Next Article in Journal
RuNi/MMO Catalysts Derived from a NiAl-NO3-LDH Precursor for CO Selective Methanation in H2-Rich Gases
Next Article in Special Issue
Noble Metal Single-Atom Coordinated to Nitrogen, Oxygen, and Carbon as Electrocatalysts for Oxygen Evolution
Previous Article in Journal
Copper-Catalyzed/Hypervalent Iodine-Mediated Functionalization of Unactivated Compounds
Previous Article in Special Issue
Optimized Ni, Co, Mn Oxides Anchored on Graphite Plates for Highly Efficient Overall Water Splitting
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanism of Methanol Synthesis from CO2 Hydrogenation over Cu/γ-Al2O3 Interface: Influences of Surface Hydroxylation

1
State Key Laboratory of Photocatalysis on Energy and Environment, College of Chemistry, Fuzhou University, Fuzhou 350116, China
2
College of Chemical and Biological Engineering, Yichun University, Yichun 336000, China
3
Fujian Provincial Key Laboratory of Theoretical and Computational Chemistry, Xiamen 361005, China
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(9), 1244; https://doi.org/10.3390/catal13091244
Submission received: 25 July 2023 / Revised: 15 August 2023 / Accepted: 24 August 2023 / Published: 27 August 2023
(This article belongs to the Special Issue Theory-Guided Electrocatalysis and Photocatalysis)

Abstract

:
The adsorption and hydrogenation of carbon dioxide on γ-Al2O3(110) surface-supported copper clusters of different sizes are investigated using density functional theory calculations. Our results show that the activation of CO2 is most obvious at the Cu/γ-Al2O3 interface containing the size-selected Cu4 cluster. It is interesting that the CO2 activation is more pronounced at the partially hydroxyl-covered interface. The catalytic mechanisms of CO2 conversion to methanol at the dry and hydroxylated Cu4/γ-Al2O3 interfaces via the formate route and the pathway initiated through the hydrogenation of carbon monoxide produced by the reverse water–gas shift reaction are further explored. On both interfaces, the formate pathway is identified as the preferred reaction pathway, in which the hydrogenation of HCOO to H2COO is the rate-limiting step (RLS). However, since the surface OH group can act as a hydrogen source in some elementary reactions, unlike the dry surface, the production of H2COOH species along the formate pathway is found at the hydroxylated interface. In addition, the introduction of OH at the interface leads to an increase in the kinetic barrier of the RLS, indicating that surface hydroxylation has a negative effect on the catalytic activity of CO2 conversion to CH3OH at the Cu/γ-Al2O3 interface.

1. Introduction

The hydrogenation of carbon dioxide to methanol has attracted considerable attention because it offers an efficient way to reduce CO2 concentrations [1,2,3,4,5], the resulting methanol is high-value-added and it can be used as a renewable fuel and a convenient feedstock for chemical processes [6,7,8]. Due to the chemical inertness of CO2, many catalysts have been proposed to achieve the conversion of CO2 to CH3OH, in which the most representative is the metal catalyst supported on the oxide surface [3,9].
Many investigations have revealed that metal (Cu, Ni and Pd) particles dispersed on the oxide supports (such as ZnO, CeO2, Al2O3, TiO2, ZrO2, In2O3, Ga2O3, etc.) can greatly promote the catalytic activity of CO2 conversion [4,9,10], which can be attributed to the synergistic effect at the interface between the metal particle and substrate. For instance, Pd/In2O3 and Au/CeOx/TiO2 catalysts facilitate the hydrogenation of CO2 to methanol [11,12]. As the main active component of Cu/ZnO/Al2O3 catalysts used in industry, Cu-based catalysts have attracted extensive attention in recent years [5,13,14]. Although CO2 is weakly adsorbed on the surface of pure copper, resulting in poor activity for CH3OH production [15,16,17], the deposition of Cu particles on oxide supports can modify the coordination environment and electronic property of Cu sites, thus, exhibiting a unique catalytic activity of CO2 conversion. Both experimental and theoretical investigations have demonstrated that Cu/TiO2 and Cu/ZrO2 catalysts can achieve CO2 hydrogenation to methanol, and the reverse water–gas shift (RWGS) + CO hydrogenation pathway has been proven to be the main reaction route for the overall conversion [18,19]. Liu et al. investigated the mechanism of CO2 hydrogenation to methanol for Cu4 nanoparticles deposited on an amorphous Al2O3 surface [5]; their results indicated that the Cu4/Al2O3 thin film showed effective catalytic activity for CO2 reduction to CH3OH at low CO2 partial pressure, in which the formate route was the most favorable pathway. The above results demonstrated that the catalytic mechanism of CO2 hydrogenation is closely related to the type of substrate. Moreover, the performance of Cu/oxide catalysts is also sensitive to the sizes of the supported metal particles [20]. As an example, Yang and coworkers studied methanol synthesis over three kinds of size-selected Cun clusters (n = 3, 4, 20) deposited on Al2O3 through experimental measurements combined with density functional theory (DFT) calculations, observing a strong cluster size dependence of catalytic activity [21]. Tao et al. explored CO2 conversion at the Cu/TiO2 interfaces formed through the deposition of Cun clusters of different sizes, and their results showed that the TiO2 surface-supported size-selected Cu4 cluster exhibited the highest activity for the production of CH3OH [22]. Therefore, the type of support and the size of the copper clusters were shown to have a decisive impact on the catalytic activity as well as the reaction mechanisms of Cu/oxide catalysts.
As an important oxide, γ-Al2O3 is widely used as a support for heterogeneous catalysis. For examples, Ni/γ-Al2O3, Pd/γ-Al2O3 and Cu/γ-Al2O3 have been demonstrated to be efficient for CO2 conversion. It is well known that the surface of aluminum is easy to hydroxylate. Experimental results show that water can be adsorbed on the Al2O3 surface in both molecular and dissociative forms. Although water molecules are easily removed, surface hydroxyl groups are still present even when the temperature reaches 100 °C [23]. Theoretical studies have found that water molecules can be easily chemisorbed at exposed Lewis acidic Al sites and then exothermically dissociated into OH and H species with a low kinetic barrier [24]. Moreover, water can be produced during the CO2 hydrogenation process, which provides a source of hydroxylation. Therefore, the surface of the γ-Al2O3 support can be hydroxylated under realistic reaction conditions, and it can be expected for hydroxyl groups to modify the properties of the aluminum surface [23]. Pan and Zhang et al. studied CO2 hydrogenation over Ni/γ-Al2O3 and Pd/γ-Al2O3, respectively [25,26], and they focused on the influence of surface hydroxylation on the selectivity of the HCOO transformation to methanol and methane. On the hydroxylated SiC surface, the OH groups directly participated in CO2 hydrogenation by providing H atoms [27], and the hydroxyls on the surface of the ZnZrOx solid solution could promote CO2 conversion to methanol [28]. In contrast, the presence of water accelerated the sintering and deactivation of the Cu/ZnO catalyst [3,29]. As for the Cu/γ-Al2O3 interface concerned in this paper, the in-depth study of how the surface hydroxyl group affects the mechanism of CO2 hydrogenation to methanol is quite limited at the present [30].
Here, we report a systematic theoretical investigation of small copper clusters supported on dry and hydroxylated γ-Al2O3(110) surfaces that were chosen as model catalysts for CO2 activation and hydrogenation to CH3OH. The structures of the Cun clusters of different sizes supported on the γ-Al2O3 surface were determined. Two main pathways, including the formate route and the RWGS + CO hydrogenation route, were taken into account for the CO2 conversion. A partially hydroxylated γ-Al2O3(110) surface that could provide Lewis acid sites for CO2 binding and hydroxyls for the hydrogen transfer was constructed to explore the influences of surface hydroxylation on methanol synthesis from CO2 hydrogenation.

