Next Article in Journal
Enhanced Visible-Light-Driven Photocatalysis of Ag/Ag2O/ZnO Nanocomposite Heterostructures
Next Article in Special Issue
Silica-Decorated NiAl-Layered Double Oxide for Enhanced CO/CO2 Methanation Performance
Previous Article in Journal
A Review of the Advances and Challenges in Measuring the Thermal Conductivity of Nanofluids
Previous Article in Special Issue
CeO2-Promoted PtSn/SiO2 as a High-Performance Catalyst for the Oxidative Dehydrogenation of Propane with Carbon Dioxide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CO2 Activation and Hydrogenation on Cu-ZnO/Al2O3 Nanorod Catalysts: An In Situ FTIR Study

1
Department of Chemical Engineering, Guangdong Technion-Israel Institute of Technology (GTIIT), Shantou 515063, China
2
Schulich Faculty of Chemistry, Technion-Israel Institute of Technology (IIT), Haifa 32000, Israel
3
Guangdong Provincial Key Laboratory of Materials and Technologies for Energy Conversion (MATEC), Guangdong Technion-Israel Institute of Technology (GTIIT), Shantou 515063, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(15), 2527; https://doi.org/10.3390/nano12152527
Submission received: 30 June 2022 / Revised: 15 July 2022 / Accepted: 15 July 2022 / Published: 23 July 2022

Abstract

:
CuZnO/Al2O3 is the industrial catalyst used for methanol synthesis from syngas (CO + H2) and is also promising for the hydrogenation of CO2 to methanol. In this work, we synthesized Al2O3 nanorods (n-Al2O3) and impregnated them with the CuZnO component. The catalysts were evaluated for the hydrogenation of CO2 to methanol in a fixed-bed reactor. The support and the catalysts were characterized, including via in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS). The study of the CO2 adsorption, activation, and hydrogenation using in situ DRIFT spectroscopy revealed the different roles of the catalyst components. CO2 mainly adsorbed on the n-Al2O3 support, forming carbonate species. Cu was found to facilitate H2 dissociation and further reacted with the adsorbed carbonates on the n-Al2O3 support, transforming them to formate or additional intermediates. Like the n-Al2O3 support, the ZnO component contributed to improving the CO2 adsorption, facilitating the formation of more carbonate species on the catalyst surface and enhancing the efficiency of the CO2 activation and hydrogenation into methanol. The synergistic interaction between Cu and ZnO was found to be essential to increase the space–time yield (STY) of methanol but not to improve the selectivity. The 3% CuZnO/n-Al2O3 displayed improved catalytic performance compared to 3% Cu/n-Al2O3, reaching a CO2 conversion rate of 19.8% and methanol STY rate of 1.31 mmolgcat−1h−1 at 300 °C. This study provides fundamental and new insights into the distinctive roles of the different components of commercial methanol synthesis catalysts.

1. Introduction

In the last two decades, global warming and related extreme weather conditions have become topical issues, and the mitigation of carbon dioxide (CO2) release into the atmosphere is a top global priority. CO2 is a major anthropogenic greenhouse gas (GHG) released during fossil fuel exploitation; thus, controlling CO2 emissions is critical to maintaining the carbon-neutral state of the atmosphere [1,2,3]. The catalytic reduction of CO2 with renewable hydrogen (H2) to value-added chemicals and fuels such as hydrocarbons and alcohols represents a potential strategy to mitigate CO2 emissions into the atmosphere [1,4,5,6]. For example, the hydrogenation of CO2 to methanol is an important reaction in CO2 conversion, as methanol can be used as a feedstock chemical and fuel [7]. However, CO2 is a thermodynamically inert molecule that needs high reaction temperatures to activate the C=O bond. At high temperatures, the undesired reverse water–gas shift (RWGS) reaction is thermodynamically favored, which reduces the product selectivity. Therefore, developing a catalyst that can efficiently drive the process under mild conditions is imperative.
Different types of catalysts have been investigated for the thermocatalytic hydrogenation of CO2 to methanol, including supported metal [8,9] and metal oxide [10,11] catalysts. Cu-based catalysts have been reported to have the best activity for methanol production under industrially relevant conditions (5–10 MPa and 200–300 °C) [12,13]. The Cu-ZnO-Al2O3 catalyst prepared via the co-precipitation method is the industrial catalyst used for methanol production from CO + H2; thus, the Cu-ZnO-based catalysts are widely investigated for CO2 hydrogenation to methanol as well [14,15]. Various supports and promoters are explored for these catalysts, and different catalyst structures and active sites are reported [15,16,17]. The common supports or promoters for Cu-based catalysts include ZnO, ZnO-Al2O3, ZrO2, TiO2, In2O3, CeO2, and SiO2 [18,19]. At times the contribution of the support is complex. For example, Nitta et al. [18] observed significantly improved selectivity towards methanol formation at a low temperature upon introducing ZnO into Cu/ZrO2. However, at high temperatures, adding ZnO promoted the methanol decomposition into CO, leading to a reduced yield for methanol.
Factors impacting the overall catalytic performance of the Cu-based catalysts include the nature of the support material, the Cu loading, the dispersion, and the preparation method [12,20]. However, their effects on the activation of CO2 are still not clear. The specific roles of the different components within the industrial Cu-ZnO-Al2O3 catalyst for methanol synthesis from CO2 were assigned in the literature [12,18]. ZnO is generally believed to be a multitasked component, acting as a physical spacer for the Cu particles and facilitating H2 dissociation; Al2O3 plays the role of a structural promoter that increases Cu dispersion, whereas Cu drives the selectivity towards methanol [21,22]. However, further investigation is required to ascertain the individual roles of the catalyst components concerning CO2 activation, as well as to better understand the observed synergetic effects within the complete system. In addition to the composition, the nature of the catalyst support and characteristics such as the morphology (size and shape), were also found to influence the CO2 conversion efficiency and selectivity towards methanol formation [23,24,25,26]. An et al. [27] found that the catalytic activities of Cu/Zn/Al/Zr catalysts depended strongly on the morphology of the support, and the utilization of a fibrous shape improved the hydrogenation performance significantly. In addition, the proper selection of the catalyst support may provide additional advantages, such as reducing the amount of required active metal, inhibiting the sintering of the active metal, and prolonging the stability [22].
Diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) is a powerful technique for analyzing in situ substrate or adsorbate interactions, and has been previously utilized to characterize systems where the interactions of gases over the surface of a catalyst elucidated the catalytic surface chemistry [28,29,30,31].
In this work, we explored porous Al2O3 nanorods as the catalyst support in the CuZnO/Al2O3 catalysts for CO2 hydrogenation. The Al2O3 nanorods were synthesized via the steam-assisted solid wet–gel method and deployed as a support for CuZnO. The CuZnO/Al2O3 catalyst was prepared using the incipient wetness impregnation method and evaluated for the hydrogenation of CO2 to methanol. The catalysts supported on alumina nanorods exhibited better CO2 reduction to methanol under the examined conditions than that supported on commercial alumina. Through a DRIFTS study, we gained a new understanding of the specific roles of Al2O3, ZnO, and Cu in CO2 adsorption, activation, and hydrogenation reaction. This information will contribute to the design of catalysts with improved performance for industrial applications.

