A combined theoretical and experimental investigation on the photocatalytic hydrogenation of CO2 on Cu/ZnO polar surface

Han Xiao ab, Yihong Lian a, Shiduo Zhang a, Minyi Zhang *ab, Jiye Zhang *c and Chunsen Li *abde
aState Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou, Fujian 350002, China. E-mail: myzhang@fjirsm.ac.cn; Chunsen.li@fjirsm.ac.cn
bUniversity of Chinese Academy of Sciences, Beijing 100049, China
cSchool of Materials Science and Engineering, Shanghai University, Shanghai 200444, China. E-mail: jychang@shu.edu.cn
dFujian Provincial Key Laboratory of Theoretical and Computational Chemistry, Xiamen University, Xiamen, Fujian 361005, China
eFujian College, University of Chinese Academy of Sciences, Fuzhou, Fujian 350002, China

Received 3rd March 2023 , Accepted 21st April 2023

First published on 21st April 2023


Abstract

The photocatalytic hydrogenation of CO2 by Cu-deposited ZnO (Cu/ZnO) polar surfaces is investigated through density functional theory (DFT) calculations combined with experimental work. The DFT results demonstrate that, without Cu-loading, CO2 and H2 present weak physisorption on the clean ZnO polar surface, except that H2 undergoes strong chemisorption on the ZnO(000[1 with combining macron]) surface. Cu deposition on the ZnO polar surface could remarkably enhance the CO2 chemisorption ability, due to the induced charge redistribution on the interface of the Cu/ZnO polar surface systems. Additionally, a Cu-nanoisland, which was simulated using a Cu(111) slab model, exhibited strong ability to chemically adsorb H2. Thus, H2 may act as an adsorption competitor to CO2 on the Cu/ZnO(000[1 with combining macron]), while, in contrast, CO2 and H2 (syngas) may have more opportunity to simultaneously adsorb on Cu/ZnO(0001) to promote the CO2 hydrogenation. These facet-dependent properties lead us to assume that Cu/ZnO(0001) should be a favorable photocatalyst for CO2 hydrogenation. This assumption is further verified by our photocatalysis experiment based on a ZnO single crystal. According to the theoretical and experimental results, the optimal HCOO* reaction pathway for the photocatalytic hydrogenation of CO2 on Cu/ZnO(0001) is proposed. In this optimal HCOO* path, the hydrogenation of CO2* step and hydrogenation of HCOO* step could be promoted by the coupling of a photo-generated spillover proton and a photoelectron on the interface of Cu/ZnO(0001). This research demonstrates the feasibility of the photocatalytic reduction of CO2 on Cu/ZnO(0001), and will help to develop related high-efficiency catalysts.


1. Introduction

From the perspective of reducing the concentration of carbon dioxide in the atmosphere to combat the greenhouse effect and global warming, the conversion of carbon dioxide into chemical feedstocks such as formic acid and methanol through catalytic hydrogenation is an important strategy.1–4 Solid catalysts such as the transition-metal-doped zinc oxide (ZnO) exhibit high activity and selectivity for the reduction of CO2, and have been considered as key chemical platforms for the production of fuels and feedstocks at pressures less than 100 bar and temperatures ranging from 473 to 573 K.5–7 However, compared with conventional catalytic strategies, photocatalytic CO2 hydrogenation is now becoming a new promising green chemistry technology that can directly convert CO2 into useful chemical products under ultraviolet and visible light irradiation without high temperature and high pressure conditions.8–10 ZnO has been extensively investigated as a photocatalyst for solar energy conversion and organic or dye pollutant treatment due to its excellent electrochemical stability, high electron mobility, and photocatalytic efficiency.11–14 The polar surfaces of ZnO, such as the Zn-terminated surface ZnO(0001) and O-terminated surface ZnO(000[1 with combining macron]), were found to be more active in photocatalytic reactions than the nonpolar surfaces of ZnO,15–17 because the dipoles associated with Zn- and O-terminated surfaces cause electrostatic instability on these surfaces. As such, much research work has been devoted to development of efficient photocatalysts for CO2 reduction using ZnO as an important composite.18–21

On the other hand, many studies have also focused on Cu-decorated ZnO, which generally exhibits reduced optical band-gap energy, increased optical absorption ability, and improved photocatalytic properties compared with bare ZnO.22–25 Pawar et al. recently reported that Cu-doped ZnO microstructures showed enhanced performance in the photocatalytic degradation of methylene blue and rhodamine B dyes due to the existence of a large amount of (001) polar surfaces.26 Moreover, Tsang et al. found that the exposed polar face (002) of Cu-deposited ZnO (Cu/ZnO) showed higher selectivity towards the hydrogenation of CO2 to methanol compared with other crystal facets,27 and thus demonstrated the important role of facet type in regulating catalytic reactivity.28 This evokes speculation that the ZnO polar surface in synergy with copper might also exhibit good photocatalytic CO2 hydrogenation ability.

It has been found that the interaction of the deposited Cu and ZnO has a significant impact on the activation and hydrogenation of CO2.5,6,29 However, little is known about the structure–activity relationship between deposited Cu and the polar surfaces of ZnO, such as the active sites, and the mechanism of CO2 activation on polar surfaces. Moreover, no consensus has been reached on the adsorption behavior of CO2 on ZnO(0001) and ZnO(000[1 with combining macron]) polar surfaces in previous experimental and theoretical work.26,27,30 Therefore, a detailed study of how CO2 interacts with clean ZnO polar surfaces as well as Cu/ZnO polar surfaces will help to provide an understanding of the intrinsic mechanism of the photocatalytic hydrogenation of CO2 by Cu/ZnO photocatalyst.