2. Computational Details

All spin-polarized DFT calculations were performed using the Vienna ab initio simulation package (VASP) [31,32]. The generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerhof (PBE) exchange–correlation functional was employed, and the projector-augmented plane wave (PAW) was used to describe the electron–ion interactions [33]. A cutoff energy of 500 eV was set for the plane-wave expansion and the dipole correction in the surface normal direction was applied. The effects of van der Waals interactions were considered by using the vdw-DF3 functional [34], and Γ-centered k-points were used for the slab calculation. The convergences of the total energy and force were set to 1 × 10−5 eV and 0.03 eV/Å, respectively.
A periodic slab including six atomic layers was employed to simulate the γ-Al2O3 (110) surface. For the depositions of a single Cu atom and Cu2 dimer, a (2 × 1) supercell was adopted, while a large (2 × 2) supercell containing 48 Al2O3 units (i.e., 96 Al and 144 O atoms) was used for the cases of supporting other Cun clusters of large sizes. During the structural optimizations, the atoms at the bottom two layers were fixed at their optimized bulk positions, whereas the positions of other atoms, including the Cun cluster and the adsorbed species, were allowed to relax. The thickness of the vacuum was set to 12 Å to avoid interactions between the adjacent slabs. The hydroxylation surface was obtained through the dissociation of adsorbed water molecules on the γ-Al2O3(110) surface. Additionally, molecular dynamics (MD) simulations with a low-energy cutoff of 300 eV were carried out to explore possible configurations for the depositions of Cun clusters on the dry and hydroxylated γ-Al2O3(110) surface using the Nosé algorithm [35]. The simulation time was 10 ps with a step of 1 fs at a temperature of 600 K. Then, typical configurations were sampled from the results of the MD simulations and, finally, further structural optimizations using the above accurate settings were performed to determine the most stable configuration of the Cun/γ-Al2O3(110) surfaces.
The adsorption energy of the reaction intermediate (Eads) was defined as,
Eads = E(Cun/γ-Al2O3) + E(X) − E(X/Cun/γ-Al2O3)
where X presents the intermediate involved in the methanol synthesis, E(X/Cun/γ-Al2O3), E(Cun/γ-Al2O3) and E(X) represent the total energies of the adsorbed system, the pristine Cun/γ-Al2O3(110) surface and the intermediate in the gas phase, respectively.
To study the reaction mechanisms of CO2 hydrogenation to CH3OH, the climbing image nudged elastic band (CI-NEB) method was adopted for investigating the reaction paths, including the structures of the transition states (TSs) by using eight images [36,37]. All the structures of TSs were further confirmed through vibrational calculations. The activation barrier (Ea) of each elementary reaction for the synthesis of methanol was defined as,
Ea = E(TS) − E(IS)
where E(TS) and E(IS) represent the total energies of the TS and the initial state (IS), respectively. The associated reaction energy (Er) of each step was expressed as the energy difference between the final state (FS) and the IS, namely,
Er = E(FS) − E(IS)
The values of Ea and Er reported in the following were corrected by considering the zero-point energy (ZPE) contribution. Furthermore, a method proposed by Campbell for microkinetic simulations was performed to determine the rate-limiting step (RLS) by evaluating the degree of rate control (DRC) [38,39]. The information about the microkinetic simulation was presented in detail elsewhere [40].

3. Results and Discussion

3.1. Structure of γ-Al2O3(110) Surface

As a material with a low crystallinity, the structure of Al2O3 in the γ phase is still under debate. Two typical crystallographic structures, named the spinel-like and nonspinel models, were proposed. Although the spinel-like model is commonly used, it does not satisfy the existing theoretical and experimental results well [41]. The nonspinel model proposed by Digne agrees well with the experimental spectra [42], especially the peaks associated with the surface features [43,44]. Herein, the nonspinel structure was adopted to simulate the (110) surface, which accounted for more than 70% of the total exposed area of γ-Al2O3. For the bulk of nonspinel γ-Al2O3 crystallized in the monoclinic phase (see the inset in Figure 1), the optimized lattice parameters were a = 5.536 Å, b = 8.345 Å, c = 8.037 Å and β = 90.6°, respectively. These values were in good agreement with the results reported in other theoretical works [45]. The top and side views of the structure of (2 × 2) the supercell of the clean γ-Al2O3(110) surface are shown in Figure 1, with an area of approximately 17.1 × 16.5 Å2. The surface energy of γ-Al2O3(110) was calculated to be 2.33 J·m−2, consistent with the value of 2.59 J·m−2 reported by Digne et al. [42]. It was clear that two-fold and three-fold coordinated O atoms were observed at the exposed (110) facet, while the surface Al atoms were three-fold, four-fold and five-fold coordinated. To facilitate the subsequent discussion, the symbols O2c, O3c, Al3c, Al4c and Al5c were used to distinguish the surface oxygen and aluminum atoms from different coordination environments (Figure 1b). Moreover, two kinds of Al4c atoms were denoted by Al4c-I and Al4c-II, respectively. It must be noted that the coordination-unsaturated Al/O atoms were the active sites for the adsorptions of molecules on the γ-Al2O3(110) surface.

3.2. The Hydroxylation of γ-Al2O3(110) Surface

It is well known that the presence of water molecules easily leads to the hydroxylation of γ-Al2O3(110); in this process, the adsorption and dissociation of H2O are closely related to the strength of the Lewis acid of the unsaturated Al sites on the surface. This can be attributed to the difference in the coordination environment of the surface Al atoms, and more specifically, Al3c has a stronger Lewis acidity than Al4c and Al5c atoms. According to the experimental results [46,47], the concentration of the OH group on the γ-Al2O3(110) surface varied in a wide range; it decreased from 11.8 to 3.0 OH·nm−2 as the temperature increased from 500 K to 1000 K.
Considering that the coordination-unsaturated sites are required for the support of copper clusters and the adsorption of CO2 molecules, in this work, a partially hydroxylated γ-Al2O3(110) surface with a coverage of 5.9 OH·nm−2 was employed to build the Cu-supported model catalyst (see Figure 2a), which was constructed through the decomposition of two H2O molecules on the surface with a (2 × 1) supercell. Since H2O is most strongly bound to the Al3c atom with an adsorption energy of Eads = 2.40 eV, the decomposition of the water on the γ-Al2O3(110) surface began at the Al3c site. As displayed in Figure 3a, in the FS, the resulting hydroxyl group was still attached to the Al3c atom, while the hydrogen atom formed a bond with a neighboring O2c atom. This step was exothermic by 1.0 eV, with a small activation energy of Ea = 0.16 eV. For the second water, it could be adsorbed at the Al4c-I or Al4c-II site; however, according to the values of Eads (1.44 eV vs. 1.09 eV), H2O molecules favor the Al4c-I site (Figure 3b). After breaking one O-H bond, the OH group occupied the bridging site between two Al4c-I through an O atom; meanwhile, the H atoms bound to the oxygen atom near the Al4c-I site (i.e., O3c atom) in the FS. This step required a small barrier of Ea = 0.20 eV, and it was also an exothermic process with Er = −0.57 eV.

