Next Article in Journal
Hollow g-C3N4@Cu0.5In0.5S Core-Shell S-Scheme Heterojunction Photothermal Nanoreactors with Broad-Spectrum Response and Enhanced Photocatalytic Performance
Previous Article in Journal
Molecular Cluster Complex of High-Valence Chromium Selenide Carbonyl as Effective Electrocatalyst for Water Oxidation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

DFT Study of CO2 Reduction Reaction to CH3OH on Low-Index Cu Surfaces

1
School of Chemistry and Chemical Engineering, Chongqing University of Technology, Chongqing 400054, China
2
School of Chemistry and Chemical Engineering, Chongqing University, Chongqing 400044, China
3
Chongqing Medical and Pharmaceutical College, Chongqing 400020, China
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(4), 722; https://doi.org/10.3390/catal13040722
Submission received: 21 March 2023 / Revised: 6 April 2023 / Accepted: 9 April 2023 / Published: 11 April 2023
(This article belongs to the Section Computational Catalysis)

Abstract

:
The electrochemical reduction of CO2 is an efficient method to convert CO2 waste into hydrocarbon fuels, among which methanol is the direct liquid fuel in the direct methanol fuel cells (DMFC). Copper is the most widely used catalyst for CO2 reduction reaction (CO2RR); the reaction is affected by the surface morphology of the copper. Here, the morphology effect and the mechanism of CO2RR on three typical low-index Cu (100), Cu (110) and Cu (111) surfaces are studied. According to our results, Cu (110) provides the optimum surface for the CO2RR via CO2 → *COOH → *CO → *CHO → *CH2O → *CH2OH → CH3OH pathway, where the reduction reaction of CO2 to *COOH is the potential-determining step (PDS). This is because Cu (110) has the highest d band center, which promotes the adsorption of *COOH.

1. Introduction

Excessive use of fossil energy sources leads to large CO2 emissions and a range of global climate problems [1,2,3]. Thus, decreasing the amount of CO2 in the atmosphere is crucial. Carbon capture and storage and carbon conversion to hydrocarbon products have become increasingly important [4,5]. However, the capture and storage of CO2 requires significant amounts of energy. In contrast, the direct conversion of CO2 into various hydrocarbon fuels is less energy intensive and is a much more attractive option. To date, thermoscatalysis, photocatalysis and electrocatalysis have been explored for converting CO2 into various C1 hydrocarbons (CH3OH, HCOOH, CH4), C2 hydrocarbons (C2H4, CH3CH2OH) and C2+ hydrocarbons under ambient conditions [6,7]. CH3OH is one of the most important hydrocarbon fuels since it can be used directly as the fuel of direct methanol fuel cells (DMFC) [8,9]; significant achievements have been attained in the development of cost-effective and product-selective electrochemical CO2 reduction reactions (CO2RR) [10]. Though both homogeneous and heterogeneous catalysts have been widely investigated [11,12], copper-based catalysts have been considered as the best catalysts for CO2RR [13,14]. Both Cu+ ion [15] and metallic Cu [16,17] are considered as the active catalytic site for CO2RR. Nevertheless, metallic Cu provides a comparable catalytic function to that of industrial catalysts during the synthesis of methanol [18]. In addition, the CO2RR can be significantly affected by the morphology of the copper surface [19].
For the CO2RR to methanol, as listed in Equations (1)–(3), three typical reduction processes, including the two-electron (2e) pathway leading to the formation of CO, the four-electron (4e) pathway leading to HCHO and the six-electron (6e) pathway leading to CH3OH, were systematically studied in this paper.
C O 2 g + 2 H + + 2 e C O g + H 2 O ( l )
C O 2 g + 4 H + + 4 e H C H O l + H 2 O ( l )
C O 2 g + 6 H + + 6 e C H 3 O H l + H 2 O ( l )
Carbon dioxide has very strong C-O bonds; as such, the dissociation of C-O bonds in carbon dioxide is not a straightforward method. The pathway including the hydrogenation of CO2 to COOH ( C O 2 + H + + e + * * C O O H ) with an indirect C-O bond dissociation and a following dissociation of COOH to CO ( * C O O H + H + + e * C O + H 2 O ) is much more favorable than direct C-O bond dissociation [20,21]. Thus, the indirect C-O bond dissociation method was considered in the two-electron reduction pathway. To elucidate the morphology-dependent electrochemical CO2RR, the adsorption and Gibbs free energy of the reaction via the CO2 → *COOH → *CO → *CHO → *CH2O → *CH2OH → CH3OH pathway were systematically investigated on the Cu (100), Cu (110) and Cu (111) surfaces. It was found that Cu (110) provides the best catalytic activity towards CO2RR, and the first step of CO2 hydrogenation to COOH is the potential-determining step (PDS), with an overpotential of 0.72 V. The method here can also be expanded to design other morphology-dependent CO2RR catalysts.