In the present paper, our study focuses on exploring the mechanism of the photocatalytic hydrogenation of CO2 on the Cu-deposited ZnO polar surfaces, ZnO(0001) and ZnO(000[1 with combining macron]). By performing theoretical simulations, we investigated the adsorption behaviors of CO2 and H2 on the polar surface of clean and Cu-deposited ZnO polar surfaces, respectively. The potential active polar surface, Cu/ZnO(0001), is proposed based on the elucidation of adsorption sites, relaxation geometries, and electrical properties, as we shall show. Then, DFT calculations were performed to investigate the reaction mechanism of CO2 hydrogenation on the Cu/ZnO(0001) surface and the favorable reaction pathway in terms of energy was identified. Furthermore, illustrative experiments were done accordingly, and the results were in good agreement with the theoretical predictions. Finally, based on the combined theoretical and experimental works, we propose a possible reaction route for the direct photocatalytic hydrogenation of CO2 on the Cu/ZnO surface.

2. Theoretical and experimental methods

2.1. Computational methods and modeling

All calculations were performed using the first-principles method based on density functional theory (DFT) and the first-principles pseudo-potential method to simulate the surface reconstruction and formation energy. The DFT calculations were carried out using the Vienna Ab Initio Simulation Package (VASP) code31,32 with the projector-augmented wave (PAW) method using the Perdew–Burke–Ernzerhof (PBE) exchange–correlation functional33,34 The GGA+U approximation with an effective Hubbard U-parameter Ueff = 7.65 eV was applied to describe the on-site Coulomb interaction among the localized zinc 3d electrons.35 The value of Ueff is close to 7.5 eV, which was used in a similar Cu/ZnO system.36 The plane wave cutoff was set as 400 eV, and the total energy convergence was set at 10−6 eV for self-consistent iterations. As shown in Fig. 1, the theoretical model used here consists of an orthogonal surface unit cell a × b (i.e., 6.506 × 5.643 Å) containing four zinc and four oxygen atoms within each Zn–O layer in a repeated slab configuration. Each slab contains 6 Zn–O layers separated by a vacuum space of 20 Å. The surface sizes are (1a × 2b) corresponding to k-point meshes of (4 × 2 × 1). In order to suppress any long-range interference between the zinc and oxygen-terminated ends of the slab, the dangling bonds of the bottom atoms of ZnO (0001) and (000[1 with combining macron]) were saturated by adding one monolayer of pseudo-hydrogen atoms with a nuclear charge of Z = 1/2|e| and Z = 3/2|e|, respectively. Only the top three Zn–O layers were relaxed, while the bottom three Zn–O layers were kept in their previously optimized crystal positions.
image file: d3nr01001e-f1.tif
Fig. 1 (a) Side view of the slab supercell for the clean ZnO(0001) surface. (b) Top view of the slab supercell for the ZnO(0001) surface and the related CO2/H2 adsorption sites; top view of the slab supercell for the ZnO (0001) surface. (c) Top view of the slab supercell for the Cu/ZnO(0001) surface and the related CO2/H2 adsorption sites. (d) Side view and (e) top view of the slab supercell for the Cu9/ZnO(0001) surface. Note: The geometry of ZnO (000[1 with combining macron]) surface can be simply imaged by O atom and Zn atom interchange.

For the adsorption of Cu atoms and adsorbate molecules (CO2 or H2) on a clean ZnO polar surface, we simulated four symmetry adsorption sites, namely a top-site, bridge-site, fcc-site, and hcp-site (Fig. 1(b)). For the adsorption of CO2 or H2 on the Cu/ZnO system, four symmetry adsorption sites are also considered. The 1st–3rd positions are the edge site, the site close to the Cu atom, and the site far from the Cu atom, respectively (see Fig. 1(c)). The 4th adsorption site is a metallic site located far from the ZnO support. For simplified calculation, an ideal Cu(111) surface with a six-layer slab and a p(4 × 4) unit cell was used. In this model, only the top three layers were relaxed, and k-point meshes of 4 × 4 × 1 were used. The energy of adsorption/binding Ead is defined as:

Ead = Eads/subEadsEsub
where Eads/sub is the total energy of the substrate covered with the adsorbate in the optimized configuration, and Eads and Esub are the total energies of the adsorbate in the gas phase and the relaxed clean substrate, respectively.

Moreover, we constructed the Cu9 cluster loading on the ZnO(0001) surface with a size of (4a × 4b) and (2 × 2 × 1) k point meshes to explore the CO2 hydrogenation pathways (Fig. 1(d) and (e)). In this model, only the top two Zn–O layers were relaxed, while the bottom four Zn–O layers were kept on their previously optimized crystal positions. The starting geometry of the Cu9/ZnO(0001) surface was first determined by running ab initio molecular dynamics (AIMD) using a Nose thermostat at a temperature of 300 K for 12 ps. The subsequent energy minimized snapshots extracted from the AIMD trajectories were then singled out and optimized.

