Next Article in Journal
Amyloid-β Tetramers and Divalent Cations at the Membrane/Water Interface: Simple Models Support a Functional Role
Previous Article in Journal
Genetic Mechanisms of Migraine: Insights from Monogenic Migraine Mutations
Previous Article in Special Issue
Towards Machine Learning in Heterogeneous Catalysis—A Case Study of 2,4-Dinitrotoluene Hydrogenation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Al-Doped Octahedral Cu2O Nanocrystal for Electrocatalytic CO2 Reduction to Produce Ethylene

State Key Laboratory of High-Efficiency Utilization of Coal and Green Chemical Engineering, College of Chemistry and Chemical Engineering, Ningxia University, Yinchuan 750021, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(16), 12680; https://doi.org/10.3390/ijms241612680
Submission received: 9 July 2023 / Revised: 4 August 2023 / Accepted: 8 August 2023 / Published: 11 August 2023
(This article belongs to the Special Issue Nanoparticle for Catalysis—2nd Edition)

Abstract

:
Ethylene is an ideal CO2 product in an electrocatalytic CO2 reduction reaction (CO2RR) with high economic value. This paper synthesised Al-doped octahedral Cu2O (Al–Cu2O) nanocrystal by a simple wet chemical method. The selectivity of CO2RR products was improved by doping Al onto the surface of octahedral Cu2O. The Al–Cu2O was used as an efficient electrocatalyst for CO2RR with selective ethylene production. The Al–Cu2O exhibited a high % Faradic efficiency (FEC2H4) of 44.9% at −1.23 V (vs. RHE) in CO2 saturated 0.1 M KHCO3 electrolyte. Charge transfer from the Al atom to the Cu atom occurs after Al doping in Cu2O, optimizing the electronic structure and facilitating CO2RR to ethylene production. The DFT calculation showed that the Al–Cu2O catalyst could effectively reduce the adsorption energy of the *CHCOH intermediate and promote the mass transfer of charges, thus improving the FEC2H4. After Al doping into Cu2O, the center of d orbitals shift positively, which makes the d–band closer to the Fermi level. Furthermore, the density of electronic states increases due to the interaction between Cu atoms and intermediates, thus accelerating the electrochemical CO2 reduction process. This work proved that the metal doping strategy can effectively improve the catalytic properties of Cu2O, thus providing a useful way for CO2 cycling and green production of C2H4.

1. Introduction

The increased CO2 emissions in the atmosphere result in a serious greenhouse effect and the elevated sea level [1,2]. The electrochemical CO2 reduction reaction (CO2RR) is a promising strategy to mitigate the global warming and energy crisis while transforming CO2 into fuels and chemical feedstocks [3,4,5]. It can use clean electric energy generated by renewable solar and wind energy to drive the conversion of CO2 under mild conditions [6,7]. Electroreduction of CO2 in molten salts, also called molten salt carbon capture and electrochemical transformation (MSCC-ET), is an advanced method which can capture carbon dioxide from the atmosphere or flue gases [8,9,10,11]. The reduction products of CO2RR include CO [12,13,14], HCOOH [15,16,17,18], alcohols [19,20,21], and various hydrocarbons [22,23,24]. Among them, C2H4 has attracted more and more attention due to its high energy density. Furthermore, C2H4 is essential in producing various plastics, solvents, and cosmetics in the chemical industry [25].
Up to now, Cu-based materials can electrocatalytic convert CO2 into C2/C2+ products. Among reported Cu-based catalysts, Cu2O nanocrystal has attracted much attention due to their electrocatalytic activity and high selectivity toward C2H4. To improve the CO2RR performance of Cu2O, great efforts have been made on its structural design. Metal ions can be used as structure-guiding agents to optimize the micro-structure of nanocrystals [26]. Cu2O nanoparticles (Cu2O NPs) exhibit good performance for CO2RR, possibly because the low coordination Cu+ on the surface contributes to C-C coupling, thus promoting the production of C2H4 [27]. Other strategies, including crystal facet controlling [28], defect engineering [29,30], alloying [31], valence adjustment [32], and surface molecular modification [24], have been employed to improve the electrocatalytic performance of CO2RR to produce C2H4. For example, Shang et al. [33] have designed a core-shell Cu@Cu2O catalyst on which a thin layer of natural oxide grows on the surface under environmental conditions. The synergistic effect between Cu+ and Cu0 on the Cu@Cu2O surface helps to improve its efficiency and selectivity for C2 products. Ning et al. [34] reported the preparation of Cu2O/nitrogen-doped carbon shell (Cu2O/NCS) composite and its application in CO2 electroreduction to selective formation of C2H4. However, copper-based catalysts still face many problems, such as inevitable competitive hydrogen evolution reaction (HER), complex reaction mechanisms diversification of products, and low selectivity of target products [35]. Therefore, it is of huge challenge to design CO2RR electrocatalysts with high activity, selectivity, and satisfied stability.
This work proposed an effective strategy to improve the CO2RR activity by doping Al on the surface of octahedron Cu2O nanocrystals. Al-doped Cu2O (Al–Cu2O) was used as an effective electrocatalyst for selective ethylene production by CO2RR. Al–Cu2O exhibits a high Faraday efficiency (FEC2H4) of 44.9% at −1.23 V (vs. RHE). During the coupling process of *CHCOH intermediate, the Al–Cu2O catalyst can significantly reduce the free energy and promote the formation of C2H4. It can also inhibit the occurrence of HER side reaction. Therefore, the doping strategy is beneficial for the adsorption of intermediates to reconstruct the internal stable state of Cu2O, thus improving the activity and selectivity of CO2 conversion to ethylene.

