spectrometer (Pfeiffer Vacuum Quadstar, 32-bit, Aßlar, Germany). The procedures for the H2-TPD and CO2-TPD were similar, with the only difference being the replacement of the adsorbed 10 vol% CO2/N2 gas to 10 vol% H2/N2, and the signal detection was performed using TCD. It should be noted that the TPD experiments for CO2 and H2 were conducted on different charges of the sample to ensure independent and accurate analysis of each gas’s adsorption and desorption behavior. Fresh samples were used for each TPD experiment, allowing for precise monitoring of the desorbed species without interference from previous gas exposures.

3.3. Activity Tests

A schematic diagram of the reaction apparatus for CO2 hydrogenation to methanol is illustrated in Figure S1. Different catalysts for CO2 hydrogenation were evaluated for their activity and selectivity using a continuous-flow fixed-bed reactor. Here, 0.3 g of catalyst with a particle size of 40–60 mesh was loaded into a stainless-steel reactor with an inner diameter of 5 mm. The catalyst was reduced in a 10 vol% H2/N2 stream at 300 °C for 3 h. Subsequently, the reaction was conducted at 240 °C, a pressure of 3 MPa, and a space velocity of 2400 mL/(gcat·h). After a stable reaction period, the products were analyzed using gas chromatography. Under the experimental conditions, the products of the CO2 hydrogenation reaction included CO and methanol, along with unreacted CO2 and N2. No other by-products were detected. The effluent from the reaction was analyzed online using an Agilent 6820 gas chromatograph. Methanol was detected using a flame ionization detector (FID) with a Porapak Q capillary column, while CO, CO2, and N2 were detected using a TCD equipped with a carbon molecular sieve column. The CO2 conversion, product selectivity, and methanol yield were calculated using the following formulas:
C O 2   c o n v e r s i o n   ( % ) = A C O f C O + A C H 3 O H f C H 3 O H A C O f C O + A C H 3 O H f C H 3 O H + A C O 2 f C O 2
M e t h a n o l   s e l e c t i v i t y % = A C H 3 O H f C H 3 O H A C O f C O + A C H 3 O H f C H 3 O H
C O   s e l e c t i v i t y % = A C O f C O A C O f C O + A C H 3 O H f C H 3 O H
M e t h a n o l   y i e l d % = C O 2   c o n v e r s i o n × M e t h a n o l   s e l e c t i v i t y × 100
In the above formulas, Ai denotes the integrated peak area of each substance in the gas chromatographic analysis at the reactor outlet and fi denotes the correction factor of each substance with respect to N2.

4. Conclusions

The Cu-ZnO-ZrO2 catalysts were prepared by the solvothermal method. The effects of different solvothermal times (1 h, 3 h, 6 h, and 12 h) on the physicochemical properties of the Cu-ZnO-ZrO2 catalysts were investigated. The results of the catalytic activity test revealed that the CZZ-6 exhibited the highest catalytic activity and methanol yield at a solvothermal time of 6 h. It can be observed that the solvothermal treatment time directly affected the size of the catalyst particles and the surface area, which in turn affected the surface properties and the distribution of active sites. The high activity of the CZZ-6 is ascribed to a suitable solvothermal time that promotes the uniform growth and high dispersion of Cu particles and increases the SCu. This work provides an in-depth study of the regulation of the solvothermal time on Cu-ZnO-ZrO2 catalysts prepared by the solvothermal method, which provides valuable guidance for high-performance catalysts.

Supplementary Materials

The following supporting information can be downloaded at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal14060390/s1, Figure S1: Schematic diagram of the reaction apparatus for CO2 hydrogenation to methanol; Table S1: Peak area parameters of the different Cu-ZnO-ZrO2 catalysts; Table S2: XPS parameters of the different Cu-ZnO-ZrO2 catalysts; Table S3: Comparison of the activity of the Cu-Zn-Zr catalysts [7,11,12,15,21,24,46,47,48,49,50,51,52,53,54].