2.2. Synthesis of Cu/ZnO polar surface photocatalysts and related characterization

Two pieces of (0001) facet oriented ZnO wafers with a size of 10 × 10 × 0.5 mm were cut from a bulk ZnO single crystal grown via a hydrothermal method37 and then polished and cleaned. The determination of polarized facets was conducted using a simple chemical etching method.38 Cu nanoislands were deposited on the ZnO wafers using a radiofrequency magnetron sputtering system. The diameter and thickness of the Cu target were 50 mm and 3 mm, respectively, and the purity was greater than 99.99%. The working pressure (Ar 99.99%) was 0.1 Pa and the substrate was maintained at 473 K during the film deposition. The distance from the target to the substrate was 9.0 cm. The sputtering time was about 20 s. One of the wafers was deposited with Cu nanoislands on its (001) facet using the magnetron sputtering method, thus forming the Cu–ZnO(0001) catalyst. The other wafer was deposited with Cu nanoislands on its (000[1 with combining macron]) facet under the same deposition conditions, thus forming the Cu–ZnO(000[1 with combining macron]) catalyst. The morphologies and distribution of Cu metal on the ZnO substrates were characterized using atomic force microscopy (AFM; Veeco Multimode, Nano scope Multimode IIIa). X-ray photoelectron spectroscopy (XPS) analysis was conducted on an ESCA-LAB 250 photoelectron spectrometer (Thermo Scientific) using an Al Kαradiation (1486.6 eV). Height images of the films were recorded using tapping-mode AFM. The photocatalytic hydrogenation of CO2 experiment was carried out in a flowing atmosphere and at room temperature using a custom-built reactor (see details in the ESI). A CO2/H2 reaction mixture with a molar ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]3 was fed at a rate of 300 stp mL min−1 (stp = standard temperature and pressure, T = 298 K, P = 101.3 kPa) through the catalyst bed. Before each test, the catalysts were pre-reduced in situ at 573 K for 2 h under the H2 flow (20 stp mL min−1). The catalysts were illuminated using a 300 W Xe light source (Solaredge700, Beijing Perfect light Co. Ltd) for 4 h. The possible residual products on the surface of the catalysts were determined using a PerkinElmer system 2000 FT-IR spectrometer. A resolution of 4.0 cm−1 was used throughout the investigation and 50 scans taken over a 20 s interval were averaged to achieve a satisfactory signal-to-noise ratio.

3. Results and discussion

3.1. Geometries of Cu/ZnO polar surfaces and their adsorption properties toward CO2/H2

The models of the clean ZnO(0001)/(000[1 with combining macron]) surface were simulated using slab models based on optimized bulk wurtzite ZnO using an orthogonal unit cell.39 The performances of ZnO polar surfaces with copper deposition were simplistically simulated using one Cu atom deposited on the clean ZnO(0001) and ZnO(000[1 with combining macron]) polar surfaces, respectively (see Fig. 2(a) and (b)). The calculated binding energies of the four high-symmetry deposition sites (i.e., top, bridge, fcc-hollow, and hcp-hollow, as shown in Fig. 1(b)) are displayed in Table S1 (see ESI). Our results found that both the ZnO(0001) surface and ZnO(000[1 with combining macron]) surface could strongly bind to the Cu atom. For both the ZnO(0001) surface or ZnO(000[1 with combining macron]) surface, the fcc-hollow site was the most stable deposition site for Cu.
image file: d3nr01001e-f2.tif
Fig. 2 The most stable geometry of Cu deposited at the fcc-site for (a) the clean ZnO(0001) and (b) ZnO(000[1 with combining macron]) surfaces. The most stable geometry of (c) CO2 and (d) H2 adsorbed on the clean surface of ZnO(0001); and the most stable geometry of (e) CO2 and (f) H2 adsorbed on the clean surface of ZnO(000[1 with combining macron]). The most stable geometry of (g) CO2 and (h) H2 adsorbed on Cu/ZnO(0001); the most stable geometry of (i) CO2 and (j) H2 adsorbed on the Cu/ZnO(000[1 with combining macron]) system; the most stable geometry of (k) CO2 and (l) H2 adsorbed on metallic Cu(111).

Further, we investigated CO2 and H2 adsorption on the clean and Cu-deposited ZnO polar surfaces (see Fig. S1 and S2 in ESI). Based on our calculations, only weak physisorption of CO2 was found on the two clean polar surfaces (see Table S1 and the stable adsorption states of CO2 on the two clean surfaces, Fig. 2(c) and (e)), which is in line with previous theoretical studies.40 In contrast, H2 adsorption on the ZnO polar surfaces showed facet-dependent properties. For the sites of ZnO(0001), H2 adsorption only resulted from the physisorption of H2. However, for the ZnO(000[1 with combining macron]) surface, H2 adsorption on the bridge-site and fcc-site could lead to strong chemisorption of H2. In these cases, the H–H bond of H2 is broken, and two dissociated H atoms are separately bound to two adjacent surface O atoms of the ZnO(000[1 with combining macron]) surface.

The characteristics of CO2 and H2 adsorption on the Cu/ZnO polar surfaces were investigated by using the most stable Cu/ZnO polar surface with Cu deposited on the fcc-hollow site of the ZnO polar surface (see Fig. S2 in ESI). The calculated adsorption energies with different deposition sites are listed in Table 1. In addition, in order to simulate the top of the deposited Cu nano-islands on the ZnO polar surfaces in the experiments, both the metallic sites of Cu/ZnO(0001) and Cu/ZnO(000[1 with combining macron]) were simplified and simulated using the Cu(111) slab models (see Fig. 2(k) and (l)).

Table 1 Adsorption energies (in eV) for CO2 and H2 adsorbed on the Cu/ZnO(0001) and Cu/ZnO(000[1 with combining macron]) system
Cu/ZnO system Edge Close Far Metallic
(0001)/CO2 −0.89 −0.06 −0.05 −0.03
(0001)/H2 −0.05 −0.02 −0.02 −0.41
(000[1 with combining macron])/CO2 −0.66 −0.58 −0.08 −0.03
(000[1 with combining macron])/H2 −3.57 −4.45 −4.67 −0.41


For the Cu/ZnO(0001) surface, CO2 adsorption on the edge site has the lowest adsorption energy of −0.89 eV (see Table 1 and Fig. 2(g)). In this case, the linear CO2 is bent to a bidentate carbonate with an internal angle O–C–O of 126.8°. Its C atom bonds to the Cu atom with C–Cu bond length of 1.98 Å, and one of its O atoms bonds to a surface Zn atom with an O–Zn bond length of 1.96 Å (see Fig. 2(g)). In the case of the Cu/ZnO(000[1 with combining macron]) surface, CO2 adsorption on the edge-site of Cu/ZnO(000[1 with combining macron]) also leads to the lowest adsorption energy of −0.66 eV (see Table 1 and Fig. 2(i)). In this structure, the C atom of CO2 binds with the O-terminated facet with a C–O bond length of 1.41 Å, and one of its O atoms bonds to the Cu atom with the Cu–O bond length of 1.88 Å. The internal O–C–O angle of CO2 is bent to 126.3°. Thus, Cu deposition on ZnO polar surfaces would highly activate the CO2 and promote the chemisorption of CO2. Moreover, the chemisorption of CO2 occurs only in close proximity to the Cu deposition sites of Cu/ZnO(0001) and Cu/ZnO(000[1 with combining macron]) surfaces. For the far-sites and metallic sites of Cu/ZnO(0001) and Cu/ZnO(000[1 with combining macron]), the adsorption of CO2 results from only weak physisorption. This feature demonstrates that Cu deposition on ZnO polar surfaces could play an important role in the chemisorption and activation of CO2.