2. Computational Details

Density functional theory (DFT) calculations were performed using the Vienna Ab initio Simulation Package [22,23]. The projector-augmented wave (PAW) method was applied to deal with the ion-electron interactions [24], and the Perdew–Burke–Ernzerhof (PBE) [25] functional within generalized gradient approximation (GGA) [26] was used to describe the exchange-correlation functional. The plane wave energy cutoff was set as 500 eV, and the Monkforst–Pack k-point mesh of 4 × 4 × 1 was employed [27]. The convergence criteria of electronic energy and forces were 1.0 × 10−5 eV and 0.02 eV/Å, respectively. A 3 × 3 slab model consisting of four Cu layers was constructed for various low-index Cu (111), Cu (110) and Cu (100) surfaces, and a vacuum of 15 Å was built to avoid any interaction between different two surfaces.
The surface energy (Esur) of various copper surface is calculated according to Equation (4):
E s u r = ( E s l a b n E C u ) / 2 A
where Eslab is the calculated energy of various optimized copper surface, n is the number of Cu atoms in the slab, ECu represents the energy of a single Cu atom in the crystal cell and A is the area of various copper surface.
The adsorption energy (Eads) of the species adsorbed on various low-index Cu surfaces was calculated according to the Equation (5):
E a d s = E C u / C E C E C u
where ECu/C, EC and ECu represent the total energy of the adsorption system, the energy of adsorbed carbon species and the energy of Cu surface, respectively. A more negative value of Eads means a more stable adsorption.
Gibbs free energy change of elementary reaction was calculated according to Equation (6):
G = E + E Z P E T S + G p H + G U
where ΔE is the change of reaction energy directly obtained from DFT total energy, ΔEZPE is the change of zero-point energy, T is the temperature (298.15 K), ΔS is the change of entropy and the results of ZPE and TS corrections to G are listed in Table S1. G U = n e U , where n is the number of transferred electrons and U is the electrode potential. ΔGpH is the correction of the H+ free energy by the concentration, G p H = k B T × ln 10 × p H , kB is the Boltzmann constant and the value of pH was set to be zero for acidic condition. Zero-point energy of every CO2RR intermediate was computed from vibrational frequencies. The vibrational modes of the adsorbate were computed explicitly, while the Cu surfaces were fixed since the vibrations of the substrate can be negligible. The entropy of thermodynamically stable species in the gas phase was taken from the National Institute of Standards and Technology (NIST) database.

3. Results and Discussion

3.1. Structures and Stability of Low-Index Cu Surfaces

As shown in Table 1, the surface energies of Cu (100), Cu (110) and Cu (111) are 1.91 J/m2, 2.04 J/m2 and 1.77 J/m2, respectively. The largest difference of surface energy among the three low index Cu surfaces is only 0.27 J/m2; this means that the three typical low-index surfaces share the comparable stability and can be synthesized easily in the experiments. The closest distance between two Cu atoms (Figure 1) is 2.555 ± 0.001 Å in all the Cu surfaces; this is consistent with that in the bulk copper. These findings can be supported by the experimental synthesis performed by Gawande et al. [28] and Ghosh et al. [29].