2. Experimental Section

2.1. Sample Preparation

(i)
Nanorods Al2O3 preparation.
Al2O3 nanorods were synthesized using the steam-assisted solid wet–gel method according to the literature [32]. Typically, 30 g Al(NO3)3·9H2O (>99.9%, SCR, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China) was dissolved in 50 mL of deionized water. Approximately 15% NH4OH aqueous solution (36%, SCR, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China) was added dropwise to the aluminum nitrate solution at room temperature under stirring to control the pH at 5.0. The resulting solid precipitate was recovered by filtration. The as-prepared solid cake-like wet gel was transferred to a glass beaker (25 mL) sitting in a Teflon vessel (200 mL), where 2 g of water was poured into the bottom of the vessel but was physically separated from the solid gel sample. The Teflon vessel was sealed and heated at 200 °C for 48 h. The obtained white solid material was washed with deionized water and recovered by centrifugation, dried at 60 °C for 24 h, and subsequently calcined in air at 600 °C for 5 h to obtain Al2O3 nanorods.
(ii)
Catalyst preparation.
The Al2O3-supported CuZnO catalysts were prepared via the incipient wetness impregnation method using Cu(NO3)2·6H2O and Zn(NO3)2·6H2O (>99.9%, SCR, Sinopharm Chemical Reagent Co., Ltd., Shanghai, China) (mass ratio of Cu: ZnO = 2:1) as the Cu and Zn precursors. The catalysts are labeled here as CuZnO/n-Al2O3 and CuZnO/c-Al2O3, where n-Al2O3 and c-Al2O3 are Al2O3 nanorods and commercial Al2O3 (γ-Al2O3, >99.9%, Shanghai Macklin Biochemical Co., Ltd., Shanghai, China), respectively. The loading of the active component is referred to as the total weight loading of Cu and ZnO (Cu+ZnO) in the catalysts. The resulting samples were dried overnight at 60 °C and subsequently calcined in air at 400 °C for 4 h. Catalysts with Cu+ZnO loading of 3–20% and 3% Cu were studied.

2.2. Characterization

The crystallinity of the samples was characterized via X-ray diffraction (XRD) experiments performed on a Bruker D8 Advance X-ray diffractometer (Cu Kα) (Bruker Corporation Inc., Billerica, MA, USA). The diffraction patterns were acquired at a scanning step of 0.01° and a scanning speed of 10°/min. Transmission electron microscopy (TEM) and scanning electron microscopy (SEM) images of the samples were recorded on a Talos F200X transmission electron microscope (Thermo Fisher Scientific Inc., Waltham, MA, USA) operated at 300 kV and a scanning electron microscope (GeminiSEM 450, Carl ZEISS, Oberkochen, Germany), respectively.
In situ DRIFTS measurements were performed on a Nicolet iS50 FTIR instrument (Thermo Fisher Scientific Inc., Waltham, MA, USA). The scans were recorded from 4000 to 600 cm−1. Here, 50 mg catalyst powder was placed in a high-pressure (0–10 MPa) DRIFT cell (HVC-DRP-5, Harrick Scientific Products Inc., NY, USA) equipped with ZnSe windows and reduced at 300 °C in a pure H2 stream (30 mL/min) for 2 h, after which the sample was flushed with N2 (40 mL/min) for 1 h and cooled to room temperature. The background subtractions were executed over different samples for testing in 40 mL/min N2 under 2 MPa. Then, the reactant gas mixture (H2 + CO2, 3:1 ratio) at a flow rate of 40 mL/min was introduced into the sample cell and the spectra were collected at different temperatures, increasing by 3 °C/min at an interval of 19 s. For the CO2 activation study, the sample cell was cooled to 250 °C. The background subtractions were similarly performed under 2 MPa. The CO2 was switched into the reaction chamber and the spectra were recorded for up to 120 min, after which the inlet gas was changed to H2 and the process lasted for another 1 h 40 min.

2.3. Catalytic Performance Evaluation

The CO2 hydrogenation to methanol was performed in a high-pressure fixed-bed flow reactor. Here, 0.1 g catalyst was fixed in a quartz tube using quartz wool and then packed into a stainless steel tubular reactor. Prior to the catalytic measurements, the catalyst was reduced in a gas stream (10% H2/N2) at 300 °C for 2 h with a flow rate of 10 mL/min under atmospheric pressure. Then, the temperature was decreased to 200 °C and the reducing gas was replaced by reaction gas (CO2 + H2, 1:3 ratio). The reaction was X(CO2) gas hourly space velocity (GHSV) of 7800 mL g−1 h−1. The products flowing out from the reactor passed through a tube connected to a temperature control box to maintain the temperature at 120 °C and were then analyzed using an online gas chromatographer (Shimadzu GC-2014C) equipped with a TCD and an FID. The data obtained from the GC measurements were used to calculate the CO2 conversion (X(CO2)), CH3OH selectivity (Si), and methanol space–time yield (STYCH3OH) using Equations (1)–(3).
X C O 2 = n c o 2 , i n n c o 2 , o u t n c o 2 , i n × 100 %
S i = n p r o d u c t s ,   i n c o 2 , i n n c o 2 , o u t × 100 %
S T Y C H 3 O H = G H S V × 0.25 22.4 × X C O 2 S i 10000