3.3. Structures of Small Copper Clusters Supported on γ-Al2O3 Surface

Firstly, the structures of small copper particles of different sizes in the gas phase were examined. Consistent with the results of other theoretical studies [48,49,50], for Cun (n = 1–6) clusters in the ground state, they all exhibited coplanar configurations; the calculated lengths of unequal Cu–Cu bonds of all clusters are provided in Figure 4.
Figure 5 shows the most stable structures of copper clusters supported on a dry γ-Al2O3(110) surface without an OH group obtained through MD simulations, and the corresponding binding energy (Eb) and average lengths of the Cu–O and Cu–Cu bonds are listed in Table 1. Similar to the case of the Al-terminated α-Al2O3(0001) surface [51], the Cu atoms of the clusters tended to bind to oxygen atoms with low coordination. For the deposition of a single Cu atom, it preferred to form bonds with two surface O2C atoms (Figure 5a), with the average length of two Cu–O bonds being approximately 1.81 Å. Both the copper dimer and trimer were anchored on the γ-Al2O3(110) surface by forming three Cu–O bonds (Figure 5b,c) with an average length of 2.02 and 2.15 Å, respectively. For the remaining copper clusters of relatively large sizes, namely, Cu4, Cu5 and Cu6, they were still dispersed on the surface with an approximate planar configuration, and the numbers of Cu–O bonds at the interface were four, six and six (Figure 5d–f), with average Cu–O bond lengths of 2.11, 2.05 and 2.07 Å, respectively. According to the Eb values given in Table 1, it was clear that the binding strength between the copper clusters and the supports increased gradually with the increase in the size of the copper clusters. Owing to the formation of a Cu–O bond at the Cu/Al2O3 interface, the electrons transferred from Cun to the support, resulting in copper particles carrying a positive charge (see Table 1). This also caused the obvious decrease in the surface work function, especially for the system containing Cu4 clusters (approximately 2.4 eV). Hence, for the Al2O3 surface supported with Cu4 clusters, it was relatively easy to donate electrons to the CO2 molecules.
Since the Cu4-supported Al2O3 surface exhibited the strongest activation of CO2 (see below), we also investigated the most stable configuration of Cu4 clusters deposited on the hydroxylated surface. As shown in Figure 2b, now, Cu4 was still anchored to the surface through four Cu–O bonds, with an average length of 2.11 Å. However, the planar structure in the gas phase (Figure 4c) turned to a tetrahedron-like configuration, indicating that Cu4 had good flexibility to accommodate changes in the surface structure. Besides the remarkable variation of the configuration, the introduction of OH groups could significantly enhance the binding strength of Cu4 on the γ-Al2O3(110) surface, and the value of Eb increased from 4.55 eV on the dry surface to 7.08 eV (Table 1). Therefore, the Cu4 clusters deposited on the hydroxylated γ-Al2O3(110) surface were more stable than that on the dry surface [52]. In this case, the Cu4 particles carried a more positive charge than if it was deposited on a dry surface. Moreover, to explore how hydroxyls affect copper clusters of various sizes, the most stable configurations of other Cun clusters were also determined, and the corresponding results are displayed in Figure S1 and Table S1 in the Supplementary Information. Similar to Cu4, now, Cu5 and Cu6 particles of large size also exhibited a three-dimensional stacking structure on the hydroxylated surfaces. This could be attributed to the fact that some surface sites were blocked by hydroxyl groups, making it difficult for large Cu clusters to highly disperse on the hydroxylated surface. By examining the value of Eb, it was interesting to see that the introduction of the OH group caused a slight reduction in the binding energy of Cu1 and Cu2, but the reverse variation was observed for those Cun clusters with n ≥ 3. For example, the Eb of Cu2 decreased slightly from 3.44 eV to 3.28 eV, while it increased obviously from 5.30 eV to 7.12 eV for Cu6 clusters. Hence, it seems that hydroxylation could improve the stability of large copper clusters on the alumina surface.

3.4. CO2 Activation on γ-Al2O3 Surface

The activation of CO2 can be viewed as an important prerequisite for CO2 conversion. For the Cu metal surface, owing to the weak interaction with CO2, it shows a low ability for CO2 activation [15,16]. In contrast, Cu/oxide catalysts can effectively activate CO2 and, correspondingly, they show good catalytic activity for CO2 hydrogenation to CH3OH [4,5]. Generally, CO2 molecules are chemisorbed on the surface to form CO2δ− species with a bent configuration. For each Cun/γ-Al2O3(110) system, different surface sites for CO2 adsorption were considered. The most stable adsorption structures are shown in Figure 6, and the value of Eads and the structural parameters of CO2 moiety are given in Table 2. As displayed in Figure 6a, the CO2 molecules were weakly adsorbed on the pristine γ-Al2O3(110) surface with a small value of Eads = 0.45 eV (Table 2), and the nearly linear configuration (∠O–C–O = 176.5°) indicated that CO2 was poorly activated. However, for those Cun-supported surfaces, the bent arrangement of the CO2 moiety was observed. After adsorption, the lengths of two C–O bond of CO2 were stretched from 1.18 Å in the gas phase to 1.2–1.3 Å; meanwhile, the ∠O–C–O bond angles were reduced to approximately 130° (Table 2). The adsorption energy of the Cun-supported surfaces was predicted to be in the range of 0.64–0.86 eV, which was obviously higher than that of Pd4/In2O3(110) (Eads = −0.37 eV) [11]. Hence, the deposition of copper clusters could noticeably promote the activation of CO2 on the alumina surface and facilitate the hydrogenation of CO2 to methanol. It could be seen from Table 2 that the value of Eads of Cu4/Al2O3 was the largest (0.86 eV), in which CO2 was adsorbed at the interface through the formation of two Cu–C bonds and one Al4c–Ob bond (Figure 6e). Due to the longest C–O bond (1.314 Å) and the smallest ∠O–C–O bond angle (127.8°) of the CO2 moiety, Cu4/Al2O3 was the most favorable candidate for CO2 hydrogenation to CH3OH among six Cun/γ-Al2O3 surfaces.
The Bader charges of the Cun clusters and CO2 moiety of different Cun/Al2O3 systems are also presented in Table 2. Due to the receipt of electrons from the surface, the CO2 moiety carried a negative charge. The most negative charge (−0.97 e) was especially obtained on Cu4/Al2O3, which confirmed the distinct activation of CO2 on the Cu4-supported surface. For the Cu4 cluster, CO2 adsorption induced a significant increase in the positive charge from 0.35 e to 0.82 e. Furthermore, by comparing the amounts of charges of the CO2 moiety and Cu4, it could be seen that some electrons (approximately 0.15 e) were transferred from the Al2O3 support to CO2. This was consistent with the fact an Al4c–O adsorption bond was formed at the interface. The above results revealed the synergistic behavior of different constituents of Cu4/γ-Al2O3 catalysts.
Moreover, the adsorption of CO2 on the hydroxylated Cu4/γ-Al2O3(110) surface was also explored. As displayed in Figure 2c, CO2 was adsorbed on the surface via two Cu–C bonds, one Al4c–Ob bond and one Cu–Oa bond. Due to the formation of an additional Cu–O bond, the value of Eads increased obviously from 0.86 eV to 1.06 eV (Table 2). The relatively strong binding of CO2 also meant that CO2 was activated more obviously on the hydroxylated surface. Accordingly, a larger average length of C–O bonds and a smaller ∠O-C-O bond angle were found after adsorption. At the same time, CO2 accepted more electrons from the surface, and now, the charge of the CO2 moiety was predicted to be −1.1 e.
Based on the above results, in the following sections, the dry and hydroxylated Cu4/γ-Al2O3(110) surfaces were employed as model catalysts to investigate the reaction mechanisms for the CO2 to CH3OH transformation.