3.2. Gibbs Free Energy Change during CO2RR on Low-Index Cu Surfaces

Gibbs free energy changes of the elementary reaction in the CO2RR via the CO2 → *COOH → *CO → *CHO → *CH2O → *CH2OH → CH3OH pathway were systematically investigated. Three typical products, including CO, HCHO and CH3OH, during the electrochemical reduction of CO2 were studied comparatively on low-index Cu (100), Cu (110) and Cu (111) surfaces. The reduction pathway via formate (HCOO) species was omitted since the formate is a ‘dead-end’ species [30]. Therefore, a proton-coupled electron transfer process that leads to the formation of O-H bond was considered in this paper [31].
All the optimized structures of CO2RR to CO through a two-electron pathway, to HCHO through a four-electron pathway and to CH3OH through a six-electron pathway are presented in Figure 2; the related free energy changes of elementary reaction for the electrochemical CO2RR in presented in Figure 3. A proton-electron pair was added into the elementary step in the pathway, and the energetic levels of all the species are given at U = 0 VRHE (RHE = reversible hydrogen electrode). The most favorable pathway is the pathway with the lowest positive change of Gibbs free energy between any two steps.
As shown in Figure 3, the free energy change of CO2RR in the three pathways has a similar trend; the first step reaction of carbon dioxide has the largest energy barrier and determines the activity of the overall carbon dioxide conversion, including CO2RR to CO, HCHO and CH3OH [32]. Thus, the hydrogenation of CO2 to *COOH is considered as the PDS for the whole reduction processes. Moreover, Cu (110) provides the best catalytic activity among the three low-index Cu surfaces, since the energy barrier of the PDS is only 0.72 eV, which is 0.17 eV and 0.50 eV lower than that on Cu (111) and Cu (100), respectively. Specifically, for the reaction of CO2 to CO via the two-electron pathway, the hydrogenation of *COOH to *CO takes place via a proton-electron transfer step on the three low-index Cu surfaces; the reduction reaction of *COOH to *CO is energy favorable since it is an exothermic reaction. Moreover, CO can desorb much easier on Cu (110) than on both Cu (100) and Cu (111) surfaces, since the CO desorptions on both Cu (100) and Cu (111) surface must overcome a larger energy barrier. For the reaction of CO2 to HCHO via the four-electron pathway, the reaction follows the two-electron pathway except for the CO desorption process. The hydrogenation of *CO to *CHO in the four-electron pathway is slightly easier than the CO2 hydrogenation to *COOH step, and the energy barriers for *CO reduction to *CHO are 0.38 eV on Cu (110), 0.81 eV on Cu (100) and 0.85 eV on Cu (111). Cu (110) still provides the best catalytic activity for the reduction reaction of CO to C1 chemicals. The reduction of *CHO to *HCHO is an exothermic process on various low-index Cu surfaces, and the *HCHO can desorb spontaneously since the free energy change for HCHO desorption is exothermic and the adsorption energy (Table 2) is very low. The reduction of CO2RR to CH3OH via the six-electron pathway shares the same procedures with that in the four-electron pathway for the initial four electrons. The hydrogenation of adsorbed *HCHO to *CH2OH is an endothermic process on various low-index Cu surfaces; as such, the three low-index Cu surfaces share similar activity for this step. The energy barriers are 0.17 eV on Cu (110), 0.07 eV on Cu (111) and 0.07 eV on Cu (100). The reduction of *CH2OH to methanol is a spontaneous reaction on all the low-index Cu surfaces.