Author Contributions

Conceptualization, J.H. and Y.L.; methodology, J.H. and Y.L.; validation, G.W. and J.Y.; writing—original draft preparation, J.H. and Y.L.; writing—review and editing, J.H. and D.M.; visualization, J.H.; supervision, G.W. and J.Y.; project administration, D.M.; funding acquisition, D.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Shanghai Municipal Science and Technology Commission of China, grant number 23010504600.

Data Availability Statement

Data are contained within the article or Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Alvarez, A.; Bansode, A.; Urakawa, A.; Bavykina, A.V.; Wezendonk, T.A.; Makkee, M.; Gascon, J.; Kapteijn, F. Challenges in the greener production of formates/formic acid, methanol, and DME by heterogeneously catalyzed CO2 hydrogenation processes. Chem. Rev. 2017, 117, 9804–9838. [Google Scholar] [CrossRef]
  2. Porosoff, M.D.; Yan, B.; Chen, J.G. Catalytic reduction of CO2 by H2 for synthesis of CO, methanol and hydrocarbons: Challenges and opportunities. Energy Environ. Sci. 2016, 9, 62–73. [Google Scholar] [CrossRef]
  3. Yang, Z.; Guo, D.; Dong, S.; Wu, J.; Zhu, M.; Han, Y.-F.; Liu, Z.-W. Catalysis for CO2 hydrogenation—What we have learned/should Learn from the hydrogenation of syngas to methanol. Catalysts 2023, 13, 1452. [Google Scholar] [CrossRef]
  4. Dang, S.; Yang, H.; Gao, P.; Wang, H.; Li, X.; Wei, W.; Sun, Y. A review of research progress on heterogeneous catalysts for methanol synthesis from carbon dioxide hydrogenation. Catal. Today 2019, 330, 61–75. [Google Scholar] [CrossRef]
  5. Zhao, L.; Zhang, L.; Wu, Z.; Huang, C.; Chen, K.; Wang, H.; Yang, F. Size effect of Cu particles on interface formation in Cu/ZnO catalysts for methanol synthesis. Catalysts 2023, 13, 1190. [Google Scholar] [CrossRef]
  6. 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]
  7. Witoon, T.; Kachaban, N.; Donphai, W.; Kidkhunthod, P.; Faungnawakij, K.; Chareonpanich, M.; Limtrakul, J. Tuning of catalytic CO2 hydrogenation by changing composition of CuO–ZnO–ZrO2 catalysts. Energy Convers. Manag. 2016, 118, 21–31. [Google Scholar] [CrossRef]
  8. Kordus, D.; Widrinna, S.; Timoshenko, J.; Lopez Luna, M.; Rettenmaier, C.; Chee, S.W.; Ortega, E.; Karslioglu, O.; Kuhl, S.; Roldan Cuenya, B. Enhanced methanol synthesis from CO2 hydrogenation achieved by tuning the Cu-ZnO interaction in ZnO/Cu2O nanocube catalysts supported on ZrO2 and SiO2. J. Am. Chem. Soc. 2024, 146, 8677–8687. [Google Scholar] [CrossRef] [PubMed]
  9. Zabilskiy, M.; Sushkevich, V.L.; Palagin, D.; Newton, M.A.; Krumeich, F.; van Bokhoven, J.A. The unique interplay between copper and zinc during catalytic carbon dioxide hydrogenation to methanol. Nat. Commun. 2020, 11, 2409. [Google Scholar] [CrossRef]
  10. Marcos, F.C.F.; Cavalcanti, F.M.; Petrolini, D.D.; Lin, L.; Betancourt, L.E.; Senanayake, S.D.; Rodriguez, J.A.; Assaf, J.M.; Giudici, R.; Assaf, E.M. Effect of operating parameters on H2/CO2 conversion to methanol over Cu-Zn oxide supported on ZrO2 polymorph catalysts: Characterization and kinetics. Chem. Eng. J. 2022, 427, 130947. [Google Scholar] [CrossRef]
  11. Bonura, G.; Cordaro, M.; Cannilla, C.; Arena, F.; Frusteri, F. The changing nature of the active site of Cu-Zn-Zr catalysts for the CO2 hydrogenation reaction to methanol. Appl. Catal. B Environ. 2014, 152–153, 152–161. [Google Scholar] [CrossRef]