3.5. CO2 Conversion to Methanol on the Cu4/γ-Al2O3 Surface

Two routes, namely, the formate pathway and the RWGS + CO hydrogenation pathway, were considered in the present work. In the formate pathway, CO2 was directly hydrogenated to CH3OH via the formation of the HCOO*, H2COO*, H2CO* and H3CO* intermediates. While, in the other pathway, carbon monoxide (CO*) was generated firstly via the RWGS process; then, CO* was sequentially hydrogenated to methanol through the formation of the HCO*, H2CO* and H3CO* intermediates.

3.5.1. Reaction Mechanism on the Dry Surface

The calculated kinetic barriers (Ea) and the reaction energies (Er) of all the elementary reaction steps involved in the two pathways on the dry Cu4/γ-Al2O3(110) surfaces are given in Table 3, and the optimized structures of the corresponding intermediates and TSs are provided in Figure S2.
The potential energy profile for CH3OH synthesis from CO2 and H2 via the formate pathway (steps F1 to F6 denoted in Table 3) is shown in Figure 7a. The formate route began with the hydrogenation of the preadsorbed CO2 to HCOO* (i.e., step F1). This process only had to overcome a small energy barrier of Ea = 0.28 eV, and it was exothermic by Er = −1.29 eV. In the next step, the second H* attacked the C atom of HCOO* to form the H2COO* intermediate (step F2, Ea = 1.18 eV, Er = 0.71 eV). The activation barrier of this elementary step was lower than that of the Cu(111) surface (1.60 eV) and Cu29 nanoparticle (1.41eV) [5]. Compared to HCOO* species, the C−Oa bond in H2COO* elongated obviously from 1.28 to 1.51 Å, indicating the easy cleavage of the C−Oa bond. Therefore, in the subsequent step, as the third H* was bonded to the Oa atom of H2COO*, the C–Oa bond was broken to yield the H2CO* intermediate (step F3, Ea = 0.58 eV, Er = −0.58 eV). After that, the separated OH* group was hydrogenated to water (step F4, Ea = 0.82 eV, Er = −0.50 eV). With the desorption of H2O*, the remaining H2CO* species was adsorbed on the Cu4 cluster through the formation of a new Cu−C bond. Then, H2CO* was hydrogenated to H3CO* by introducing the fifth H* atom. The results of Ea = 0.13 eV and Er = −1.25 eV indicated that step F5 could easily occur from the point of view of thermodynamics and dynamics. For step F6, it corresponded to the hydrogenation of H3CO* to generate the target product of CH3OH with Ea = 0.68 eV and Er = −0.09 eV. According to the energy profile shown in Figure 7a, similar to the cases of the Cu(111) surface and Cu29 nanoparticles, step F2 was the key step with the highest energy barrier of Ea = 1.18 eV. Moreover, the binding energies of the intermediate were predicted to be 3.50 (HCOO*), 4.82 (H2COO*), 3.87 (H2COOH*), 2.20 (H2CO*) and 4.08 eV (H3CO*), respectively.
In order to identify the RLS of the formate pathway at different temperatures, microkinetic simulations were performed to determine the values of the degree of rate control (DRC) proposed by Campbell [38,39], in which the total pressure was set to 1 atm with a mixture of CO2 and H2 feed gases in a ratio of 1:3. From the DRC curves shown in Figure 8a, it was clear that the hydrogenation of HCOO* (i.e., step F2) was the RLS at temperatures below 560 K, while the elementary reaction for the generation of H2CO* species (i.e., step F4) became the RLS at high temperatures.
As for the RWGS + CO hydrogenation pathway, it is initiated by CO2* hydrogenation to COOH* intermediate (i.e., step R1 in Table 3). Now, the preadsorbed H atom formed an O−H bond with the Oa atom of CO2*, resulting in an obvious elongation of the C−Oa bond from 1.21 Å to 1.34 Å. The calculated values of Ea and Er of step R1 were 0.78 eV and −0.46 eV (Table 3), respectively. In the next step, CO* was produced by breaking the C−Oa bond of COOH*. The high energy barrier of Ea = 1.55 eV with a large positive reaction energy of Er = 0.79 eV indicated that step R2 was kinetically and thermodynamically unfavorable. For step R3, it was associated with the hydrogenation of CO* to produce the HCO* intermediate with Ea = 0.88 eV and Er = −0.09 eV. Then, the HCO* species could be further hydrogenated to H2CO* (step R4, Ea = 0.44 eV, Er = −0.06 eV). The remaining process followed the same steps F5 and F6 of the formate pathway, which converted the H2CO* intermediate to methanol. Therefore, the entire process of the RWGS + CO hydrogenation route was to form intermediates in the sequence of COOH*, CO*, HCO*, H2CO* and H3CO*, and the corresponding binding energies were 3.60, 1.21, 3.30, 2.20 and 4.08 eV, respectively. According to the energy profile displayed in Figure 7a, the elementary reaction for the produce of the CO* intermediate via the dissociation of the COOH* species (i.e., step R2) was the key step with the highest energy barrier of Ea = 1.55 eV. The results of the DRC analysis (Figure 8b) also confirmed that the step R2 was the RLS of the RWGS + CO hydrogenation pathway in the temperature range investigated (430–650 K).

3.5.2. Reaction Mechanism on the Hydroxylated Surface

The values of the Ea and Er of each elementary reaction of two pathways on the hydroxylated Cu4/γ-Al2O3(110) surfaces are listed in Table 4, and the optimized structures of the intermediates and TSs are shown in Figure S3. It is worth noting that in some steps, such as steps F3 and F4, the hydrogen atom (denoted by Hs in Table 4) of the OH group anchored on the alumina surface could participate in the hydrogenation processes.
Owing to the more obvious activation of CO2 on the hydroxylated surface as mentioned above, step F1 of the formate pathway of CO2* hydrogenation to HCO2* was more likely to occur by overcoming a smaller kinetic barrier of Ea = 0.15 eV and releasing more energy of Er = −1.50 eV (Table 4). However, the further conversion of HCOO* species to H2COO* through step F2 required climbing a high energy barrier of Ea = 1.34 eV, approximately 0.16 eV larger than the dry surface. Unlike the above steps, where the H atom was preadsorbed on the Cu4 clusters, for step F3, where H2COO* was hydrogenated to H2COOH*, the hydrogen atom came from the surface OH* group. The optimized structures involved in step F3 are shown in Figure S3; the migration of the Hs atom from OH* to the Oa atom of the H2COO* species was observed. Meanwhile, the nearby hydrogen atom moved towards the hydroxyl O atom lacking H; then, the OH* group was reformed. The small activation energy of Ea = 0.02 eV and the negative value of Er = −0.56 eV indicated that step F3 was prone to occur. For comparison, a different product (i.e., H2CO*) was obtained for step F3 on the dry surface (Table 3). Therefore, the using of surface OH* as the source of hydrogen had an effect on the detailed reaction mechanism of CO2 conversion to methanol. Subsequently, the H2CO* intermediate and H2O* were created through the further hydrogenation of H2COOH* (step F4, Ea = 0.79 eV, Er = 0.57 eV), in which the H atom still originated from the surface hydroxyl group. After the desorption of H2O*, H2CO* was hydrogenated to form H3CO* through step F5 with Ea = 0.17 eV and Er = −0.58 eV. In the last elementary reaction, methanol was produced from H3CO* (step F6, Ea = 0.73 eV, Er = 0.25 eV). According to the energy profile displayed in Figure 7b, similar to the dry surface, step F2 was still the key step with the highest energy barrier (Ea = 1.34 eV).
For the RWGS + CO hydrogenation pathway on the hydroxylated Cu4/γ-Al2O3(110) surface, the COOH* species was still generated through the hydrogenation of CO2* (step R1, Ea = 0.73 eV, Er = −0.82 eV). Then, the second H* atom attacked the Oa of COOH*, leading to the formation of CO* and H2O* (step R2, Ea = 1.91 eV, Er = 1.22 eV). Comparing the results of the same step occurring on the dry surface (Table 3), it seemed that the covering of the OH group prevented the formation of the CO intermediate. The next two elementary reactions corresponded to the hydrogenation of CO* (step R3, Ea = 0.56 eV, Er = −0.43 eV) and HCO* (step R4, Ea = 0.65 eV, Er = −0.37 eV), respectively. After that, methanol was produced through steps F5 and F6 of the formate route in turn. Obviously (Figure 7b), the RLS of the RWGS + CO hydrogenation pathway was the hydrogenation of COOH* (i.e., step R2).