3.3. Adsorption of Various Intermediates

The adsorption of various intermediates were studied further. It was found that the C atom of every intermediate is always bound to Cu atom for the three low-index Cu surfaces; this is consistent with previous findings [33,34]. As shown in Table 2, the largest adsorption energy for *COOH can be obtained among all the intermediates, and the values are −2.04 eV on Cu (110), −1.86 eV Cu (100) and −1.59 eV on Cu (111). Charge density difference (Figure 4) accounts for the strongest *COOH adsorption on Cu (110), since more electrons transfer from Cu (110) to adsorbed *COOH than that which can be observed on Cu (111) and Cu (100). Furthermore, the d band center of Cu (110), as shown in Figure 5, is −2.84 eV; this is 0.11 eV and 0.52 eV closer to the Fermi level than that of both Cu (111) and Cu (100), respectively. In addition, Cu (110) shows the highest peak of d band near the Fermi level when compared with Cu (111) and Cu (100). This can account for the largest adsorption energy for the COOH adsorbed on Cu (110) among the three low-index surfaces.
*COOH can be further reduced to the adsorbed *CO according to the CO2RR mechanism. As shown in Figure 2, CO* prefers to adsorb at the top site of both Cu (111) and Cu (100), with adsorption energy of −0.68 eV and −0.80 eV, respectively. This is consistent with the rule that the closer the d band center to the Fermi level, the stronger adsorption energy for the adsorbate. However, the adsorption energy of CO is only −0.54 eV on Cu (110), which is not consistent with the d band center rule. The reason may lie in the fact that CO prefers to adsorb at the bridge position of Cu (110) instead of the top site on Cu (111) and Cu (100).
The adsorption of CHO and CH2OH is similar with that of COOH. The adsorption energies of every species were stronger on Cu (110) than that on either Cu (111) or Cu (100), which follows the d band center rule. The adsorption of HCHO on various Cu surfaces is very weak. It is only −0.32 eV on Cu (110) and −0.17 eV on Cu (100). However, it is close to 0 on the Cu (111) surface. This may be related to the maximized electron localization of the HCHO molecule.

4. Conclusions

DFT calculations were performed to understand the CO2RR via the two-electron pathway to CO, the four-electron pathway to HCHO and the six-electron pathway to CH3OH. Our results show that the initial activation of carbon dioxide is the PDS and that this determines the activity of the overall carbon dioxide conversion. Moreover, Cu (110) provides the best catalytic activity among the three low-index surfaces, since the energy barrier in the PDS is only 0.72 eV; this is 0.17 eV and 0.50 eV lower than that on Cu (111) and Cu (100), respectively. Furthermore, the d band center of Cu (110) is −2.84 eV; this is 0.11 eV and 0.52 eV closer to the Fermi level than that of both Cu (111) and Cu (100), respectively. Based on these observations, we propose that the Cu (110) surface can be a promising catalyst for the efficient and selective reduction of CO2RR to CH3OH. As such, this method can be used to design other catalysts used for CO2RR.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13040722/s1, Table S1: Thermal correction of ZPE and TS to free energy (eV) of different adsorption species at 298.15 K.

Author Contributions

Conceptualization, Q.X. and X.Q. (Xueqiang Qi); methodology, Q.X. and X.Q. (Xuede Qi); software, Q.X.; validation, Q.X., T.Y. and X.Q. (Xuede Qi); formal analysis, J.J.; investigation, K.Z. and F.X.; resources, Q.X. and X.Q. (Xuede Qi); data curation, Y.Z.; writing—original draft preparation, Q.X. and X.Q. (Xueqiang Qi); writing—review and editing, Q.X. and X.Q. (Xueqiang Qi); visualization, K.L.; supervision, X.Q. (Xueqiang Qi). All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by China Postdoctoral Science Foundation (2021M700621).