On the other hand, similar to the case of the clean ZnO(0001) surface, the adsorption of H2 on the Cu/ZnO(0001) surface was unfeasible (see Table 1), except for the metallic sites. H2 adsorption on the Cu(111) surface leads to strong chemisorption of H2 with an adsorption energy of −0.41 eV (see Table 1). Therefore, H2 adsorption on any of the sites of Cu/ZnO(000[1 with combining macron]) could result in the strong chemisorption of H2. Similar to the cases of H2 adsorption on the clean ZnO(000[1 with combining macron]) surface and Cu/ZnO(000[1 with combining macron]), the chemisorption of H2 on the Cu(111) surface results in the dissociation of H2. As shown in Fig. 2(l), the two dissociated H atoms are loaded on two neighbouring Cu atoms of the Cu(111) surface (dH–Cu = 1.75 Å). The dissociated H atom might further evolve to the spillover proton among the Cu nano-islands by the photo-generated holes.41

Thus, the above results show that Cu deposition on ZnO polar surfaces would promote CO2 chemisorption on the edge sites of Cu/ZnO(0001) and Cu/ZnO(000[1 with combining macron]). These chemisorbed states rarely occur on the clean ZnO polar surfaces. Moreover, Cu deposition on ZnO(0001) could also promote the chemisorption H2 by inducing the Cu nano-islands on the Cu/ZnO(0001) surface. However, Cu deposition on ZnO(000[1 with combining macron]) has a minor effect on the H2 adsorption, since H2 adsorption on any of the sites of Cu/ZnO(000[1 with combining macron]) could result in the strong chemisorption of H2. Therefore, when Cu/ZnO(0001) and Cu/ZnO(000[1 with combining macron]) are exposed to CO2 and H2, the Cu/ZnO(000[1 with combining macron]) system could be saturated and poisoned by the strong chemisorption of H2. In contrast, the adsorption competitors CO2 and H2 could be simultaneously chemically adsorbed and activated on Cu/ZnO(0001) (on which the H2 undergoes chemisorption on metallic sites and CO2 chemisorption occurs on edge sites). Thus, we propose that Cu/ZnO(0001) is an ideal surface for the hydrogenation of CO2.

3.2. Electronic structures of the chemisorbed CO2 states of Cu/ZnO polar surfaces

In this section, detailed electronic structures have been obtained to provide further understanding of the physical and chemical origins of the interaction between CO2 and the Cu/ZnO polar surfaces. The ZnO(0001) surface has n-type-like conductivity, as the Fermi level enters into its conduction bands; see the density-of-states (TDOS) (see Fig. S3(a) in ESI). The ZnO(000[1 with combining macron]) surface exhibits p-type conductive characters, as its Fermi level is lower and enters the valence bands (see Fig. S3(c) in ESI). Similar results were reported in ref. 42 and 43. With Cu deposition on the ZnO(0001) surface, the Cu 3d states contribute to the valence band maxima of the Cu/ZnO(0001) surface, leading the TDOS of the Cu/ZnO(0001) surface to shift to the higher energy level and enlarging its band gap (see Fig. S3(b) in ESI). In contrast, in the case of Cu/ZnO(000[1 with combining macron]), no energy shift of TDOS is found for the Cu/ZnO(000[1 with combining macron]) surface. However, the contribution of the Cu 3d orbitals to the valence band maxima is still remarkable in Cu/ZnO(000[1 with combining macron]) (see Fig. S3(d) in ESI). The DOS analysis of the Cu/ZnO polar surface systems reveals that Cu deposition would induce charge redistribution on the interfaces of Cu/ZnO(0001) and Cu/ZnO(000[1 with combining macron]).

In the chemisorbed CO2 states of the Cu/ZnO polar surfaces, the 2p orbitals of the O and C atoms from CO2 could introduce some states located in the fundamental gaps of these two systems (Fig. 3(a) and (b)). For the chemisorbed CO2 state of Cu/ZnO(0001), the O(CO2) 2p orbital hybridizes remarkably with the valence orbitals of Cu in the top of the valence band. The C(CO2) 2p orbitals contribute to the conduction band, which is mainly dominated by the 4s orbitals of the Zn atoms. In the case of the chemisorbed CO2 state of the Cu/ZnO(000[1 with combining macron]) surface, the O(CO2) 2p orbital hybridizes with the Cu 3d orbital in the valence band and the surface O 2p in the top of the valence band (ranging from −4.6 to 0.2 eV), while the C(CO2) 2p contributes to the conduction band in the high-energy-level region (i.e., from 5 to 6 eV). The PDOS results illustrate that there are substantial covalent interactions between CO2 and the Cu/ZnO polar surface systems, which clearly confirm the formation of chemical bonds between the adsorbed CO2 and the Cu/ZnO polar surface systems.


image file: d3nr01001e-f3.tif
Fig. 3 Partial density of states (PDOS) for (a) CO2 chemisorption on edge site of Cu/ZnO(0001) and (b) CO2 chemisorption on edge site of Cu/ZnO(000[1 with combining macron]), and contour plots of electronic charge density difference for (c) CO2 chemisorption on edge site of Cu/ZnO(000[1 with combining macron]) and (d) CO2 chemisorption on edge-site of Cu/ZnO(0001).