  12. 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]
  13. Witoon, T.; Numpilai, T.; Phongamwong, T.; Donphai, W.; Boonyuen, C.; Warakulwit, C.; Chareonpanich, M.; Limtrakul, J. Enhanced activity, selectivity and stability of a CuO-ZnO-ZrO2 catalyst by adding graphene oxide for CO2 hydrogenation to methanol. Chem. Eng. J. 2018, 334, 1781–1791. [Google Scholar] [CrossRef]
  14. 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]
  15. Chen, D.; Mao, D.; Xiao, J.; Guo, X.; Yu, J. CO2 hydrogenation to methanol over CuO–ZnO–TiO2–ZrO2: A comparison of catalysts prepared by sol–gel, solid-state reaction and solution-combustion. J. Sol-Gel Sci. Technol. 2018, 86, 719–730. [Google Scholar] [CrossRef]
  16. Liang, Y.; Mao, D.; Guo, X.; Yu, J.; Wu, G.; Ma, Z. Solvothermal preparation of CuO-ZnO-ZrO2 catalysts for methanol synthesis via CO2 hydrogenation. J. Taiwan Inst. Chem. Eng. 2021, 121, 81–91. [Google Scholar] [CrossRef]
  17. Liang, Y.; Han, J.; Yu, J.; Wu, G.; Mao, D. Methanol synthesis from CO2 hydrogenation on CuO–ZnO–ZrO2 prepared by solvothermal method: The influence of solvent on catalyst properties and catalytic behavior. Top. Catal. 2023, 66, 1503–1514. [Google Scholar] [CrossRef]
  18. Yang, Y.; Liang, Y.; Zhang, Z.; Zhang, Y.; Wu, H.; Hu, Z. Morphology well-controlled synthesis of NiO by solvothermal reaction time and their morphology-dependent pseudocapacitive performances. J. Alloys Compd. 2016, 658, 621–628. [Google Scholar] [CrossRef]
  19. Ni, X.; Zhang, J.; Zhao, L.; Wang, F.; He, H.; Dramou, P. Study of the solvothermal method time variation effects on magnetic iron oxide nanoparticles (Fe3O4) features. J. Phys. Chem. Solids 2022, 169, 110855. [Google Scholar] [CrossRef]
  20. Wang, G.; Mao, D.; Guo, X.; Yu, J. Enhanced performance of the CuO-ZnO-ZrO2 catalyst for CO2 hydrogenation to methanol by WO3 modification. Appl. Surf. Sci. 2018, 456, 403–409. [Google Scholar] [CrossRef]
  21. 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]
  22. Samson, K.; Śliwa, M.; Socha, R.P.; Góra-Marek, K.; Mucha, D.; Rutkowska-Zbik, D.; Paul, J.F.; Ruggiero-Mikołajczyk, M.; Grabowski, R.; Słoczyński, J. Influence of ZrO2 structure and copper electronic state on activity of Cu/ZrO2 catalysts in methanol synthesis from CO2. ACS Catal. 2014, 4, 3730–3741. [Google Scholar] [CrossRef]
  23. Shi, L.; Shen, W.; Yang, G.; Fan, X.; Jin, Y.; Zeng, C.; Matsuda, K.; Tsubaki, N. Formic acid directly assisted solid-state synthesis of metallic catalysts without further reduction: As-prepared Cu/ZnO catalysts for low-temperature methanol synthesis. J. Catal. 2013, 302, 83–90. [Google Scholar] [CrossRef]
  24. Huang, C.; Mao, D.; Guo, X.; Yu, J. Microwave-assisted hydrothermal synthesis of CuO–ZnO–ZrO2 as catalyst for direct synthesis of nethanol by carbon dioxide hydrogenation. Energy Technol. 2017, 5, 2100–2107. [Google Scholar] [CrossRef]
  25. Liu, H.; Huang, W.; Yu, Z.; Wang, X.; Jia, Y.; Huang, M.; Yang, H.; Li, R.; Wei, Q.; Zhou, Y. High-performance CuMgAl catalysts derived from hydrotalcite for CO2 hydrogenation to methanol: Effects of Cu-MgO interaction. Mol. Catal. 2024, 558, 114002. [Google Scholar] [CrossRef]
  26. Xue, H.; Guo, X.; Mao, D.; Meng, T.; Yu, J.; Ma, Z. Phosphotungstic acid-modified MnOx for selective catalytic reduction of NOx with NH3. Catalysts 2022, 12, 1248. [Google Scholar] [CrossRef]