3. Results and Discussion

3.1. Catalyst Structure, Morphology and Textural Properties

The XRD patterns of the n-Al2O3 support with various levels of CuZnO loadings are presented in Figure 1. The pattern of the catalyst support shows diffractions at the 2θ values of around 37.3, 39.4, 46.0, and 67.0°, characteristic of the γ-Al2O3 (JCPDS 04-0858). The patterns of calcined catalyst samples at low loadings (3% and 6% CuZnO) are quite similar to that of the n-Al2O3 support. No XRD peaks of CuO or ZnO can be observed, suggesting the high dispersion of CuO and ZnO, or possibly because the amount yielded in the XRD signal was close to the detection limit. In contrast to the observations for low loadings, prominent CuO XRD peaks can be observed for 10% and 20% CuZnO loadings at 35.7°, 38.6°, and 48.7° (JCPDS 80-1268).
Figure 2a shows the morphology of the n-Al2O3 support, which can be described as γ-Al2O3 nanorods with lengths and widths of around 250–300 nm and 15–25 nm, respectively, in comparison with the spherical nano-sized particles of c-Al2O3 (figure not shown). The observed morphology is similar to that of the alumina nanorods reported in the literature [32,33]. Figure 2b shows the morphology of the n-Al2O3 impregnated with 3% CuZnO, and the corresponding EDX result presented in Table 1 shows a Cu/ZnO ratio of 2.25, which is reasonably close to the theoretical value of 2. The TEM micrographs of the n-Al2O3 and 3% CuZnO/n-Al2O3 catalysts are shown in Figure 3. After impregnating the n-Al2O3 support with 3% CuZnO, the morphology and structure of n-Al2O3 remained unchanged. Both materials consist of nanorods with pores of about 3–10 nm in diameter. Due to their low loadings and low contrast, it is difficult to visualize the CuO and ZnO particles on the surface or inside the pores of the alumina nanorods with traditional wide-field TEM [34]. In Figure 4a, the linear isotherms of Cu/n-Al2O3, CuZnO/n-Al2O3, ZnO/n-Al2O3, and CuZnO/c-Al2O3 with corresponding BET surface areas of 94.4, 89.4, 67.6, and 99.3 m2/g, respectively, are shown. Except for CuZnO/c-Al2O3, other samples exhibit type IV isotherms with H1-type hysteresis loops, suggesting a typical mesoporous structure related to the n-Al2O3 support [35]. CuZnO/c-Al2O3 exhibits a type II isotherm with H4-type hysteresis loops, indicating an otherwise smaller mesoporous structure. According to the PSD curves (Figure 4b), CuZnO/c-Al2O3 possesses pores exhibiting a broad range but with an average pore width of 68 Å. Despite the n-Al2O3-supported samples exhibiting a wider pore size distribution, more of the pores are concentrated in the low-diameter region up to 150 Å. Cu/n-Al2O3 and CuZnO/n-Al2O3 have a nearly similar total pore volume of approximately 0.5 cm3/g, which is greater than those of ZnO/n-Al2O3 and CuZnO/c-Al2O3, 0.35 and 0.21 cm3/g, respectively.

3.2. Catalytic Performance Evaluation

3.2.1. CO2 Hydrogenation to Methanol

The catalytic performances, including the CO2 conversion efficiency, selectivity towards methanol, and space–time yield for methanol formation, are presented in Figure 5. The influence of the temperature was examined in the range of 200 °C to 300 °C for the catalysts with CuZnO loadings from 3 to 20%, and the results are presented in Figure 5a. As expected, the CO2 conversion increased with the temperature for all of the examined catalysts because the thermodynamically stable CO2 molecule can be more efficiently broken down at high temperatures. However, at high reaction temperatures, the selectivity towards CH3OH formation presented in Figure 5b decreased, given the competitive advantage of the RWGS reaction, resulting in CO formation [36]. The increasing CuZnO loading increased the CH3OH selectivity from 58.5% for 3% CuZnO to 79.5% for 20% CuZnO at 200 °C, yet the improved selectivity did not essentially originate from increased CO2 conversion efficiency. The catalysts with 6–20% CuZnO had relatively similar CO2 conversion levels, higher than that of the catalyst containing 3% CuZnO by 12.6% at 300 °C. In addition, the methanol selectivity and STY of the former were higher than that of the latter at all temperatures (Figure 5b,c).
Since our aim was not to pursue a high efficiency for CO2 conversion to methanol but rather to explore and understand the influence of the catalyst structure on its catalytic performance, as well as the fact that the catalyst with 3% CuZnO showed a more homogeneous distribution of CuZnO on the catalyst support than the catalyst with higher loadings, we focused on the catalytic performance of the 3% CuZnO catalyst. As shown in Figure 5d–f, different samples containing 3% Cu, 3% ZnO, or 3% CuZnO were compared to better ascertain the role of each component and the synergistic effect. In addition, the catalytic performance of 3% CuZnO/n-Al2O3 was compared with that of 3% CuZnO/c-Al2O3 (commercial Al2O3). As shown in Figure 5a–f, improved catalytic performance was observed, with better CO2 conversion (19.8%) and STYCH3OH (1.31 mmol·gcat−1h−1) at 300 °C while utilizing the 3% CuZnO/n-Al2O3 sample. This combination outperformed those of the separate components, as well as the sample supported on the commercial Al2O3. However, the methanol selectivity on 3% CuZnO/n-Al2O3 was lower than those of the Cu/n-Al2O3 and CuZnO/c-Al2O3 at temperatures above 220 °C (Figure 5e). The 3% ZnO/n-Al2O3 exhibited the lowest catalytic activity under all studied temperatures. This observation was not unexpected because Cu is the main component that contributes to methanol selectivity. The pure n-Al2O3 support was found to have negligible activity for CH3OH production (data not shown). Overall, CuZnO exhibited better reactivity than either Cu or ZnO. While Cu is the methanol-selective component, ZnO and Al2O3 play other roles in the catalytic process [12].