Analysis of the electronic density difference (see the contour diagrams in Fig. 3(c) and (d)) again confirms the covalent contribution of the interaction between CO2 and the Cu/ZnO polar surface system. For the edge sites of Cu/Zn(0001), the charge density changes mainly emerge on the 2p orbitals of the C and O of CO2, and the Cu atom, as well as the relevant surface Zn atom. Similar charge density changes have been found in the chemisorbed CO2 states of the Cu/ZnO(000[1 with combining macron]) surface, with the charge density changes mainly emerging on the 2p orbitals of C and O of CO2, the 3d orbitals of the Cu atom, and the 2p orbitals of the linked surface O. The charge density changes of the two Cu/ZnO systems indicate the covalent interaction between CO2 and the deposited Cu atom. Moreover, the Cu atom donates electrons to CO2, and the C–O double bonds in CO2 are greatly weakened and elongated. The depleted charge density of Cu and the accumulated charge density on CO2 imply that Cu deposition could result in charge redistribution and improve the electron donation ability of ZnO polar surfaces. As a result, the chemisorption of CO2 is remarkably promoted on the edge-sites of Cu/ZnO polar surfaces.

Bader charge analyses44,45 were applied to further identify the charge states of CO2 adsorption on Cu/ZnO polar surfaces (see Table 2). Without CO2 adsorption, there is significant net negative charge distributed on the Cu atom of Cu/ZnO(0001), indicating that there is electron transfer from ZnO to Cu. This is consistent with our XPS results for Cu/ZnO(0001) (see section 3.4). When CO2 is adsorbed on the Cu/ZnO(0001), CO2 gains electrons with a partially negative charge of −0.9|e|, which is donated from Cu/ZnO(0001). The Cu atom and the surface Zn atoms of Cu/ZnO(0001) inject electrons to CO2 with 0.34|e| and 0.32|e|, respectively. In the case of Cu/ZnO(000[1 with combining macron]) system, Cu is still the primary electron donor, contributing 0.10|e|; however, surface O atoms prefer to accept the negative charges of −0.03|e|, leading to much less negative partial charges of −0.04|e| in CO2. This difference in charge transfer in between Cu/ZnO(0001) and Cu/ZnO(000[1 with combining macron]) is mainly attributed to their different surface atoms, i.e., the surface Zn atoms of Cu/ZnO(0001) and surface O atoms of Cu/ZnO(000[1 with combining macron]), which are characterized by electrophobic and electrophilic properties, respectively, as mentioned in ref. 46.

Table 2 Bader charges Q (in |e|) of the atoms (or entities) on the clean substrate (i.e., Cu/ZnO) and the adsorbed complex (i.e., CO2–Cu@ZnO), and their Bader charge differences ΔQ when going from CO2 in the gas phase plus the clean Cu@ZnO substrate to the adsorbed complexa
  CO2 chemisorbed on Cu/ZnO(0001) CO2 chemisorbed on Cu/ZnO(000[1 with combining macron])
Cu Znb Surface Znc CO2 Cu Ob Surface Oc CO2
a Negative and positive values of ΔQ indicate electron gain and donation, respectively. b For the surface atom involved in the interaction between the adsorbate and substrate. c Sum of the charges of all surface atoms in a unit cell except the atom bonding to the CO2.
Q Cu/ZnO −0.31 0.89 6.98 0.00 0.92 −1.07 −7.49 0.00
Q CO2−Cu/ZnO 0.03 1.21 7.08 −0.90 1.02 −1.11 −7.52 −0.04
ΔQ 0.34 0.32 0.10 −0.90 0.10 −0.04 −0.03 −0.04


As a result, Cu deposition on the ZnO polar surface systems would encourage CO2 activation through its electron donor ability. Moreover, Cu/ZnO(0001) possesses surface Zn atoms with electrophobic properties, which would provide more electron transfer to CO2 chemisorption than Cu/ZnO(000[1 with combining macron]). Thus, Cu/ZnO(0001) could induce more electron transfer to promote the formation of chemisorbed CO2 species on its edge site than Cu/ZnO(000[1 with combining macron]).

3.3. Reaction pathway of CO2 hydrogenation to formic acid

Based on our previous studies in the above section, Cu/ZnO(0001) is the most feasible candidate for the photocatalytic hydrogenation of CO2. In this section, the reaction pathway of CO2 hydrogenation by Cu/ZnO(0001) is investigated. Considering that the H2 molecule tends to dissociate on Cu nanoislands, we constructed a large-size slab model of Cu9/ZnO(0001) with a Cu9 cluster on the ZnO(0001) surface to simulate the CO2 hydrogenation reaction (see Fig. 1(d) and (c)). The electron location function (ELF) and charge difference distribution calculation of Cu9/ZnO(0001) reveal that the Cu9 cluster and ZnO(0001) surface have a strong interaction (Fig. 4(a) and (b)). Much of the electronic charge could be accumulated around the interface of Cu9 and ZnO(0001), as well as the photogenerated electrons arising from ZnO(0001). This great accumulated charge ability might enhance the reactivity of Cu9/ZnO(0001) toward the CO2 reduction.
image file: d3nr01001e-f4.tif
Fig. 4 (a) The electron location function and (b) the differential charge density of the Cu-cluster-loaded ZnO(0001) surface (blue represents positive charge and yellow represents negative charge). (c) The Gibbs free energy diagrams of CO2 reduction.