  27. Han, J.; Yu, J.; Xue, Z.; Wu, G.; Mao, D. Highly efficient CO2 hydrogenation to methanol over Cu–Ce1−xZrxO2 catalysts prepared by an eco-friendly and facile solid-phase grinding method. Renew. Energy 2024, 222, 119951. [Google Scholar] [CrossRef]
  28. Wang, H.; Zhang, G.; Fan, G.; Yang, L.; Li, F. Fabrication of Zr–Ce oxide solid solution surrounded Cu-based catalyst assisted by a microliquid film reactor for efficient CO2 hydrogenation to produce methanol. Ind. Eng. Chem. Res. 2021, 60, 16188–16200. [Google Scholar] [CrossRef]
  29. Sripada, P.; Kimpton, J.; Barlow, A.; Williams, T.; Kandasamy, S.; Bhattacharya, S. Investigating the dynamic structural changes on Cu/CeO2 catalysts observed during CO2 hydrogenation. J. Catal. 2020, 381, 415–426. [Google Scholar] [CrossRef]
  30. Hu, X.; Mao, D.; Yu, J.; Xue, Z. Low-temperature CO oxidation on CuO-CeO2-ZrO2 catalysts prepared by a facile surfactant-assisted grinding method. Fuel 2023, 340, 127529. [Google Scholar] [CrossRef]
  31. Chen, G.; Yu, J.; Li, G.; Zheng, X.; Mao, H.; Mao, D. Cu+-ZrO2 interfacial sites with highly dispersed copper nanoparticles derived from Cu@UiO-67 hybrid for efficient CO2 hydrogenation to methanol. Int. J. Hydrogen Energy 2023, 48, 2605–2616. [Google Scholar] [CrossRef]
  32. Cui, X.; Yan, W.; Yang, H.; Shi, Y.; Xue, Y.; Zhang, H.; Niu, Y.; Fan, W.; Deng, T. Preserving the active Cu–ZnO interface for selective hydrogenation of CO2 to dimethyl ether and methanol. ACS Sustain. Chem. Eng. 2021, 9, 2661–2672. [Google Scholar] [CrossRef]
  33. 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]
  34. Wang, W.; Zhang, X.; Guo, M.; Li, J.; Peng, C. An investigation of the CH3OH and CO selectivity of CO2 hydrogenation over Cu−Ce−Zr catalysts. Front. Chem. Sci. Eng. 2022, 16, 950–962. [Google Scholar] [CrossRef]
  35. Wang, W.; Qu, Z.; Song, L.; Fu, Q. CO2 hydrogenation to methanol over Cu/CeO2 and Cu/ZrO2 catalysts: Tuning methanol selectivity via metal-support interaction. J. Energy Chem. 2020, 40, 22–30. [Google Scholar] [CrossRef]
  36. 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] [PubMed]
  37. Mao, D.; Zhang, H.; Zhang, J.; Wu, D. The influence of the compositions and structures of Cu-ZrO2 catalysts on the catalytic performance of CO2 hydrogenation to CH3OH. Chem. Eng. J. 2023, 471, 144605. [Google Scholar] [CrossRef]
  38. Song, L.; Wang, H.; Wang, S.; Qu, Z. Dual-site activation of H2 over Cu/ZnAl2O4 boosting CO2 hydrogenation to methanol. Appl. Catal. B Environ. 2023, 322, 122137. [Google Scholar] [CrossRef]
  39. An, B.; Zhang, J.; Cheng, K.; Ji, P.; Wang, C.; Lin, W. Confinement of ultrasmall Cu/ZnOx nanoparticles in metal–organic frameworks for selective methanol synthesis from catalytic hydrogenation of CO2. J. Am. Chem. Soc. 2017, 139, 3834–3840. [Google Scholar] [CrossRef]
  40. Dasireddy, V.D.B.C.; Likozar, B. The role of copper oxidation state in Cu/ZnO/Al2O3 catalysts in CO2 hydrogenation and methanol productivity. Renew. Energy 2019, 140, 452–460. [Google Scholar] [CrossRef]
  41. 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]
  42. Natesakhawat, S.; Lekse, J.W.; Baltrus, J.P.; Ohodnicki, P.R.; 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]
  43. Zhang, C.; Wang, L.; Etim, U.J.; Song, Y.; Gazit, O.M.; Zhong, Z. Oxygen vacancies in Cu/TiO2 boost strong metal-support interaction and CO2 hydrogenation to methanol. J. Catal. 2022, 413, 284–296. [Google Scholar] [CrossRef]