Data Availability Statement

The data presented in this study are available in this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kumar, B.; Llorente, M.; Froehlich, J.; Dang, T.; Sathrum, A.; Kubiak, C.P. Photochemical and photoelectrochemical reduction of CO2. Annu. Rev. Phys. Chem. 2012, 63, 541–569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Talapaneni, S.N.; Singh, G.; Kim, I.Y.; AlBahily, K.; Al-Muhtaseb, A.H.; Karakoti, A.S.; Tavakkoli, E.; Vinu, A. Nanostructured Carbon Nitrides for CO2 Capture and Conversion. Adv. Mater. 2020, 32, e1904635. [Google Scholar] [CrossRef] [PubMed]
  3. Wu, W.; Tan, L.; Chang, H.; Zhang, C.; Tan, X.; Liao, Q.; Zhong, N.; Zhang, X.; Zhang, Y.; Ho, S.-H. Advancements on process regulation for microalgae-based carbon neutrality and biodiesel production. Renew. Sustain. Energy Rev. 2023, 171, 112969. [Google Scholar] [CrossRef]
  4. Roy, S.C.; Varghese, O.K.; Paulose, M.; Grimes, C.A. Toward Solar Fuels: Photocatalytic Conversion ofCarbon Dioxide to Hydrocarbons. ACS Nano 2010, 4, 1259–1278. [Google Scholar] [CrossRef]
  5. Xue, Q.; Qi, X.; Yang, T.; Jiang, J.; Zhou, Q.; Fu, C.; Yang, N. DFT Study on the CO(2) Reduction to C(2) Chemicals Catalyzed by Fe and Co Clusters Supported on N-Doped Carbon. Nanomaterials 2022, 12, 2239. [Google Scholar] [CrossRef]
  6. Li, X.; Yu, J.; Jaroniec, M.; Chen, X. Cocatalysts for Selective Photoreduction of CO2 into Solar Fuels. Chem. Rev. 2019, 119, 3962–4179. [Google Scholar] [CrossRef]
  7. Habisreutinger, S.N.; Schmidt-Mende, L.; Stolarczyk, J.K. Photocatalytic reduction of CO2 on TiO2 and other semiconductors. Angew. Chem. Int. Ed. 2013, 52, 7372–7408. [Google Scholar] [CrossRef]
  8. Goeppert, A.; Czaun, M.; Jones, J.P.; Surya Prakash, G.K.; Olah, G.A. Recycling of carbon dioxide to methanol and derived products–closing the loop. Chem. Soc. Rev. 2014, 43, 7995–8048. [Google Scholar] [CrossRef]
  9. Yang, T.; Xue, Q.; Yu, X.; Qi, X.; Wu, R.; Lu, S.; Gu, Z.; Jiang, J.; Nie, Y. DFT Study on Methanol Oxidation Reaction Catalyzed by PtmPdn Alloys. Coatings 2022, 12, 918. [Google Scholar] [CrossRef]
  10. Torrente-Murciano, L.; Mattia, D.; Jones, M.D.; Plucinski, P.K. Formation of hydrocarbons via CO2 hydrogenation—A thermodynamic study. J. CO2 Util. 2014, 6, 34–39. [Google Scholar] [CrossRef]
  11. Mandal, S.C.; Rawat, K.S.; Nandi, S.; Pathak, B. Theoretical insights into CO2 hydrogenation to methanol by a Mn–PNP complex. Catal. Sci. Technol. 2019, 9, 1867–1878. [Google Scholar] [CrossRef]
  12. Shohei Tada, S.K.; Honma, T.; Kamei, H.; Nariyuki, A.; Kon, K.; Toyao, T.; Shimizu, K.-I.; Satokawa, S. Design of Interfacial Sites between Cu and Amorphous ZrO2 Dedicated to CO2-to-Methanol Hydrogenation. ACS Catal. 2018, 8. [Google Scholar] [CrossRef]
  13. 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] [PubMed]
  14. Hussain, J.; Skúlason, E.; Jónsson, H. Computational Study of Electrochemical CO2 Reduction at Transition Metal Electrodes. Procedia Computer Sci. 2015, 51, 1865–1871. [Google Scholar] [CrossRef]
  15. Szanyi, J.; Goodman, D.W. Methanol synthesis on a Cu(100) catalyst. Catal. Lett. 1991, 10, 383–390. [Google Scholar] [CrossRef]
  16. Chinchen, G.C.; Waugh, K.C.; Whan, D.A. The activity and state of the copper surface in methanol synthesis catalysts. Applied. Catal. 1986, 25, 101–107. [Google Scholar] [CrossRef]