Fig. 4(c) shows the Gibbs energy profile and intermediate structures of CO2 hydrogenation by Cu9/ZnO(0001). The initial step corresponds to the chemisorption of CO2 on the interface of Cu9/ZnO(0001). It is energetically favourable to form the exothermic intermediate CO2* with an energy of −0.3 eV. Here, molecules that have been activated are denoted with a star. Subsequently, the hydrogenation of CO2* step could generate two different intermediates, HOCO* and HCOO*, due to the fact that the target atom attacked by H can be either the O or C atom of CO2*. Our calculation results reveal that the formation of both HCOO* and HOCO* are endothermic processes. However, the formation of the intermediate HCOO* requires a smaller activation energy than that of HOCO*. Moreover, since the C atom of CO2* is combined with the interfacial Cu atom, the C atom could be activated by the greatly accumulated charge around the interface of Cu9/ZnO(0001). Particularly, in the HCOO* path, the spillover proton on the Cu nano-islands could couple with the photoelectron to attack the C atom of CO2*, which could greatly promote the formation of the intermediate HCOO*. The intermediate HCOO* is metastable and, subsequently, HCOO* would undergo structural rearrangement to form the bi-HCOO* intermediate with an energetic exothermicity of 0.55 eV. In this structural rearrangement step, the upper O atom of HCOO* would shift to the interface of Cu9/ZnO(0001) and combine with the interfacial Cu and Zn atoms to form the intermediate bi-HCOO*. In our calculations, bi-HCOO* should overcome a large activation energy to accept another H to form the final product trans-HCOOH*. However, the interfacial O atom of bi-HCOO* could be further attacked by the second coupling of a spillover proton and photoelectron, which could promote the process of bi-HCOO* evolving to the final product trans-HCOOH*.29 In view of the HCOO* path, the hydrogenation of CO2* step and hydrogenation of bi-HCOO* step are both feasible to be assisted by the coupling of a spillover proton and a photoelectron. By contrast, in the hydrogenation of CO2* step in the HOCO* path, the upper O of CO2*, as the target of spillover proton attack in the cis-HOCO* formation step, is far from the interface of Cu9/ZnO(0001). Thus, the formation of the intermediate HOCO* is less assisted by the coupling spillover proton and photoelectron near the interface. Therefore, based on our calculations, the HCOO* path is the optimal reaction pathway for the photocatalytic hydrogenation of CO2 by Cu/ZnO(0001).

3.4. Experimental study of the direct photocatalytic hydrogenation of CO2 on the Cu/ZnO(0001) surface

In order to further interpret the above theoretical results and the suggested reaction route of the direct photocatalytic hydrogenation of CO2 by the Cu/ZnO polar surface, we constructed an illustrative experiment using a custom-built reaction system (see details in the ESI, Fig. S4). The AFM images of the Cu metal on ZnO substrates are shown in Fig. S5 of the ESI.Fig. 5(a) and (b) represent the high-resolution Cu 2p3/2 and O 1s XPS data of the Cu/ZnO(0001), respectively. It can be observed that when Cu was deposited on the ZnO(0001), the binding energy of Cu 2p3/2 shifted to a lower value, indicating an electron transfer process from ZnO(0001) to Cu, which agrees with the above theoretical Bader charge analysis. Correspondingly, the O 1s binding energy of the Cu-loaded ZnO(0001) surface shows an upward shift compared to that of the clean ZnO(0001) surface. The fed H2 and CO2 gases reacted on the clean ZnO or Cu/ZnO polar surface under the radiation of a 300 W xenon lamp. After 4 hours of reaction, the irradiated surfaces of the photocatalysts were dried in a vacuum oven at 60 °C and then scraped with a razor blade to obtain the residual reaction products. The powders scraped from the reacted catalyst surface were then analysed using infrared (IR) transmission spectroscopy. Fig. 5(c) presents the FT-IR spectrum of the surface product on Cu/ZnO(0001) after reaction. The characteristic peaks located at 2970, 2880, 2740, 1668, 1580, 1397, 1378, and 1353 cm−1 are in good agreement with those of zinc formate,47 while the characteristic peaks of copper formate (at 2930, 2850, 1620, and 1350 cm−1) are absent.48 Thus, it can be speculated that formic acid is a possible reduction product of the photocatalytic hydrogenation of CO2 on the Cu/ZnO polar surface. However, for the clean ZnO and Cu/ZnO(000[1 with combining macron]) photocatalysts, no obvious products related to the reduction of CO2 can be identified in the scraped powders from the FTIR spectrum. Therefore, the Cu/ZnO(0001) surface is an efficient photocatalyst for the direct hydrogenation of CO2, which is consistent with the above theoretical discussions.
image file: d3nr01001e-f5.tif
Fig. 5 (a) XPS Cu 2p3/2 and (b) O 1s core level spectra; (c) IR spectrum data of the Cu/ZnO(0001) catalyst after the photocatalytic reaction.

Combining the theoretical and experimental works, we propose the optimal reaction route of the direct photocatalytic hydrogenation of CO2 by Cu/ZnO(0001) as illustrated in Fig. 6. In the oxidation half-reaction, an H2 molecule is adsorbed by Cu nanoislands and then is bound to a Cu atom after disassociation. These two disassociated H atoms are further oxidized to two H+ by the photo-generated holes. In the reduction half-reaction, CO2 is absorbed at the edge site of Cu/ZnO(0001) and becomes a bidentate carbonate structure (intermediate CO2*, structure (1) in Fig. 6). The intermediate CO2* is then attacked by H+ coupling with a photoelectron on the interface, via simultaneous C–Cu bond cleavage and C–H bond formation steps, resulting in the HCOO* intermediate (structure (2) in Fig. 6). The intermediate HCOO* is metastable, and it is energetic favourable for it to undergo structural rearrangement to form the intermediate bi-HCOO* (structure (3) in Fig. 6). Finally, the secondly spillover proton coupled with the photoelectron would attack the interfacial O of the bi-HCOO* intermediate, leading to the final trans-HCOOH* product (formic acid).


image file: d3nr01001e-f6.tif
Fig. 6 The proposed schematic illustration of the reaction route of the direct photocatalytic hydrogenation of CO2 on the Cu/ZnO(0001) surface.