4. Conclusions

In the present work, model catalysts of small copper clusters supported on the γ-Al2O3(110) surface were constructed to investigate CO2 activation and its conversion to methanol using DFT calculations. The reaction mechanisms for CH3OH synthesis on the dry and hydroxylated surface were explored to understand the influences of surface hydroxylation for such a kind of Cu/oxide interface. Our results showed that copper clusters tended to be deposited on the alumina surface in the form of a monolayer. CO2 molecules could be activated effectively at the interface, which was composed of Cun clusters and exposed Lewis acidic Al sites, especially the systems containing the Cu4 cluster. For two possible reaction routes, the formate pathway was more favorable than the RWGS + CO hydrogenation pathway, in which the RLS was the hydrogenation of HCOO* to the H2COO species. On the hydroxylated surface, the hydroxyl group could provide its H atom to participate in some hydrogenation steps, resulting in the formation of the H2COOH* intermediate. Although the RLSs on two surfaces were the same, the corresponding kinetic barrier of the hydroxylated surface was higher than that of the dry surface. In addition, the OH group would occupy Al3c sites on the surface responsible for CO2 activation via the synergy of copper clusters and the support. In summary, surface hydroxylation was not favorable for CO2 conversion to methanol over Cu/Al2O3 catalysts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13091244/s1., Figure S1: Top and side views of the most stable configurations of Cun clusters supported on hydroxylated γ-Al2O3(110) surface, Figure S2: Optimized structures of the initial state, the transition state and the final state of each elementary reaction on the dry γ-Al2O3(110) surface, Figure S3: Optimized structures of the initial state, the transition state, and the final state of each elementary reaction on the hydroxylated Cu4/γ-Al2O3(110) surface; Table S1: Calculated results of Cun clusters supported on the hydroxylated γ-Al2O3(110) surface.

Author Contributions

Investigation and writing, H.Z. and H.J.; investigation, Y.L. (Yanli Li); funding acquisition, Y.L. (Yi Li), S.H. and W.C.; supervision, W.L.; review and editing, Y.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (grant nos. 21773030, 21973014, 21863004, 21703036 and 21373048), and the Natural Science Foundations of Fujian province (no. 2022J01066) and Jiangxi province (no. 20192BAB206035). The numerical calculations in this paper were performed on the supercomputing system in the Supercomputing Center of Fujian of China.