  44. Hou, X.-X.; Xu, C.-H.; Liu, Y.-L.; Li, J.-J.; Hu, X.-D.; Liu, J.; Liu, J.-Y.; Xu, Q. Improved methanol synthesis from CO2 hydrogenation over CuZnAlZr catalysts with precursor pre-activation by formaldehyde. J. Catal. 2019, 379, 147–153. [Google Scholar] [CrossRef]
  45. Huang, J.F.; Zhu, J.; Cao, L.Y.; Fei, J.; Wu, J.P. Influence of solvothermal time on oxidation resistance of carbon/carbon composites modified by borate sol. Surf. Eng. 2012, 28, 351–356. [Google Scholar] [CrossRef]
  46. Raudaskoski, R.; Niemelä, M.V.; Keiski, R.L. The effect of ageing time on co-precipitated Cu/ZnO/ZrO2 catalysts used in methanol synthesis from CO2 and H2. Top. Catal. 2007, 45, 57–60. [Google Scholar] [CrossRef]
  47. Zou, T.; Araújo, T.P.; Krumeich, F.; Mondelli, C.; Pérez-Ramírez, J. ZnO-promoted inverse ZrO2–Cu catalysts for CO2-based methanol synthesis under mild conditions. ACS Sustain. Chem. Eng. 2021, 10, 81–90. [Google Scholar] [CrossRef]
  48. Angelo, L.; Girleanu, M.; Ersen, O.; Serra, C.; Parkhomenko, K.; Roger, A.-C. Catalyst synthesis by continuous coprecipitation under micro-fluidic conditions: Application to the preparation of catalysts for methanol synthesis from CO2/H2. Catal. Today 2016, 270, 59–67. [Google Scholar] [CrossRef]
  49. Arena, F.; Barbera, K.; Italiano, G.; Bonura, G.; Spadaro, L.; Frusteri, F. Synthesis, characterization and activity pattern of Cu–ZnO/ZrO2 catalysts in the hydrogenation of carbon dioxide to methanol. J. Catal. 2007, 249, 185–194. [Google Scholar] [CrossRef]
  50. Ding, Z.; Xu, Y.; Yang, Q.; Hou, R. Pd-modified CuO–ZnO–ZrO2 catalysts for CH3OH synthesis from CO2 hydrogenation. Int. J. Hydrogen Energy 2022, 47, 24750–24760. [Google Scholar] [CrossRef]
  51. Ma, Y.; Sun, Q.; Wu, D.; Fan, W.-H.; Zhang, Y.-L.; Deng, J.-F. A practical approach for the preparation of high activity Cu/ZnO/ZrO2 catalyst for methanol synthesis from CO2 hydrogenation. Appl. Catal. A Gen. 1998, 171, 45–55. [Google Scholar] [CrossRef]
  52. Li, L.; Mao, D.; Yu, J.; Guo, X. Highly selective hydrogenation of CO2 to methanol over CuO–ZnO–ZrO2 catalysts prepared by a surfactant-assisted co-precipitation method. J. Power Sources 2015, 279, 394–404. [Google Scholar] [CrossRef]
  53. Marcos, F.C.F.; Lin, L.; Betancourt, L.E.; Senanayake, S.D.; Rodriguez, J.A.; Assaf, J.M.; Giudici, R.; Assaf, E.M. Insights into the methanol synthesis mechanism via CO2 hydrogenation over Cu-ZnO-ZrO2 catalysts: Effects of surfactant/Cu-Zn-Zr molar ratio. J. CO2 Util. 2020, 41, 101215. [Google Scholar] [CrossRef]
  54. Li, Z.; Du, T.; Li, Y.; Jia, H.; Wang, Y.; Song, Y.; Fang, X. Water-and reduction-free preparation of oxygen vacancy rich Cu-ZnO-ZrO2 catalysts for promoted methanol synthesis from CO2. Fuel 2022, 322, 124264. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the different Cu-ZnO-ZrO2 catalysts: (a) calcined, and (b) reduced.
Figure 1. XRD patterns of the different Cu-ZnO-ZrO2 catalysts: (a) calcined, and (b) reduced.
Catalysts 14 00390 g001
Figure 2. SEM and EDX mapping images of CCZ-6.
Figure 2. SEM and EDX mapping images of CCZ-6.
Catalysts 14 00390 g002
Figure 3. N2 adsorption–desorption isotherms (a) and the corresponding pore size distribution curves (b) of the different Cu-ZnO-ZrO2 catalysts.
Figure 3. N2 adsorption–desorption isotherms (a) and the corresponding pore size distribution curves (b) of the different Cu-ZnO-ZrO2 catalysts.