3.2.2. DRIFTS Study on the CO2 Activation

The CO2 activation was investigated via in situ DRIFT spectroscopy [37,38]. First, CO2 was adsorbed on the various fresh catalysts (at 250 °C, 2 MPa) that were reduced in H2 at 300 °C and the spectra were recorded. After sufficient adsorption of CO2 and attaining a stable state for up to 120 min, the inlet gas was changed to H2. The spectra were again recorded for up to 220 min under the same conditions. The different catalyst components were studied during the experiments, and their specific roles in CO2 activation and hydrogenation were elucidated.
(i)
CO2 activation on n-Al2O3
The pure n-Al2O3 in the CO2 stream evolves bands at 1661, 1611, 1585, 1548, 1447, 1393, 1327, and 1228 cm−1 attributed to carbonate species, indicating adsorption of CO2 on Al2O3 (Figure 6a). Specifically, the bands at 1661, 1585, and 1447 cm−1 are those of bicarbonates (HCO3) species [39]. Other bands at 1548 and 1327 cm−1 can be assigned to polydentate carbonate, while the band at 1393 cm−1 belongs to monodentate carbonate. The 1228 cm−1 band can be attributed to the δ(OH) of bicarbonate species. This band decreases or even disappears with time. On the C-H stretching region in Figure 6b, bands can be observed at 3002, 2929, 2899-2896, 2875, 2769, and 2741 cm−1. According to the literature, the bands at 3002 and 2769 cm−1 were assigned to bidentate formate species on Al2O3-supported catalyst [40,41,42]; however, these bands were formed on pure alumina nanorods in this study. Here, we question the formation of formate species, since no H2 gas was passed through the sample during the first 120 min, except that certain reactions exist with OH groups on the Al2O3 surface [37]. Upon passing H2 gas through the sample, the band frequencies in the spectra did not change significantly, indicating that no substantial reaction occurred with the absorbed carbon species and pointing to the fact that Al2O3 cannot or can poorly activate H2. In addition, the presence of H2 may promote CO2 desorption from the Al2O3 surface, since there was no strong reaction between them (CO2 and Al2O3), as indicated by the disappearance of the band at 1228 cm−1. This result indicates that nanorod Al2O3 can be used as a support for the CO2 conversion catalyst due to its ability to adsorb CO2 in its activated forms.
(ii)
CO2 activation on Cu/Al2O3 and ZnO/n-Al2O3
The formation of carbonate species on the surface of Cu/Al2O3 was obvious at 120 min, as evidenced by the bands at 1644, 1442, and 1228 cm−1 [39,42] (Figure 6c). Interestingly, these carbonate species disappeared and formate species formed after switching to H2. This indicates that Cu can activate H2 and promote its reaction with adsorbed CO2 species [43]. The formate species showed bands at 1662, 1585, 1395, and 1329 cm−1. These bands were attributed to bidentate formate species on both Cu and Al2O3 and could be confirmed by the new bands at 2999, 2952, and 2742 cm−1 in the C-H stretching spectra (Figure 6d) [41,44]. Wang et al. [37] assigned the band at around 2999 cm−1 to the C-H bonds in unsaturated CHx on CuNi catalysts, and its presence facilitated the deep hydrogenation of CO2 to form methanol. For ZnO/n-Al2O3 ((Figure 6e,f), after switching the feed gas from CO2 to H2, there was no strong indication of H2 activation; however, the appearance of bands of higher intensities and at nearly similar frequencies to that of pure n-Al2O3 is an indication that ZnO promoted the CO2 adsorption on the catalyst surface. The prominence of bands at all frequencies, which originated when only CO2 was adsorbed and maintained in the presence of H2, prove that ZnO activates CO2 mainly as carbonates. However, some bands exhibited slight shifts in their positions, which indicates some sort of interaction and formation of formate species. In a previous study, the bands at 1385 and 1328 cm−1 were assigned to bidentate formate on ZnO and Al2O3, whereas that at 1613 cm−1 was attributed to Vas(OCO) of a bidentate carbonate species on ZnO [45]. The formation of formate species on ZnO can be confirmed by the evolution of 2946 and 2841 cm−1 bands in the C-H stretch spectra.
(iii)
CO2 activation on CuZnO/Al2O3.
In the case of CuZnO/n-Al2O3, in the CO2 gas stream, similar spectral features to Cu/n-Al2O3 were observed. The peaks at 1643 and 1434 cm−1 were those of Vs(OCO) of bicarbonate species [42]. The band at 1228 cm−1 was assigned to carboxylate (CO2δ−) species in previous studies [38,46], but we related the band to the δ(OH) of bicarbonate species, as earlier mentioned. However, the peak intensities were found to be higher than those obtained from Cu/n-Al2O3. This was attributed to the contribution of ZnO, which promotes CO2 adsorption. After switching the gas to H2, most of the formed carbonate species converted into the formate species by reacting with H-atoms. This was evident from analyzing the bands at 1616, 1585, 1388, and 1326 cm−1, which indicated formate species on Al2O3, ZnO, and Cu [41,45]. The increased band intensities at 2897, 2921, and 2995 cm−1 indicated that the evolution of bidentate formate on the catalyst increased with time. It follows that CuZnO/n-Al2O3 exhibited the best CO2 activation, good H2 activation, and catalytic activity for converting carbonate into formate, which is an important intermediate for methanol production. The increase in peak intensity means that the CO2 adsorption intensified with time and reached a maximum (steady-state) at 120 min.

3.2.3. DRIFTS Study of Methanol Formation from CO2 Hydrogenation

The key surface species and intermediates involved in forming methanol over the catalysts are shown in Figure 7. The in situ DRIFT spectra collected at different temperatures on n-Al2O3, ZnO/Al2O3, Cu/n-Al2O3, and CuZnO/n-Al2O3 dosed with the reaction gas mixture indicated that the hydrogenation of CO2 was initiated at a temperature above 50 °C. The initiation temperature decreases when the catalyst contains an active phase (Figure 7d,f,h). Except for the band around 1228 cm−1, which is associated with δ(OH) of bicarbonate species or weakly adsorbed carbonate species (CO32−) that are desorbed at high temperatures, there is an increase in the intensities of the bands of hydrogenated species as the temperature rises. This means that the formation of key surface species increases with the temperature. The peaks at 1644 and 1416 cm−1 are attributed to bicarbonate species formed when the reaction gas mixture passed through n-Al2O3 at 50 °C (Figure 7a). At higher temperatures, formate species formation is evident, typically indicated by the bands at 1612, 1583, 1391, and 1329 cm−1. The formate species are also confirmed with the C-H stretching bands at 2997 and 2902 cm−1 in Figure 7b, which are identical to those observed when formic acid is adsorbed and evacuated at 473 K on Al2O3 [41]. In Figure 7e,f, the bands at 1661, 1608, 1582, 1389, and 1325 cm−1 are attributed to formate species on Cu and Al2O3 [40]. Specifically, the band at 1325 cm−1 is for Vs(OCO) of bidentate formate species on Cu [40,41]. Other bands also evolved at 2931 and 2740 cm−1, related to v(C-H)+vas(OCO) and v(C-H)+vs(OCO) of bidentate formate on Al2O3 [40]. On CuZnO/n-Al2O3 (Figure 7g,h), the reaction gas is adsorbed as carbonate species up to 100 °C, as indicated by the adsorption bands around 1526 and 1415 cm−1 for polydentate carbonate and bicarbonate species, respectively [40]. The formation of formate species can be related to the bands at 1666 and 1628 cm−1. The bidentate formate species on ZnO and Al2O3 showed bands of higher intensities at 1575, 1388, 1314, 2888, and 2740 cm−1 [41,45]. Meanwhile, the 2921 cm−1 band can be assigned to the v(C-H) and δ(C-H)+vas(OCO) of bidentate formate on Cu [41,44]. The peaks at 2948 and 2843 in the CuZnO/n-Al2O3 spectra and at 2931 and 2846 in the Cu/n-Al2O3 spectra indicate the formation of methoxy species [30]. At 150 °C or higher temperatures, the band intensities increased compared with those of Cu/n-Al2O3. For these catalysts, the carbonate conversion to formate species is seen on both CuZnO and Cu, but the formation of formate species begins at a low temperature (100 °C) exclusively on Cu/n-Al2O3. The key intermediate species formed at different temperatures are summarized in Table 2. The introduction of ZnO enhances the CO2 adsorption and facilitates more carbonate species formation. This, in turn, boosts the formate production through hydrogenation via dissociated H atoms, which spillover to the catalyst surface. The formation of formate at a lower temperature on the CuZnO/n-Al2O3 catalyst might be ascribed to a more effective H2 dissociation in the presence of ZnO [40]. Generally, two reaction pathways for the hydrogenation of the CO2 to methanol have been identified: RWGS + CO-hydro, along with formate pathways [47]. However, many studies on Cu-based catalysts conform to the formate pathway [30,38]. In this study, CO2 is activated as carbonate species on the catalyst surface under the examined reaction conditions. As the temperature increases, H2 is better activated and spills over to the catalyst surface, where it reacts with the carbonate and forms formate and methoxy species. Further hydrogenation of the surface species produces methanol. This mechanism is, thus, in good agreement with literature reports for other Cu-based systems.