4. Conclusions

In this paper, the adsorption of CO2 and H2 on the polar surfaces of ZnO(0001), ZnO(000[1 with combining macron]), Cu/ZnO(0001), and Cu/ZnO(000[1 with combining macron]) have been systematically investigated using DFT simulations. Our calculations showed that CO2 can only be weakly physisorbed on clean ZnO(0001) and ZnO(000[1 with combining macron]) polar surfaces. In contrast, the Cu deposited on the ZnO polar surfaces Cu/ZnO(0001) and Cu/ZnO(000[1 with combining macron]) could greatly promote the chemical adsorption of CO2 on the surfaces. The electronic structure studies clearly show that Cu deposition induces charge redistribution on the Cu/ZnO polar surfaces and improves the charge donor ability of the ZnO substrate near Cu atoms, and thereby the chemical adsorption of CO2 on Cu/ZnO(0001) is enhanced. H2 prefers to chemisorb on the deposited Cu nanoislands, ZnO(000[1 with combining macron]) surface and Cu/ZnO(000[1 with combining macron]). Therefore, when Cu/ZnO(0001) and Cu/ZnO(000[1 with combining macron]) are exposed to CO2 and H2, Cu/ZnO(000[1 with combining macron]) could be saturated and poisoned by the strong chemisorption of H2, while, in contrast, Cu/ZnO(0001) has more opportunity to adsorb H2 and CO2 simultaneously near the edge site to promote the hydrogenation of CO2. Thus, it can be deduced that Cu/ZnO(0001) should be a favourable surface for the photocatalytic hydrogenation of CO2. This deduction is then confirmed by the photocatalytic hydrogenation experiment. The optimal HCOO* reaction pathway for the photocatalytic hydrogenation of CO2 by Cu/ZnO(0001) is proposed based on the theoretical and experimental results. In this optimal HCOO* path, the hydrogenation of CO2* step and hydrogenation of HCOO* step could be promoted by the coupling of a photo-generated spillover proton and a photoelectron on the interface of Cu/ZnO(0001). This research demonstrated the feasibility of the photocatalytic reduction of CO2 on Cu/ZnO(0001), and will be helpful for the reasonable design of high-efficiency catalysts.

Author contributions

Han Xiao carried out the DFT calculations and analysed the data. Yihong Lian and Shiduo Zhang performed Bader charge analyses and calculated the electronic location function. Jiye Zhang conducted the experiments. Minyi Zhang and Chunsen Li designed the project, analysed the data, and wrote the manuscript.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This investigation was based on work supported by the National Natural Science Foundation of China (Grant No. 21933009 and 51472241) and the Natural Science Foundation of Fujian Province (Grant No. 2021J01525).