Catalysts 14 00390 g003
Figure 4. Cu 2p XPS spectra of the (a) calcined and (b) in situ reduced catalysts.
Figure 4. Cu 2p XPS spectra of the (a) calcined and (b) in situ reduced catalysts.
Catalysts 14 00390 g004
Figure 5. Cu LM2 XAES spectra of the in situ reduced catalysts.
Figure 5. Cu LM2 XAES spectra of the in situ reduced catalysts.
Catalysts 14 00390 g005
Figure 6. O 1s spectra of the in situ reduced catalysts.
Figure 6. O 1s spectra of the in situ reduced catalysts.
Catalysts 14 00390 g006
Figure 7. H2-TPR patterns of the various catalysts.
Figure 7. H2-TPR patterns of the various catalysts.
Catalysts 14 00390 g007
Figure 8. CO2-TPD profiles of the pre-reduced catalysts.
Figure 8. CO2-TPD profiles of the pre-reduced catalysts.
Catalysts 14 00390 g008
Figure 9. H2-TPD profiles of the pre-reduced catalysts.
Figure 9. H2-TPD profiles of the pre-reduced catalysts.
Catalysts 14 00390 g009
Figure 10. Catalytic performance of the different Cu-ZnO-ZrO2 catalysts. Reaction conditions: P = 3.0 MPa, H2:CO2 = 3 (v/v), T = 240 °C, GHSV = 2400 mL/(gcat·h).
Figure 10. Catalytic performance of the different Cu-ZnO-ZrO2 catalysts. Reaction conditions: P = 3.0 MPa, H2:CO2 = 3 (v/v), T = 240 °C, GHSV = 2400 mL/(gcat·h).
Catalysts 14 00390 g010
Figure 11. The relationship between the methanol yield and the total CO2 adsorption (a), and the SCu (b).
Figure 11. The relationship between the methanol yield and the total CO2 adsorption (a), and the SCu (b).
Catalysts 14 00390 g011
Figure 12. Schematic illustration of the effect of the solvothermal time for CO2 hydrogenation to methanol.
Figure 12. Schematic illustration of the effect of the solvothermal time for CO2 hydrogenation to methanol.
Catalysts 14 00390 g012
Table 1. Physicochemical properties of different Cu-ZnO-ZrO2 catalysts.
Table 1. Physicochemical properties of different Cu-ZnO-ZrO2 catalysts.
CatalystSBET
(m2/g)
Pore Volume
(cm3/g)
Pore Diameter
(nm)
dCuO1
(nm)
dCu1
(nm)
SCu2
(m2/g)
CZZ-120.70.05510.626.830.24.6
CZZ-332.70.0698.122.227.212.0
CZZ-640.50.0878.619.125.715.7
CZZ-1216.40.05413.327.729.14.5
1 Determined by XRD. 2 Determined by the N2O chemisorption method.
Table 2. The temperature and H2 consumption in the H2-TPR patterns of the various catalysts.
Table 2. The temperature and H2 consumption in the H2-TPR patterns of the various catalysts.
CatalystTemperature of Peaks (°C)H2 Consumption of Peaks (μmol/g)Degree of Reduction 1
TαTβαβTotal
CZZ-1211262138267628140.81
CZZ-3210-2802-28020.80
CZZ-6201-3306-33060.94
CZZ-12242-3378-33780.96
1 Determined by the ratio of the actual H2 consumption of the catalysts to the theoretical H2 consumption.
Table 3. The data of the CO2-TPD profiles of the different Cu-ZnO-ZrO2 catalysts.
Table 3. The data of the CO2-TPD profiles of the different Cu-ZnO-ZrO2 catalysts.
Catalystα Peakβ Peakγ Peak(Aα + Aβ + Aγ)/(a.u.)
Tα/(°C)Aα/(a.u.)Tβ/(°C)Aβ/(a.u.)Tγ/(°C)Aγ/(a.u.)
CZZ-1105471507238742161
CZZ-311411616520738946369
CZZ-610570151144380101315
CZZ-12111461606839042156
Table 4. The data of the H2-TPD profiles of the different Cu-ZnO-ZrO2 catalysts.
Table 4. The data of the H2-TPD profiles of the different Cu-ZnO-ZrO2 catalysts.
Catalystα Peakβ PeakAα + Aβ/(a.u.)
Tα/(°C)Aα/(a.u.)Tβ/(°C)Aβ/(a.u.)
CZZ-11372028580100
CZZ-31482532481106
CZZ-610630282172202
CZZ-121612532782107
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