3.3. Roles of Al2O3, ZnO, and Cu in CO2 Hydrogenation to Methanol

Understanding the structure–function relationship of a composite catalyst is important to develop more efficient catalysts for industrial applications. The individual roles of Al2O3, ZnO, and Cu and their synergistic effects in the CuZnO/Al2O3 catalyst used for the CO2-to-methanol synthesis were investigated using DRIFT spectroscopy. Al2O3 is a large surface area support material with an appropriate surface for dispersing and stabilizing the active phase and slowing the sintering during the reaction at high temperatures [40]. Herein, we found Al2O3 to be a good material for CO2 adsorption and activation to produce bicarbonate, carbonate, and formate species; the bicarbonate and carbonate species are converted into formate and methoxy in the presence of metal species that can activate H2. The 3% Cu/n-Al2O3 and 3% CuZnO/n-Al2O3 catalysts possess nearly similar BET surface areas of 94.4 and 89.4 m2/g, respectively. This shows that both active phases had the same impact on the surface area; however, the catalysts exhibit slight differences in their CO2 adsorption and catalytic activities. Methoxy and increasing amounts of formate species are produced for both catalysts. The conversion of carbonates to formate is evident, as H-atoms are available for the reaction after activation on Cu particles following the spillover effect. According to the catalytic performance and in situ DRIFT results, CuZnO/n-Al2O3 exhibited the best performance due to the synergistic effect resulting from the interaction of Cu and ZnO. Apparently, as ascertained from the in situ DRIFT spectra, the presence of ZnO with Cu promotes CO2 adsorption and formate formation. The enhanced formate formation is thought to be due to additional H2 dissociation activity. We, thus, suggest that ZnO also promotes H2 activation. It was previously reported that a defective ZnOx overlayer formed in a Cu/ZnO catalyst could dissociate H2, even at room temperature [40]. From the above discussion, it will be reasonable to conclude from our study that the different components of the industrial catalyst for methanol synthesis from CO2 play different roles but with a cooperative effect in methanol formation, as shown in Scheme 1.

4. Conclusions

Cu-based catalysts supported on Al2O3 nanorods for CO2 activation and subsequent hydrogenation to methanol are discussed in this study. Compared with the commercial Al2O3 support, catalysts using Al2O3 nanorods as the support show higher efficiency for CO2 conversion between 200 and 300 °C, but the methanol selectivity becomes lower above 220 °C. There is improved CO2 adsorption on the Al2O3 nanorods. CO2 is adsorbed and activated on the catalyst surface as carbonate species, which upon subsequent H2 dissociation on Cu hydrogenates to formate species by reacting with the H atom. ZnO promotes CO2 adsorption and probably facilitates H2 dissociation on Cu, also leading to the more efficient hydrogenation of activated CO2 species on CuZnO/Al2O3. This study will help to tailor the properties and composition of the Cu-based catalyst system to achieve higher efficiency for methanol synthesis from CO2.

Author Contributions

Conceptualization, Z.Z.; Investigation, L.W. and C.Z.; Supervision, L.A. and Z.Z.; Writing—original draft, L.W.; Writing—editing and review, U.J.E., Z.Z. and L.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Guangdong Provincial Key Laboratory of Materials and Technologies for Energy Conversion (MATEC), by the GTIIT Project KD2000079, Guangdong Technion—Israel Institute of Technology, the 2020 Li Ka Shing Foundation Cross-Disciplinary Research Grant (2020LKSFG09A), and the Guangdong Province Key discipline fund (GTIIT) in 2022.