Data Availability Statement

The data presented in this study are available on request from the corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Karelovic, A.; Ruiz, P. The Role of Copper Particle Size in Low Pressure Methanol Synthesis Via CO2 Hydrogenation over Cu/ZnO Catalysts. Catal. Sci. Technol. 2015, 5, 869–881. [Google Scholar] [CrossRef]
  2. Sun, K.; Fan, Z.; Ye, J.; Yan, J.; Ge, Q.; Li, Y.; He, W.; Yang, W.; Liu, C.-J. Hydrogenation of CO2 to Methanol over In2O3 Catalyst. J. CO2 Util. 2015, 12, 1–6. [Google Scholar] [CrossRef]
  3. Rodriguez, J.A.; Liu, P.; Stacchiola, D.J.; Senanayake, S.D.; White, M.G.; Chen, J.G.G. Hydrogenation of CO2 to Methanol: Importance of Metal-Oxide and Metal-Carbide Interfaces in the Activation of CO2. ACS Catal. 2015, 5, 6696–6706. [Google Scholar] [CrossRef]
  4. Kapiamba, K.F.; Otor, H.O.; Viamajala, S.; Alba-Rubio, A.C. Inverse Oxide/Metal Catalysts for CO2 Hydrogenation to Methanol. Energ. Fuel. 2022, 36, 11691–11711. [Google Scholar] [CrossRef]
  5. Liu, C.; Yang, B.; Tyo, E.; Seifert, S.; DeBartolo, J.; von Issendorff, B.; Zapol, P.; Vajda, S.; Curtiss, L.A. Carbon Dioxide Conversion to Methanol over Size-Selected Cu4 Clusters at Low Pressures. J. Am. Chem. Soc. 2015, 137, 8676–8679. [Google Scholar] [CrossRef]
  6. Aresta, M. (Ed.) Carbon Dioxide as Chemical Feedstock; Wiley-VCH: New York, NY, USA, 2010. [Google Scholar]
  7. Kondratenko, E.V.; Mul, G.; Baltrusaitis, J.; Larrazabal, G.O.; Perez-Ramirez, J. Status and Perspectives of CO2 Conversion into Fuels and Chemicals by Catalytic, Photocatalytic and Electrocatalytic Processes. Energ. Environ. Sci. 2013, 6, 3112–3135. [Google Scholar] [CrossRef]
  8. Porosoff, M.D.; Yan, B.; Chen, J.G. Catalytic reduction of CO2 by H2 for synthesis of CO, methanol and hydrocarbons: Challenges and opportunities. Energ. Environ. Sci. 2016, 9, 62–73. [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. Fayisa, B.A.; Yang, Y.; Zhen, Z.; Wang, M.-Y.; Lv, J.; Wang, Y.; Ma, X. Engineered Chemical Utilization of CO2 to Methanol via Direct and Indirect Hydrogenation Pathways: A Review. Ind. Eng. Chem. Res. 2022, 61, 10319–10335. [Google Scholar] [CrossRef]
  11. Ye, J.; Liu, C.-j.; Mei, D.; Ge, Q. Methanol Synthesis from CO2 Hydrogenation over a Pd4/In2O3 Model Catalyst: A Combined DFT and Kinetic Study. J. Catal. 2014, 317, 44–53. [Google Scholar] [CrossRef]
  12. Yang, X.F.; Kattel, S.; Senanayake, S.D.; Boscoboinik, J.A.; Nie, X.W.; Graciani, J.; Rodriguez, J.A.; Liu, P.; Stacchiola, D.J.; Chen, J.G.G. Low Pressure CO2 Hydrogenation to Methanol over Gold Nanoparticles Activated on a CeOx/TiO2 Interface. J. Am. Chem. Soc. 2015, 137, 10104–10107. [Google Scholar] [CrossRef]
  13. Murthy, P.S.; Liang, W.; Jiang, Y.; Huang, J. Cu-Based Nanocatalysts for CO2 Hydrogenation to Methanol. Energ. Fuel. 2021, 35, 8558–8584. [Google Scholar] [CrossRef]
  14. Cui, X.; Liu, Y.; Yan, W.; Xue, Y.; Mei, Y.; Li, J.; Gao, X.; Zhang, H.; Zhu, S.; Niu, Y.; et al. Enhancing methanol selectivity of commercial Cu/ZnO/Al2O3 catalyst in CO2 hydrogenation by surface silylation. Appl. Catal. B Environ. 2023, 339, 123099. [Google Scholar] [CrossRef]
  15. Grabow, L.C.; Mavrikakis, M. Mechanism of Methanol Synthesis on Cu through CO2 and CO Hydrogenation. ACS Catal. 2011, 1, 365–384. [Google Scholar] [CrossRef]
  16. Yang, Y.X.; Evans, J.; Rodriguez, J.A.; White, M.G.; Liu, P. Fundamental Studies of Methanol Synthesis from CO2 Hydrogenation on Cu(111), Cu Clusters, and Cu/ZnO(000ī). Phys. Chem. Chem. Phys. 2010, 12, 9909–9917. [Google Scholar] [CrossRef] [PubMed]
  17. Qiu, M.; Fang, Z.; Li, Y.; Zhu, J.; Huang, X.; Ding, K.; Chen, W.; Zhang, Y. First-Principles Investigation of the Activation of CO2 Molecule on Tm/Cu (Tm=Fe, Co and Ni) Surface Alloys. Appl. Surf. Sci. 2015, 353, 902–912. [Google Scholar] [CrossRef]
  18. 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]
  19. Tang, Q.L.; Hong, Q.J.; Liu, Z.P. CO2 fixation into methanol at Cu/ZrO2 interface from first principles kinetic Monte Carlo. J. Catal. 2009, 263, 114–122. [Google Scholar] [CrossRef]
  20. Natesakhawat, S.; Lekse, J.W.; Baltrus, J.P.; Ohodnicki, P.R.; Howard, B.H.; Deng, X.Y.; 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]
  21. Yang, B.; Liu, C.; Halder, A.; Tyo, E.C.; Martinson, A.B.F.; Seifert, S.; Zapol, P.; Curtiss, L.A.; Vajda, S. Copper Cluster Size Effect in Methanol Synthesis from CO2. J. Phys. Chem. C 2017, 121, 10406–10412. [Google Scholar] [CrossRef]
  22. Tao, H.L.; Li, Y.; Cai, X.; Zhou, H.; Li, Y.; Lin, W.; Huang, S.; Ding, K.; Chen, W.; Zhang, Y. What Is the Best Size of Subnanometer Copper Clusters for CO2 Conversion to Methanol at Cu/TiO2 Interfaces? A Density Functional Theory Study. J. Phys. Chem. C 2019, 123, 24118–24132. [Google Scholar] [CrossRef]
  23. Ahn, J.; Rabalais, J.W. Composition and Structure of the Al2O3{0001}-(1×1) Surface. Surf. Sci. 1997, 388, 121–131. [Google Scholar] [CrossRef]
  24. Ranea, V.A.; Schneider, W.F.; Carmichael, I. DFT Characterization of Coverage Dependent Molecular Water Adsorption Modes on Alpha-Al2O3(0001). Surf. Sci. 2008, 602, 268–275. [Google Scholar] [CrossRef]
  25. Pan, Y.-X.; Liu, C.-J.; Ge, Q. Effect of Surface Hydroxyls on Selective CO2 Hydrogenation over Ni4/γ-Al2O3: A Density Functional Theory Study. J. Catal. 2010, 272, 227–234. [Google Scholar] [CrossRef]
  26. Zhang, R.; Liu, H.; Wang, B.; Ling, L. Insights into the Effect of Surface Hydroxyls on CO2 Hydrogenation over Pd/γ-Al2O3 Catalyst: A Computational Study. Appl.Catal. B Environ. 2012, 126, 108–120. [Google Scholar] [CrossRef]
  27. Peng, Y.; Wang, L.; Luo, Q.; Cao, Y.; Dai, Y.; Li, Z.; Li, H.; Zheng, X.; Yan, W.; Yang, J.; et al. Molecular-Level Insight into How Hydroxyl Groups Boost Catalytic Activity in CO2 Hydrogenation into Methanol. Chemistry 2018, 4, 613–625. [Google Scholar] [CrossRef]
  28. Sha, F.; Tang, S.; Tang, C.; Feng, Z.; Wang, J.; Li, C. The role of surface hydroxyls on ZnZrOx solid solution catalyst in CO2 hydrogenation to methanol. Chin. J. Catal. 2023, 45, 162–173. [Google Scholar] [CrossRef]
  29. Jadhav, S.G.; Vaidya, P.D.; Bhanage, B.M.; Joshi, J.B. Catalytic carbon dioxide hydrogenation to methanol: A review of recent studies. Chem. Eng. Res. Des. 2014, 92, 2557–2567. [Google Scholar] [CrossRef]
  30. Zhang, R.; Wang, B.; Liu, H.; Ling, L. Effect of Surface Hydroxyls on CO2 Hydrogenation Over Cu/γ-Al2O3 Catalyst: A Theoretical Study. J. Phys. Chem. C 2011, 115, 19811–19818. [Google Scholar] [CrossRef]
  31. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  32. Kresse, G.; Furthmuller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  33. Blochl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [PubMed]