Han, J.; Liang, Y.; Yu, J.; Wu, G.; Mao, D. Exploring the Effect of the Solvothermal Time on the Structural Properties and Catalytic Activity of Cu-ZnO-ZrO2 Catalysts Synthesized by the Solvothermal Method for CO2 Hydrogenation to Methanol. Catalysts 2024, 14, 390. https://0-doi-org.brum.beds.ac.uk/10.3390/catal14060390

AMA Style

Han J, Liang Y, Yu J, Wu G, Mao D. Exploring the Effect of the Solvothermal Time on the Structural Properties and Catalytic Activity of Cu-ZnO-ZrO2 Catalysts Synthesized by the Solvothermal Method for CO2 Hydrogenation to Methanol. Catalysts. 2024; 14(6):390. https://0-doi-org.brum.beds.ac.uk/10.3390/catal14060390

Chicago/Turabian Style

Han, Jian, Yannan Liang, Jun Yu, Guisheng Wu, and Dongsen Mao. 2024. "Exploring the Effect of the Solvothermal Time on the Structural Properties and Catalytic Activity of Cu-ZnO-ZrO2 Catalysts Synthesized by the Solvothermal Method for CO2 Hydrogenation to Methanol" Catalysts 14, no. 6: 390. https://0-doi-org.brum.beds.ac.uk/10.3390/catal14060390

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