Acknowledgments

The authors acknowledge the financial support by Guangdong Provincial Key Laboratory of Materials and Technologies for Energy Conversion (MATEC), GTIIT; the Li Ka Shing Foundation for Cross-Disciplinary Research, and the Guangdong Province.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. 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]
  2. Wang, Q.; Luo, J.; Zhong, Z.; Borgna, A. CO2 capture by solid adsorbents and their applications: Current status and new trends. Energy Environ. Sci. 2011, 4, 42–55. [Google Scholar] [CrossRef]
  3. Wang, J.; Huang, L.; Yang, R.; Zhang, Z.; Wu, J.; Gao, Y.; Wang, Q.; O’Hare, D.; Zhong, Z. Recent advances in solid sorbents for CO2 capture and new development trends. Energy Environ. Sci. 2014, 7, 3478–3518. [Google Scholar] [CrossRef]
  4. Aresta, M.; Dibenedetto, A.; Angelini, A. Catalysis for the Valorization of Exhaust Carbon: From CO2 to Chemicals, Materials, and Fuels. Technological Use of CO2. Chem. Rev. 2014, 114, 1709–1742. [Google Scholar] [CrossRef]
  5. Porosoff, M.D.; Yang, X.; Boscoboinik, J.A.; Chen, J.G. Molybdenum Carbide as Alternative Catalysts to Precious Metals for Highly Selective Reduction of CO2 to CO. Angew. Chem. Int. Ed. 2014, 53, 6705–6709. [Google Scholar] [CrossRef]
  6. Studt, F.; Sharafutdinov, I.; Abild-Pedersen, F.; Elkjær, C.F.; Hummelshøj, J.S.; Dahl, S.; Chorkendorff, I.; Nørskov, J.K. Discovery of a Ni-Ga catalyst for carbon dioxide reduction to methanol. Nat. Chem. 2014, 6, 320–324. [Google Scholar] [CrossRef]
  7. Sehested, J. Industrial and scientific directions of methanol catalyst development. J. Catal. 2019, 371, 368–375. [Google Scholar] [CrossRef]
  8. De, S.; Dokania, A.; Ramirez, A.; Gascon, J. Advances in the Design of Heterogeneous Catalysts and Thermocatalytic Processes for CO2 Utilization. ACS Catal. 2020, 10, 14147–14185. [Google Scholar] [CrossRef]
  9. Kattel, S.; Liu, P.; Chen, J.G. Tuning selectivity of CO2 hydrogenation reactions at the metal/oxide interface. J. Am. Chem. Soc. 2017, 139, 9739–9754. [Google Scholar] [CrossRef]
  10. Wang, J.; Zhang, G.; Zhu, J.; Zhang, X.; Ding, F.; Zhang, A.; Guo, X.; Song, C. CO2 hydrogenation to methanol over In2O3-based catalysts: From mechanism to catalyst development. ACS Catal. 2021, 11, 1406–1423. [Google Scholar] [CrossRef]
  11. Lin, L.; Wang, G.; Zhao, F. CO2 Hydrogenation to Methanol on ZnO/ZrO2 Catalysts: Effects of Zirconia Phase. ChemistrySelect 2021, 6, 2119–2125. [Google Scholar] [CrossRef]
  12. Etim, U.; Song, Y.; Zhong, Z. Improving the Cu/ZnO-Based Catalysts for Carbon Dioxide Hydrogenation to Methanol, and the Use of Methanol as a Renewable Energy Storage Media. Front. Energy Res. 2020, 8. [Google Scholar] [CrossRef]
  13. Etim, U.J.; Semiat, R.; Zhong, Z. CO2 Valorization Reactions over Cu-Based Catalysts: Characterization and the Nature of Active Sites. Am. J. Chem. Eng. 2021, 9, 53–78. [Google Scholar] [CrossRef]
  14. 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]
  15. 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]
  16. Huš, M.; Kopač, D.; Štefančič, N.S.; Jurković, D.L.; Dasireddy, V.D.; Likozar, B. Unravelling the mechanisms of CO2 hydrogenation to methanol on Cu-based catalysts using first-principles multiscale modelling and experiments. Catal. Sci. Technol. 2017, 7, 5900–5913. [Google Scholar] [CrossRef] [Green Version]
  17. 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] [Green Version]
  18. Nitta, Y.; Suwata, O.; Ikeda, Y.; Okamoto, Y.; Imanaka, T. Copper-zirconia catalysts for methanol synthesis from carbon dioxide: Effect of ZnO addition to Cu-ZrO2 catalysts. Catal. Lett. 1994, 26, 345–354. [Google Scholar] [CrossRef]
  19. 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]
  20. Dasireddy, V.D.; Likozar, B. The role of copper oxidation state in Cu/ZnO/Al2O3 catalysts in CO2 hydrogenation and methanol productivity. Renew. Energ. 2019, 140, 452–460. [Google Scholar] [CrossRef]
  21. 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]
  22. Liang, B.; Ma, J.; Su, X.; Yang, C.; Duan, H.; Zhou, H.; Deng, S.; Li, L.; Huang, Y. Investigation on Deactivation of Cu/ZnO/Al2O3 Catalyst for CO2 Hydrogenation to Methanol. Ind. Eng. Chem. Res. 2019, 58, 9030–9037. [Google Scholar] [CrossRef] [Green Version]
  23. Narayanan, R.; El-Sayed, M.A. Catalysis with Transition Metal Nanoparticles in Colloidal Solution: Nanoparticle Shape Dependence and Stability; ACS Publications: Washington, DC, USA, 2005; pp. 12663–12676. [Google Scholar]
  24. Sakpal, T.; Lefferts, L. Structure-dependent activity of CeO2 supported Ru catalysts for CO2 methanation. J. Catal. 2018, 367, 171–180. [Google Scholar] [CrossRef]
  25. Lei, H.; Nie, R.; Wu, G.; Hou, Z. Hydrogenation of CO2 to CH3OH over Cu/ZnO catalysts with different ZnO morphology. Fuel 2015, 154, 161–166. [Google Scholar] [CrossRef]
  26. Liao, F.; Huang, Y.; Ge, J.; Zheng, W.; Tedsree, K.; Collier, P.; Hong, X.; Tsang, S.C. Morphology—Dependent Interactions of ZnO with Cu Nanoparticles at the Materials’ Interface in Selective Hydrogenation of CO2 to CH3OH. Angew. Chem. Int. Ed. 2011, 50, 2162–2165. [Google Scholar] [CrossRef]
  27. An, X.; Li, J.; Zuo, Y.; Zhang, Q.; Wang, D.; Wang, J. A Cu/Zn/Al/Zr fibrous catalyst that is an improved CO2 hydrogenation to methanol catalyst. Catal. Lett. 2007, 118, 264–269. [Google Scholar] [CrossRef]