  17. Rawat, K.S.; Mahata, A.; Pathak, B. Thermochemical and electrochemical CO2 reduction on octahedral Cu nanocluster: Role of solvent towards product selectivity. J. Catal. 2017, 349, 118–127. [Google Scholar] [CrossRef]
  18. Rasmussen, P.B.; Holmblad, P.M.; Askgaard, T.; Ovesen, C.V.; Stoltze, P.; Nørskov, J.K. Chorkendorff Methanol synthesis on Cu(100) from a binary gas mixture of CO2 and H2. Catal. Lett. 1994, 26, 373–381. [Google Scholar] [CrossRef]
  19. Zhang, X.; Liu, J.-X.; Zijlstra, B.; Filot, I.A.W.; Zhou, Z.; Sun, S.; Hensen, E.J.M. Optimum Cu nanoparticle catalysts for CO2 hydrogenation towards methanol. Nano Energy 2018, 43, 200–209. [Google Scholar] [CrossRef]
  20. Ko, J.; Kim, B.-K.; Han, J.W. Density Functional Theory Study for Catalytic Activation and Dissociation of CO2 on Bimetallic Alloy Surfaces. J. Phys. Chem. C 2016, 120, 3438–3447. [Google Scholar] [CrossRef]
  21. Muttaqien, F.; Hamamoto, Y.; Inagaki, K.; Morikawa, Y. Dissociative adsorption of CO2 on flat, stepped, and kinked Cu surfaces. J. Chem. Phys. 2014, 141, 034702. [Google Scholar] [CrossRef] [PubMed]
  22. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169. [Google Scholar] [CrossRef] [PubMed]
  23. Kresse, G.; Furthmiiller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  24. Kresse, G.; Furthmiiller, J. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B Condens Matter 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  25. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  26. White, J.A.; Bird, D.M. Implementation of gradient-corrected exchange-correlation potentials in Car-Parrinello total-energy calculations. Phys. Rev. B Condens Matter 1994, 50, 4954–4957. [Google Scholar] [CrossRef]
  27. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  28. Gawande, M.B.; Goswami, A.; Felpin, F.X.; Asefa, T.; Huang, X.; Silva, R.; Zou, X.; Zboril, R.; Varma, R.S. Cu and Cu-Based Nanoparticles: Synthesis and Applications in Catalysis. Chem. Rev. 2016, 116, 3722–3811. [Google Scholar] [CrossRef] [Green Version]
  29. Ghosh, R.; Sahoo, A.K.; Ghosh, S.S.; Paul, A.; Chattopadhyay, A. Blue-emitting copper nanoclusters synthesized in the presence of lysozyme as candidates for cell labeling. ACS Appl. Mater. Interfaces 2014, 6, 3822–3828. [Google Scholar] [CrossRef]
  30. Vesselli, E.; Rizzi, M.; De Rogatis, L.; Ding, X.; Baraldi, A.; Comelli, G.; Savio, L.; Vattuone, L.; Rocca, M.; Fornasiero, P.; et al. Hydrogen-Assisted Transformation of CO2 on Nickel: The Role of Formate and Carbon Monoxide. J. Phys. Chem. Lett. 2009, 1, 402–406. [Google Scholar] [CrossRef]
  31. Ma, M.; Trzesniewski, B.J.; Xie, J.; Smith, W.A. Selective and Efficient Reduction of Carbon Dioxide to Carbon Monoxide on Oxide-Derived Nanostructured Silver Electrocatalysts. Angew. Chem. Int. Ed. 2016, 55, 9748–9752. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Guo, C.; Wei, S.; Zhou, S.; Zhang, T.; Wang, Z.; Ng, S.-P.; Lu, X.; Wu, C.-M.L.; Guo, W. Initial Reduction of CO2 on Pd-, Ru-, and Cu-Doped CeO2(111) Surfaces: Effects of Surface Modification on Catalytic Activity and Selectivity. ACS Appl. Mater. Interfaces 2017, 9, 26107–26117. [Google Scholar] [CrossRef] [PubMed]
  33. Zijlstra, B.; Zhang, X.; Liu, J.-X.; Filot, I.A.W.; Zhou, Z.; Sun, S.; Hensen, E.J.M. First-principles microkinetics simulations of electrochemical reduction of CO2 over Cu catalysts. Electrochim. Acta 2020, 335, 135665. [Google Scholar] [CrossRef]