  34. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef]
  35. Nosé, S. A unified formulation of the constant temperature molecular dynamics method. J. Chem. Phys. 1984, 81, 511–519. [Google Scholar] [CrossRef]
  36. Henkelman, G.; Uberuaga, B.P.; Jónsson, H. A climbing image nudged elastic band method for finding saddle points and minimum energy paths. J. Chem. Phys. 2000, 113, 9901–9904. [Google Scholar] [CrossRef]
  37. Henkelman, G.; Jónsson, H. Improved tangent estimate in the nudged elastic band method for finding minimum energy paths and saddle points. J. Chem. Phys. 2000, 113, 9978–9985. [Google Scholar] [CrossRef]
  38. Campbell, C.T. Future directions and industrial perspectives micro- and macro-kinetics: Their relationship in heterogeneous catalysis. Top. Catal. 1994, 1, 353–366. [Google Scholar] [CrossRef]
  39. Campbell, C.T. Finding the rate-determining step in a mechanism: Comparing DeDonder relations with the “degree of rate control”. J. Catal. 2001, 204, 520–524. [Google Scholar] [CrossRef]
  40. Filot, I.A.; Broos, R.J.; van Rijn, J.P.; van Heugten, G.J.; van Santen, R.A.; Hensen, E.J. First-Principles-Based Microkinetics Simulations of Synthesis Gas Conversion on a Stepped Rhodium Surface. ACS Catal. 2015, 5, 5453–5467. [Google Scholar] [CrossRef]
  41. Jennison, D.R.; Schultz, P.A.; Sullivan, J.P. Evidence for Interstitial Hydrogen as the Dominant Electronic Defect in Nanometer Alumina Films. Phys. Rev. B 2004, 69, 041405. [Google Scholar] [CrossRef]
  42. Digne, M.; Sautet, P.; Raybaud, P.; Euzen, P.; Toulhoat, H. Use of DFT to Achieve a Rational Understanding of Acid-Basic Properties of Gamma-Alumina Surfaces. J. Catal. 2004, 226, 54–68. [Google Scholar] [CrossRef]
  43. Digne, M.; Raybaud, P.; Sautet, P.; Guillaume, D.; Toulhoat, H. Atomic Scale Insights on Chlorinated Gamma-Alumina Surfaces. J. Am. Chem. Soc. 2008, 130, 11030–11039. [Google Scholar] [CrossRef] [PubMed]
  44. Fu, L.; Yang, H. Tailoring the Electronic Structure of Mesoporous Spinel γ-Al2O3 at Atomic Level: Cu-Doped Case. J. Phys. Chem. C 2014, 118, 14299–14315. [Google Scholar] [CrossRef]
  45. Ferreira, A.R.; Martins, M.J.F.; Konstantinova, E.; Capaz, R.B.; Souza, W.F.; Chiaro, S.S.X.; Leitão, A.A. Direct Comparison between Two Structural Models by DFT Calculations. J. Solid State Chem. 2011, 184, 1105–1111. [Google Scholar] [CrossRef]
  46. Mchale, J.M.; Navrotsky, A.; Perrotta, A.J. Effects of Increased Surface Area and Chemisorbed H2O on the Relative Stability of Nanocrystalline γ-Al2O3 and α-Al2O3. J. Phys. Chem. B 1997, 101, 603–613. [Google Scholar] [CrossRef]
  47. Zuo, Z.J.; Han, P.D.; Hu, J.S.; Huang, W. Effect of Surface Hydroxyls on DME and Methanol Adsorption over γ-Al2O3 (hkl) Surfaces and Solvent Effects: A Density Functional Theory Study. J. Mol. Model 2012, 18, 5107–5111. [Google Scholar] [CrossRef]
  48. Cogollo-Olivo, B.H.; Seriani, N.; Montoya, J.A. Unbiased structural search of small copper clusters within DFT. Chem. Phys. 2015, 461, 20–24. [Google Scholar] [CrossRef]
  49. Jug, K.; Zimmermann, B.; Calaminici, P.; Köster, A.M. Structure and stability of small copper clusters. J. Chem. Phys. 2002, 116, 4497–4507. [Google Scholar] [CrossRef]
  50. Fernández, E.M.; Soler, J.M.; Garzón, I.L.; Balbás, L.C. Trends in the structure and bonding of noble metal clusters. Phys. Rev. B 2004, 70, 165403. [Google Scholar] [CrossRef]
  51. Hernández, N.C.; Graciani, J.; Márquez, A.; Sanz, J.F. Cu, Ag and Au Atoms Deposited on the α-Al2O3(0001) Surface: A Comparative Density Functional Study. Surf. Sci. 2005, 575, 189–196. [Google Scholar] [CrossRef]
  52. Li, J.; Zhang, R.; Wang, B. Influence of the Hydroxylation of γ-Al2O3 Surfaces on the Stability and Growth of Cu for Cu/γ-Al2O3 Catalyst: A DFT Study. Appl. Surf. Sci. 2013, 270, 728–736. [Google Scholar] [CrossRef]
Figure 1. (a) Top and (b) side views of the optimized structures of clean γ-Al2O3(110) surface with nonspinel phase. The red and purple balls represent O and Al atoms, respectively. The structure of the unit cell of bulk γ-Al2O3 is shown in inset.
Figure 1. (a) Top and (b) side views of the optimized structures of clean γ-Al2O3(110) surface with nonspinel phase. The red and purple balls represent O and Al atoms, respectively. The structure of the unit cell of bulk γ-Al2O3 is shown in inset.
Catalysts 13 01244 g001
Figure 2. Side (upper panel) and top (bottom panel) views of (a) the optimized structure of the γ-Al2O3(110) surface with a hydroxyl coverage of 5.9 OH·nm−2, the most stable configuration of (b) Cu4 cluster and (c) CO2 molecule adsorbed on the hydroxylated γ-Al2O3(110) surface. Distances are in Å. The sienna, red, grey and purple spheres represent Cu, O, C and Al atoms, respectively. For clarity, the symbols Oa and Ob were used to denote two oxygen atoms of CO2 moiety.
Figure 2. Side (upper panel) and top (bottom panel) views of (a) the optimized structure of the γ-Al2O3(110) surface with a hydroxyl coverage of 5.9 OH·nm−2, the most stable configuration of (b) Cu4 cluster and (c) CO2 molecule adsorbed on the hydroxylated γ-Al2O3(110) surface. Distances are in Å. The sienna, red, grey and purple spheres represent Cu, O, C and Al atoms, respectively. For clarity, the symbols Oa and Ob were used to denote two oxygen atoms of CO2 moiety.
Catalysts 13 01244 g002
Figure 3. Potential energy profiles for the dissociation of H2O at the (a) Al3c site and (b) Al4c-I site to form the hydroxylated γ-Al2O3(110) surface. The red, purple and white spheres represent O, Al and H atoms, respectively.
Figure 3. Potential energy profiles for the dissociation of H2O at the (a) Al3c site and (b) Al4c-I site to form the hydroxylated γ-Al2O3(110) surface. The red, purple and white spheres represent O, Al and H atoms, respectively.
Catalysts 13 01244 g003
Figure 4. The optimized geometries of Cun clusters in gas phase. The Cu–Cu bond lengths (in Å) are shown for each cluster.
Figure 4. The optimized geometries of Cun clusters in gas phase. The Cu–Cu bond lengths (in Å) are shown for each cluster.
Catalysts 13 01244 g004
Figure 5. Side (upper panel) and top (bottom panel) views of the most stable structure of Cun clusters supported on a dry γ-Al2O3(110) surface. The sienna, red and purple balls represent Cu, O and Al atoms, respectively.
Figure 5. Side (upper panel) and top (bottom panel) views of the most stable structure of Cun clusters supported on a dry γ-Al2O3(110) surface. The sienna, red and purple balls represent Cu, O and Al atoms, respectively.
Catalysts 13 01244 g005
Figure 6. Side (upper panel) and top (bottom panel) views of the optimized structures of CO2 adsorption on Cun/γ-Al2O3 (110) surface. The sienna, red, grey and purple balls represent Cu, O, C and Al atoms, respectively. For clarity, the symbols Oa and Ob were used to distinguish two oxygen atoms of CO2 moiety.