  28. Huttunen, P.K.; Labadini, D.; Hafiz, S.S.; Gokalp, S.; Wolff, E.P.; Martell, S.M.; Foster, M. DRIFTS investigation of methanol oxidation on CeO2 nanoparticles. Appl. Surf. Sci. 2021, 554, 149518. [Google Scholar] [CrossRef]
  29. Reddy, K.P.; Choi, H.; Kim, D.; Choi, M.; Ryoo, R.; Park, J.Y. The facet effect of ceria nanoparticles on platinum dispersion and catalytic activity of methanol partial oxidation. Chem Commun 2021, 57, 7382–7385. [Google Scholar] [CrossRef]
  30. Chen, K.; Fang, H.; Wu, S.; Liu, X.; Zheng, J.; Zhou, S.; Duan, X.; Zhuang, Y.; Tsang, S.C.E.; Yuan, Y. CO2 hydrogenation to methanol over Cu catalysts supported on La-modified SBA-15: The crucial role of Cu–LaOx interfaces. Appl. Catal. B Environ. 2019, 251, 119–129. [Google Scholar] [CrossRef]
  31. Zhu, J.; Su, Y.; Chai, J.; Muravev, V.; Kosinov, N.; Hensen, E.J. Mechanism and nature of active sites for methanol synthesis from CO/CO2 on Cu/CeO2. ACS Catal. 2020, 10, 11532–11544. [Google Scholar] [CrossRef]
  32. Shen, S.; Chen, Q.; Chow, P.; Tan, G.; Zeng, X.; Wang, Z.; Tan, R.B. Steam-assisted solid wet-gel synthesis of high-quality nanorods of boehmite and alumina. J. Phy. Chem. C 2007, 111, 700–707. [Google Scholar] [CrossRef]
  33. Bai, P.; Su, F.; Wu, P.; Wang, L.; Lee, F.Y.; Lv, L.; Yan, Z.-f.; Zhao, X.S. Copolymer-Controlled Homogeneous Precipitation for the Synthesis of Porous Microfibers of Alumina. Langmuir 2007, 23, 4599–4605. [Google Scholar] [CrossRef]
  34. Martínez, A.; Prieto, G.; Rollán, J. Nanofibrous γ-Al2O3 as support for Co-based Fischer–Tropsch catalysts: Pondering the relevance of diffusional and dispersion effects on catalytic performance. J. Catal. 2009, 263, 292–305. [Google Scholar] [CrossRef]
  35. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef] [Green Version]
  36. 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]
  37. Wang, L.-X.; Guan, E.; Wang, Z.; Wang, L.; Gong, Z.; Cui, Y.; Yang, Z.; Wang, C.; Zhang, J.; Meng, X. Dispersed nickel boosts catalysis by copper in CO2 hydrogenation. ACS Catal. 2020, 10, 9261–9270. [Google Scholar] [CrossRef]
  38. Kattel, S.; Yan, B.; Yang, Y.; Chen, J.G.; Liu, P. Optimizing binding energies of key intermediates for CO2 hydrogenation to methanol over oxide-supported copper. J. Am. Chem. Soc. 2016, 138, 12440–12450. [Google Scholar] [CrossRef]
  39. 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]
  40. Hu, J.; Li, Y.; Zhen, Y.; Chen, M.; Wan, H. In situ FTIR and ex situ XPS/HS-LEIS study of supported Cu/Al2O3 and Cu/ZnO catalysts for CO2 hydrogenation. Chin. J. Catal. 2021, 42, 367–375. [Google Scholar] [CrossRef]
  41. Bando, K.K.; Sayama, K.; Kusama, H.; Okabe, K.; Arakawa, H. In-situ FT-IR study on CO2 hydrogenation over Cu catalysts supported on SiO2, Al2O, and TiO2. Appl. Catal. A Gen. 1997, 165, 391–409. [Google Scholar] [CrossRef]
  42. Baltrusaitis, J.; Jensen, J.H.; Grassian, V.H. FTIR spectroscopy combined with isotope labeling and quantum chemical calculations to investigate adsorbed bicarbonate formation following reaction of carbon dioxide with surface hydroxyl groups on Fe2O3 and Al2O3. J. Phys. Chem. B 2006, 110, 12005–12016. [Google Scholar] [CrossRef] [PubMed]
  43. Sakong, S.; Groß, A. Dissociative adsorption of hydrogen on strained Cu surfaces. Surf. Sci. 2003, 525, 107–118. [Google Scholar] [CrossRef]
  44. Fujita, S.-I.; Usui, M.; Ohara, E.; Takezawa, N. Methanol synthesis from carbon dioxide at atmospheric pressure over Cu/ZnO catalyst. Role of methoxide species formed on ZnO support. Catal. Lett. 1992, 13, 349–358. [Google Scholar] [CrossRef]
  45. Bailey, S.; Froment, G.; Snoeck, J.-W.; Waugh, K. A DRIFTS study of the morphology and surface adsorbate composition of an operating methanol synthesis catalyst. Catal. Lett. 1994, 30, 99–111. [Google Scholar] [CrossRef]
  46. Kim, Y.; Trung, T.S.B.; Yang, S.; Kim, S.; Lee, H. Mechanism of the surface hydrogen induced conversion of CO2 to methanol at Cu (111) step sites. ACS Catal. 2016, 6, 1037–1044. [Google Scholar] [CrossRef]
  47. Liu, C.; Liu, P. Mechanistic study of methanol synthesis from CO2 and H2 on a modified model Mo6S8 cluster. ACS Catal. 2015, 5, 1004–1012. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of n-Al2O3 and catalyst samples.
Figure 1. XRD patterns of n-Al2O3 and catalyst samples.
Nanomaterials 12 02527 g001
Figure 2. SEM images of (a) n-Al2O3 and (b) 3% CuZnO/n-Al2O3.
Figure 2. SEM images of (a) n-Al2O3 and (b) 3% CuZnO/n-Al2O3.
Nanomaterials 12 02527 g002
Figure 3. TEM images of (a,b) n-Al2O3 and (c,d) 3% CuZnO/n-Al2O3.
Figure 3. TEM images of (a,b) n-Al2O3 and (c,d) 3% CuZnO/n-Al2O3.
Nanomaterials 12 02527 g003
Figure 4. (a) Nitrogen adsorption–desorption isotherms and (b) pore size distribution curves of sample catalysts.
Figure 4. (a) Nitrogen adsorption–desorption isotherms and (b) pore size distribution curves of sample catalysts.
Nanomaterials 12 02527 g004
Figure 5. CO2 hydrogenation performances of catalyst samples. (a) CO2 conversion, (b) CH3OH selectivity, and (c) STYCH3OH of CuZnO with various loadings, supported on n-Al2O3. (d) CO2 conversion, (e) CH3OH selectivity, and (f) STYCH3OH of 3% Cu, 3% ZnO, and 3% CuZnO on n-Al2O3, and 3% CuZnO on c-Al2O3. Reaction conditions: GHSV = 7800 mL g−1 h−1, gas flow rate = 13 mL min−1, CO2:H2 = 1:3, P = 3.0 MPa.