  34. Ouyang, T.; Fan, W.; Guo, J.; Zheng, Y.; Yin, X.; Shen, Y. DFT study on Ag loaded 2H-MoS2 for understanding the mechanism of improved photocatalytic reduction of CO2. Phys. Chem. Chem. Phys. 2020, 22, 10305–10313. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Top view and side view of the optimized structures of (a) Cu (111), (b) Cu (100) and (c) Cu (110), respectively.
Figure 1. Top view and side view of the optimized structures of (a) Cu (111), (b) Cu (100) and (c) Cu (110), respectively.
Catalysts 13 00722 g001
Figure 2. Top and side views of CO2RR to CO through a two-electron pathway, to HCHO through a four-electron pathway and to CH3OH through a six-electron pathway on Cu (111), Cu (100) and Cu (110), respectively.
Figure 2. Top and side views of CO2RR to CO through a two-electron pathway, to HCHO through a four-electron pathway and to CH3OH through a six-electron pathway on Cu (111), Cu (100) and Cu (110), respectively.
Catalysts 13 00722 g002
Figure 3. Free energy diagram of CO2RR to CO, HCHO and CH3OH over Cu (111), Cu (100) and Cu (110), respectively.
Figure 3. Free energy diagram of CO2RR to CO, HCHO and CH3OH over Cu (111), Cu (100) and Cu (110), respectively.
Catalysts 13 00722 g003
Figure 4. Charge density difference of COOH binding on (a) Cu (111), (b) Cu (100) and (c) Cu (110), respectively. The yellow region represents the electron accumulation while the blue region represents the electron depletion. The isosurface value is 0.0018 e/Bohr3.
Figure 4. Charge density difference of COOH binding on (a) Cu (111), (b) Cu (100) and (c) Cu (110), respectively. The yellow region represents the electron accumulation while the blue region represents the electron depletion. The isosurface value is 0.0018 e/Bohr3.
Catalysts 13 00722 g004
Figure 5. Projected densities of states (PDOS) of Cu (111), Cu (100) and Cu (110). The dotted line represents the d band center.
Figure 5. Projected densities of states (PDOS) of Cu (111), Cu (100) and Cu (110). The dotted line represents the d band center.
Catalysts 13 00722 g005
Table 1. Surface energy of Cu (111), Cu (100) and Cu (110) surface, respectively.
Table 1. Surface energy of Cu (111), Cu (100) and Cu (110) surface, respectively.
SurfaceCu (111)Cu (100)Cu (110)
Esur (J/m2)1.771.912.04
Table 2. The adsorption energy of various intermediate species on Cu (111), Cu (100) and Cu (110), respectively.
Table 2. The adsorption energy of various intermediate species on Cu (111), Cu (100) and Cu (110), respectively.
SurfaceEads/eV
*COOH*CO*CHO*CH2O*CH2OH
Cu (111)−1.59−0.68−1.22−0.02−0.92
Cu (100)−1.86−0.80−1.40−0.17−1.10
Cu (110)−2.04−0.54−1.59−0.32−1.20
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

Xue, Q.; Qi, X.; Li, K.; Zeng, Y.; Xu, F.; Zhang, K.; Yang, T.; Qi, X.; Jiang, J. DFT Study of CO2 Reduction Reaction to CH3OH on Low-Index Cu Surfaces. Catalysts 2023, 13, 722. https://doi.org/10.3390/catal13040722

AMA Style

Xue Q, Qi X, Li K, Zeng Y, Xu F, Zhang K, Yang T, Qi X, Jiang J. DFT Study of CO2 Reduction Reaction to CH3OH on Low-Index Cu Surfaces. Catalysts. 2023; 13(4):722. https://doi.org/10.3390/catal13040722

Chicago/Turabian Style

Xue, Qian, Xuede Qi, Kun Li, Yi Zeng, Feng Xu, Kai Zhang, Tingting Yang, Xueqiang Qi, and Jinxia Jiang. 2023. "DFT Study of CO2 Reduction Reaction to CH3OH on Low-Index Cu Surfaces" Catalysts 13, no. 4: 722. https://doi.org/10.3390/catal13040722

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