Figure 6. Side (upper panel) and top (bottom panel) views of the optimized structures of CO2 adsorption on Cun/γ-Al2O3 (110) surface. The sienna, red, grey and purple balls represent Cu, O, C and Al atoms, respectively. For clarity, the symbols Oa and Ob were used to distinguish two oxygen atoms of CO2 moiety.
Catalysts 13 01244 g006
Figure 7. Potential energy diagrams for CH3OH synthesis via the formate pathway (black line) and the RWGS + CO hydrogenation pathway (blue line) on (a) dry and (b) hydroxylated Cu4/γ-Al2O3 (110) surfaces.
Figure 7. Potential energy diagrams for CH3OH synthesis via the formate pathway (black line) and the RWGS + CO hydrogenation pathway (blue line) on (a) dry and (b) hydroxylated Cu4/γ-Al2O3 (110) surfaces.
Catalysts 13 01244 g007
Figure 8. Degree of rate control (DRC) of methanol synthesis as a function of temperatures via (a) the formate pathway and (b) RGWS + CO hydrogenation pathway on the dry surface with a mixture of CO2 and H2 feed gases in a 1:3 ratio at 1 atm total pressure.
Figure 8. Degree of rate control (DRC) of methanol synthesis as a function of temperatures via (a) the formate pathway and (b) RGWS + CO hydrogenation pathway on the dry surface with a mixture of CO2 and H2 feed gases in a 1:3 ratio at 1 atm total pressure.
Catalysts 13 01244 g008
Table 1. Calculated binding energy (Eb) (in eV), average lengths of Cu–O, Cu–Cu and Cu–Cu bonds (in Å), the Bader charge of Cun clusters (in e) and the work function change (in eV) of Cun clusters supported on the dry surface, as well as the corresponding results of Cu4 deposited on the hydroxylated γ-Al2O3(110) surface.
Table 1. Calculated binding energy (Eb) (in eV), average lengths of Cu–O, Cu–Cu and Cu–Cu bonds (in Å), the Bader charge of Cun clusters (in e) and the work function change (in eV) of Cun clusters supported on the dry surface, as well as the corresponding results of Cu4 deposited on the hydroxylated γ-Al2O3(110) surface.
SystemsEbdCuOdCuAldCuCuBader Charge of
Cun Cluster a
Work Function Change b
Cu12.511.810+0.54−0.62
Cu23.402.0152.5422.309+0.20−0.71
Cu34.442.1492.6202.365+0.68−2.08
Cu44.552.1052.4742.359+0.35−2.38
Cu54.742.0542.5622.478+0.60−2.04
Cu65.302.0742.4492.373+0.26−2.16
Cu4 (hydroxylated surface)7.082.1132.6102.433+0.80−1.59
a The positive value denotes the loss of electrons of copper clusters after deposition. b The negative value means the decrease in surface work function with respect to the pristine surface.
Table 2. Calculated adsorption energy (Eads) (in eV), the lengths of two C–O bonds (in Å), the ∠O-C-O bond angle (in degree), the lengths of Al–Ob and Cu–C adsorption bonds (in Å), Bader charges (in e) of Cun cluster and CO2 moiety after the adsorption of CO2 on the Cun/γ-Al2O3 (110) a.
Table 2. Calculated adsorption energy (Eads) (in eV), the lengths of two C–O bonds (in Å), the ∠O-C-O bond angle (in degree), the lengths of Al–Ob and Cu–C adsorption bonds (in Å), Bader charges (in e) of Cun cluster and CO2 moiety after the adsorption of CO2 on the Cun/γ-Al2O3 (110) a.
SystemsEadsC–OaC–Ob∠O–C–OAl4c–ObCu–CCharge of
Cun Cluster
Charge of CO2 Moiety
Pristine surface0.451.1621.191176.52.063
Cu10.651.2181.288133.71.8731.914+0.69−0.71
Cu20.771.2261.312128.51.8672.056 (ave.)+0.85−0.93
Cu30.641.2151.311130.11.8902.076 (ave.)+1.24−0.79
Cu40.861.2281.314127.81.8542.121 (ave.)+0.82−0.97
Cu50.831.2161.283133.91.8871.947+1.07−0.72
Cu60.741.2311.312128.41.8392.112 (ave.)+1.00−0.95
Cu4 (hydroxylated surface)1.061.2891.304122.31.8512.042 (ave.)+1.39−1.10
a For the free CO2 molecule, the optimized length of C–O bond was 1.176 Å. The positive and negative charges denote the loss and gain of electrons, respectively.
Table 3. Calculated reaction energy (Er) and the activation barrier (Ea) of each elementary reaction involved in the formate pathway (F1 to F6) and the RWGS + CO hydrogenation pathway (R1 to R4) for methanol synthesis on the dry Cu4/γ-Al2O3(110) surface.
Table 3. Calculated reaction energy (Er) and the activation barrier (Ea) of each elementary reaction involved in the formate pathway (F1 to F6) and the RWGS + CO hydrogenation pathway (R1 to R4) for methanol synthesis on the dry Cu4/γ-Al2O3(110) surface.
StepElementary ReactionEr (eV)Ea (eV)
F1CO2* + H* → HCOO* + * a −1.290.28 (TS1)
F2HCOO* + H* → H2COO* + *+0.711.18 (TS2)
F3H2COO* + H* → H2CO* + OH*−0.580.58 (TS3)
F4H2CO* + OH* + H*→ H2CO* + H2O* + *−0.500.82 (TS4)
F5H2CO* + H* → H3CO* + *−1.250.13 (TS5)
F6H3CO* + H* → H3COH* + * −0.090.68 (TS6)
R1CO2* + H* → COOH* + *−0.460.78 (TS1′)
R2COOH* + H* → CO* + H2O(g) + *+0.791.55 (TS2′)
R3CO* + H* → HCO* + *−0.100.88 (TS3′)
R4HCO* + H* → H2CO* + *−0.060.44 (TS4′)
a Symbol * represents the adsorption site on the surface.
Table 4. Calculated reaction energy (Er) and the activation barrier (Ea) of each elementary reaction involved in the formate pathway (F1 to F6) and the RWGS + CO hydrogenation pathway (R1 to R4) for methanol synthesis on the hydroxylated Cu4/γ-Al2O3(110) surface a.
Table 4. Calculated reaction energy (Er) and the activation barrier (Ea) of each elementary reaction involved in the formate pathway (F1 to F6) and the RWGS + CO hydrogenation pathway (R1 to R4) for methanol synthesis on the hydroxylated Cu4/γ-Al2O3(110) surface a.
StepElementary ReactionEr (eV)Ea (eV)
F1CO2* + H* → HCOO* + *−1.500.15 (TS1)
F2HCOO* + H* → H2COO* + *+0.411.34 (TS2)
F3H2COO*+ Hs*→ H2COOH* + *−0.560.02 (TS3)
F4H2COOH* + Hs*→ H2CO* + H2O*+0.570.79 (TS4)
F5H2CO* + H* → H3CO* + *−0.580.17 (TS5)
F6H3CO* + H* → H3COH* + * +0.250.73 (TS6)
R1CO2* + H* → COOH* + *−0.820.73 (TS1′)
R2COOH* + H* → CO* + H2O(g) + *+1.221.91 (TS2′)
R3CO* + H* → HCO* + *−0.430.56 (TS3′)
R4HCO* + H* → H2CO* + *−0.370.65 (TS4′)
a The hydrogen atom of OH group is denoted by Hs.
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

Zhou, H.; Jin, H.; Li, Y.; Li, Y.; Huang, S.; Lin, W.; Chen, W.; Zhang, Y. Mechanism of Methanol Synthesis from CO2 Hydrogenation over Cu/γ-Al2O3 Interface: Influences of Surface Hydroxylation. Catalysts 2023, 13, 1244. https://doi.org/10.3390/catal13091244

AMA Style

Zhou H, Jin H, Li Y, Li Y, Huang S, Lin W, Chen W, Zhang Y. Mechanism of Methanol Synthesis from CO2 Hydrogenation over Cu/γ-Al2O3 Interface: Influences of Surface Hydroxylation. Catalysts. 2023; 13(9):1244. https://doi.org/10.3390/catal13091244

Chicago/Turabian Style

Zhou, Hegen, Hua Jin, Yanli Li, Yi Li, Shuping Huang, Wei Lin, Wenkai Chen, and Yongfan Zhang. 2023. "Mechanism of Methanol Synthesis from CO2 Hydrogenation over Cu/γ-Al2O3 Interface: Influences of Surface Hydroxylation" Catalysts 13, no. 9: 1244. https://doi.org/10.3390/catal13091244

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