Figure 5. CO2 hydrogenation performances of catalyst samples. (a) CO2 conversion, (b) CH3OH selectivity, and (c) STYCH3OH of CuZnO with various loadings, supported on n-Al2O3. (d) CO2 conversion, (e) CH3OH selectivity, and (f) STYCH3OH of 3% Cu, 3% ZnO, and 3% CuZnO on n-Al2O3, and 3% CuZnO on c-Al2O3. Reaction conditions: GHSV = 7800 mL g−1 h−1, gas flow rate = 13 mL min−1, CO2:H2 = 1:3, P = 3.0 MPa.
Nanomaterials 12 02527 g005
Figure 6. In situ DRIFT spectra at different times for the fresh catalysts (reduced in H2 at 300 °C, purged in N2, and then switched to CO2 gas): (a,b) n-Al2O3, (c,d) 3% Cu/n-Al2O3, (e,f) 3% ZnO/n-Al2O3, and (g,h) 3% CuZnO/n-Al2O3 at 250 °C under 2 MPa for both pure CO2 gas (0–120 min) and after switching the feed gas to H2 (120–220 min).
Figure 6. In situ DRIFT spectra at different times for the fresh catalysts (reduced in H2 at 300 °C, purged in N2, and then switched to CO2 gas): (a,b) n-Al2O3, (c,d) 3% Cu/n-Al2O3, (e,f) 3% ZnO/n-Al2O3, and (g,h) 3% CuZnO/n-Al2O3 at 250 °C under 2 MPa for both pure CO2 gas (0–120 min) and after switching the feed gas to H2 (120–220 min).
Nanomaterials 12 02527 g006
Figure 7. In situ DRIFT spectra of the reaction gas mixture of (a,b) n-Al2O3, (c,d) 3% ZnO/n-Al2O3, (e,f) 3% Cu/n-Al2O3, and (g,h) 3% CuZnO/n-Al2O3. Reaction conditions: CO2:H2 = 1:3, gas flow rate = 30 mL min−1, P = 2.0 MPa.
Figure 7. In situ DRIFT spectra of the reaction gas mixture of (a,b) n-Al2O3, (c,d) 3% ZnO/n-Al2O3, (e,f) 3% Cu/n-Al2O3, and (g,h) 3% CuZnO/n-Al2O3. Reaction conditions: CO2:H2 = 1:3, gas flow rate = 30 mL min−1, P = 2.0 MPa.
Nanomaterials 12 02527 g007
Scheme 1. Formation of key intermediate species in hydrogenation of CO2 to methanol ( Nanomaterials 12 02527 i001 bicarbonate, Nanomaterials 12 02527 i002 formate, Nanomaterials 12 02527 i003 methoxy, Nanomaterials 12 02527 i004 hydrogen).
Scheme 1. Formation of key intermediate species in hydrogenation of CO2 to methanol ( Nanomaterials 12 02527 i001 bicarbonate, Nanomaterials 12 02527 i002 formate, Nanomaterials 12 02527 i003 methoxy, Nanomaterials 12 02527 i004 hydrogen).
Nanomaterials 12 02527 sch001
Table 1. EDX analysis of the 3% CuZnO/n-Al2O3 catalyst.
Table 1. EDX analysis of the 3% CuZnO/n-Al2O3 catalyst.
ElementCuZnAlO
(At%)0.810.2940.5758.33
(Wt%)2.460.8852.1844.48
Table 2. Formation of the key intermediates and their peak intensity changes with temperature and catalyst composition *.
Table 2. Formation of the key intermediates and their peak intensity changes with temperature and catalyst composition *.
Temperature
(°C)
n-Al2O33% ZnO/n-Al2O33% Cu/n-Al2O33% CuZnO/n-Al2O3
50BicarbonateBicarbonate (ZnO),
bicarbonate (Al2O3)
Bicarbonate (Al2O3)Bicarbonate (ZnO)
100Bicarbonate ↓,
monodentate, carbonate
bidentate formate
Bidentate carbonate (ZnO) ↓,
bicarbonate (Al2O3) ↓,
polydentate carbonate (Al2O3 or ZnO), bidentate formate (ZnO and Al2O3)
Bicarbonate (Al2O3) ↓, bidentate formate (Al2O3),
polydentate carbonate
Bicarbonate (ZnO) ↓,
polydentate carbonate (Al2O3)
150Bicarbonate ↓,
polydentate carbonate,
bidentate formate ↑
Bidendentate formate (ZnO and Al2O3) ↑,
monodentate carbonate (Al2O3) ↓,
bicarbonate (Al2O3) ↓,
polydentate carbonate (Al2O3) ↑
Bicarbonate (Al2O3) ↓,
bidentate formate (Cu),
bidendentate formate (Al2O3) ↑
Bicarbonate (ZnO) ↑,
polydentate carbonate (Al2O3),
bidentate formate (Al2O3 and ZnO),
bidentate formate (Cu), bidendate formate (ZnO) ↑, methoxy
200Bicarbonate ↓,
monodentate carbonate ↓,
bidentate formate ↑
Bicarbonate (ZnO) ↑,
bidentate carbonate (ZnO) ↑, bidentate formate (ZnO and Al2O3) ↑, polydentate carbonate (ZnO) ↑
Bicarbonate (Al2O3) ↓,
bidentate formate (Cu),
bidentate formate (Al2O3) ↑,
methoxy
Bidentate formate (Al2O3 and ZnO) ↑,
bidentate formate (Cu) ↑, bidentate formate (ZnO) ↑,
bidentate formate (Al2O3), methoxy
250Bicarbonate ↓,
nonodentate carbonate ↓,
polydentate carbonate ↓, bidentate formate ↑
Bicarbonate (ZnO) ↑,
bidentate carbonate (ZnO) ↑, bidentate formate (ZnO and Al2O3) ↑,
formate (Al2O3 ↑)
Bidendate formate (Cu) ↑,
bidentate formate (Al2O3) ↑, methoxy ↑
Bidentate formate (Al2O3 and ZnO) ↑,
bidentate formate (Cu) ↑, bidentate formate (ZnO) ↑,
bidentate formate (Al2O3) ↑, methoxy ↑
300Same as in 250 °CBicarbonate (ZnO) ↑,
bidentate carbonate (ZnO) ↑, bidentate formate (ZnO and Al2O3) ↑,
formate (Al2O3) ↑
Same as in 250°CBidendate formate (Al2O3 and ZnO) ↑,
bidentate formate (Cu) ↑, Bidentate formate (ZnO) ↑, bidentate formate (Al2O3) ↑, methoxy ↑
* ↓—Decrease; ↑—increase.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, L.; Etim, U.J.; Zhang, C.; Amirav, L.; Zhong, Z. CO2 Activation and Hydrogenation on Cu-ZnO/Al2O3 Nanorod Catalysts: An In Situ FTIR Study. Nanomaterials 2022, 12, 2527. https://doi.org/10.3390/nano12152527

AMA Style

Wang L, Etim UJ, Zhang C, Amirav L, Zhong Z. CO2 Activation and Hydrogenation on Cu-ZnO/Al2O3 Nanorod Catalysts: An In Situ FTIR Study. Nanomaterials. 2022; 12(15):2527. https://doi.org/10.3390/nano12152527

Chicago/Turabian Style

Wang, Letian, Ubong Jerome Etim, Chenchen Zhang, Lilac Amirav, and Ziyi Zhong. 2022. "CO2 Activation and Hydrogenation on Cu-ZnO/Al2O3 Nanorod Catalysts: An In Situ FTIR Study" Nanomaterials 12, no. 15: 2527. https://doi.org/10.3390/nano12152527

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