References

  1. E. J. Carrington, C. A. McAnally, A. J. Fletcher, S. P. Thompson, M. Warren and L. Brammer, Nat. Chem., 2017, 9, 882–889 CrossRef CAS PubMed.
  2. F. Zha, J. Ding, Y. Chang, J. F. Ding, J. Y. Wang and J. Ma, Ind. Eng. Chem. Res., 2012, 51, 345 CrossRef CAS.
  3. F. C. Meunier, Angew. Chem., Int. Ed., 2011, 50, 4053 CrossRef CAS PubMed.
  4. P. Gao, S. G. Li, X. A. Bu, S. S. Dang, Z. Y. Liu, H. Wang, L. S. Zhong, M. H. Qiu, C. G. Yang, J. Cai, W. Wei and Y. H. Sun, Nat. Chem., 2017, 9, 1019–1024 CrossRef CAS PubMed.
  5. S. Patial, R. Kumar, P. Raizada, P. Singh, Q. V. Le, E. Lichtfouse, D. L. T. Nguyen and V. H. Nguyen, Environ. Res., 2021, 197, 111134 CrossRef CAS PubMed.
  6. R. G. Zhang, H. Y. Liu, B. J. Wang and L. X. Ling, Appl. Catal., B, 2012, 126, 108 CrossRef CAS.
  7. S. A. Kondrat, P. J. Smith, J. H. Carter, J. S. Hayward, G. J. Pudge, G. Shaw, M. S. Spencer, J. K. Bartley, S. H. Taylora and G. J. Hutchings, Faraday Discuss., 2017, 197, 287–307 RSC.
  8. X. Li, J. Q. Wen, J. X. Low, Y. P. Fang and J. G. Yu, Sci. China Mater., 2014, 57, 70–100 CrossRef.
  9. M. Marszewski, S. W. Cao, J. G. Yu and M. Jaroniec, Mater. Horiz., 2015, 2, 261–278 RSC.
  10. Q. J. Xiang, B. Cheng and J. G. Yu, Angew. Chem., Int. Ed., 2015, 54, 11350–11366 CrossRef CAS PubMed.
  11. E. S. Jang, J. H. Won, S. J. Hwang and J. H. Choy, Adv. Mater., 2006, 18, 3309 CrossRef CAS.
  12. N. Kislov, J. Lahiri, H. Verma, D. Y. Goswami, E. Stefanakos and M. Batzill, Langmuir, 2009, 25, 3310 CrossRef CAS PubMed.
  13. H. Fatima, M. R. Azhar, Y. J. Zhong, Y. Arafat, M. Khiadani and Z. P. Shao, J. Colloid Interface Sci., 2022, 614, 538–546 CrossRef CAS PubMed.
  14. A. X. Wang, W. J. Wang, J. Y. Chen, R. D. Mao, Y. P. Pang, Y. G. Li, W. Chen, D. C. Chen, D. Hao, B. J. Ni, M. Saunders and G. H. Jia, J. Phys. Chem. Lett., 2020, 11, 4990–4997 CrossRef CAS PubMed.
  15. L. Y. Yang, S. Y. Dong, J. H. Sun, J. L. Feng, Q. H. Wu and S. P. Sun, J. Hazard. Mater., 2010, 179, 438 CrossRef CAS PubMed.
  16. S. He, S. T. Zhang, J. Lu, Y. F. Zhao, J. Ma, M. Wei, D. G. Evans and X. Duan, Chem. Commun., 2011, 47, 10797 RSC.
  17. A. Mclaren, T. Valdes-Solis, G. Q. Li and S. C. Tsang, J. Am. Chem. Soc., 2009, 131, 12540 CrossRef CAS PubMed.
  18. Q. S. Guo, Q. H. Zhang, H. Z. Wang, Z. F. Liu and Z. Zhao, Catal. Commun., 2016, 77, 118–122 CrossRef CAS.
  19. X. Y. He, M. Liu, Z. Liang, Z. Y. Wang, P. Wang, Y. Y. Liu, H. F. Cheng, Y. Dai, Z. K. Zheng and B. B. Huang, J. Solid State Chem., 2021, 298, 122113 CrossRef CAS.
  20. D. Weibel, Z. R. Jovanovic and A. Steinfeld, Chem. Eng. Sci., 2017, 165, 96–107 CrossRef CAS.
  21. M. Dorraja, M. Alizadehb, N. A. Sairia, W. J. Basiruna, B. T. Gohd, P. M. Woia and Y. Aliasa, Appl. Surf. Sci., 2017, 414, 251–261 CrossRef.
  22. Y. Zhang, Y. J. Ji, J. Li, H. Z. Liu, Z. Y. Zhong and F. B. Su, J. Catal., 2017, 348, 233–245 CrossRef CAS.
  23. S. Singh, R. Pendurthi, M. Khanuja, S. S. Islam, S. Rajput and S. M. Shivaprasad, Appl. Phys. A, 2017, 123, 184 CrossRef.
  24. R. Gleißner, H. Noei, S. M. Chung, G. D. L. Semione, E. E. Beck, A. C. Dippel, O. Gutowski, G. Gizer, V. Vonk and A. Stierle, J. Phys. Chem. C, 2021, 125, 23561–23569 CrossRef.
  25. F. Jiang, Y. Yang, L. Wang, Y. F. Li, Z. H. Fang, Y. B. Xu, B. Liu and X. H. Liu, Catal. Sci. Technol., 2022, 12, 551–564 RSC.
  26. R. C. Pawar, D. H. Choi, J. S. Lee and C. S. Lee, Mater. Chem. Phys., 2015, 151, 167–180 CrossRef CAS.
  27. F. L. Liao, Y. Q. Huang, J. W. Ge, W. R. Zheng, K. Tedsree, P. Collier, X. L. Hong and S. C. Tsang, Angew. Chem., Int. Ed., 2011, 50, 2162–2165 CrossRef CAS PubMed.
  28. R. Narayanan and M. A. El-Sayed, J. Phys. Chem. B, 2005, 109, 12663–12676 CrossRef CAS PubMed.
  29. Q. L. Tang, W. T. Zou, R. K. Huang, Q. Wang and X. X. Duan, Phys. Chem. Chem. Phys., 2015, 17, 7317–7333 RSC.
  30. M. D. Higham, D. Mora-Fonz, A. A. Sokol, S. M. Woodley and C. R. A. Catlow, J. Mater. Chem. A, 2020, 8, 22840–22857 RSC.
  31. G. Kresse and J. Furthmüller, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169 CrossRef CAS PubMed.
  32. G. Kresse and J. Furthmüller, Comput. Mater. Sci., 1996, 6, 15 CrossRef CAS.
  33. P. E. Blöchl, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953–17979 CrossRef PubMed.
  34. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
  35. H. Li, L. K. Schirra, J. Shim, H. Cheun, B. Kippelen, O. L. A. Monti and J. L. Bredas, Chem. Mater., 2012, 12, 3044 CrossRef.
  36. Q. L. Tang, W. T. Zou, R. K. Huang, Q. Wang and X. X. Duan, Phys. Chem. Chem. Phys., 2015, 17, 7317 RSC.
  37. W. W. Lin, D. G. Chen, J. Y. Zhang, Z. Lin, J. K. Huang, W. Li, Y. H. Wang and F. Huang, Cryst. Growth Des., 2009, 9, 4378 CrossRef CAS.
  38. A. N. Mariano and R. E. Hanneman, J. Appl. Phys., 1963, 34, 384 CrossRef CAS.
  39. E. H. Kisi and M. M. Elcombe, Acta Crystallogr., Sect. C: Cryst. Struct. Commun., 1989, 45, 1867 CrossRef.
  40. Q. L. Tang and Q. H. Luo, J. Phys. Chem. C, 2013, 117, 22954 CrossRef CAS.
  41. F. Jiang, Y. Yang, L. Wang, Y. F. Li, Z. H. Fang, Y. B. Xu, B. Liu and X. H. Liu, Catal. Sci. Technol., 2022, 12, 551–564 RSC.
  42. L. Qiao, Y. Zeng, C. Q. Qu, H. Z. Zhang, X. Y. Hu, L. J. Song, D. M. Bi and S. J. Liu, Phys. E, 2013, 48, 7 CrossRef CAS.
  43. L. Qiao, Y. Zeng, C. Q. Qu, X. Y. Hu, L. J. Song, S. J. Liu and Y. M. Sui, Nano, 2014, 9, 1450006 CrossRef.
  44. R. F. W. Bader, Acc. Chem. Res., 1985, 18, 9 CrossRef CAS.
  45. G. Henkelman, A. Arnaldsson and H. Jónsson, Comput. Mater. Sci., 2006, 36, 354 CrossRef.
  46. C. T. Au, W. Hirsch and W. Hirschwald, Surf. Sci., 1988, 197, 391 CrossRef CAS.
  47. S. Fujita, M. Usui, H. Ito and N. Takezawa, J. Catal., 1995, 157, 403–413 CrossRef CAS.
  48. W. R. A. M. Robinson and J. C. Mol, Appl. Catal., A, 1993, 98, 81–97 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3nr01001e

This journal is © The Royal Society of Chemistry 2023