2. Results and Discussion

2.1. Morphology and Structure Analysis

As shown in Figure 1, Al–Cu2O–X nanocrystals were prepared by a simple one-step method (Experimental section for details). To further characterize Al–Cu2O–X catalysts, the XRD pattern was used to study Cu2O, Al–Cu2O, and Al–Cu2O–2 crystal structure. From the XRD pattern in Figure 2a, the peaks at 29°, 36°, 42°, 61°, 73°, and 77° correspond to the (110), (111), (200), (220), (311), and (222) planes of Cu2O, respectively, which agree well with the octahedral Cu2O (PDF#75–1535). The XRD patterns of Al–Cu2O–X (3-6) are shown in Figure S1. During the preparation process, the catalyst was synthesized by adjusting the amount of Al3+, the concentration of NaOH, and the reaction time. Al–Cu2O–X (3-6) were all single-phase Cu2O nanocrystals. The morphologies of the Cu2O, Al–Cu2O, and Al–Cu2O–2 were monitored by scanning electron microscopy (SEM). Cu2O nanocrystals without Al doping showed an octahedron shape with a smooth surface (Figure S2a). Due to the doping effect of Al, the Al–Cu2O nanocrystal presented an octahedral shape with a more rough surface and formed a defect structure (Figure 2b), which may provide abundant active sites for CO2RR [36]. When the concentration of Al3+ increased from 0.02 M to 0.03 M, Al–Cu2O–2 catalyst exhibits a cube shape (Figure S2b). However, it was reported that the resulting cube Al–Cu2O catalyst is not conducive to forming C2H4 [27]. The better-performing Al–Cu2O with a homo-octahedral shape was observed by TEM (Figure 2c), which was consistent with the SEM image. The high-resolution transmission electron microscopy (HRTEM) image in Figure 2d presented that the lattice stripe spacing d marked was 0.304 nm, corresponding to the (110) crystal plane of Cu2O. The HAADF-STEM image (Figure 2e) also exhibited an octahedral shape. The composition of Al–Cu2O was reconfirmed by elemental mapping (Figure 2f). The Al (red), Cu (blue), and O (green) elements are uniformly distributed over the Al–Cu2O nanocrystals.
The surface composition and valence of Cu2O and Al–Cu2O nanocrystals were characterized by X-ray photoelectron spectroscopy (XPS). As shown in Figure 3a,b, four peaks were observed in Cu 2p spectrum for both Cu2O and Al–Cu2O samples. For Cu2O, the peaks at 932.78 and 952.62 eV corresponded to the binding energies of Cu 2p3/2 and Cu 2p1/2 of Cu2O or Cu, respectively. The binding energies at 935.28 eV and 944.48 eV were ascribed to the peaks of Cu2+. For Al–Cu2O, the 932.89 and 952.73 eV peaks corresponded to the Cu 2p3/2 and Cu2 p1/2 of Cu2O or Cu, respectively. The binding energy of 935.26 and 944.46 eV belonged to the peak of Cu2+. The above results showed that the existence of Cu0 may be due to the partial reduction of Cu2O in the CO2RR process [37]. The existence of trace CuO may be mainly due to the oxidation of a small amount of Cu2O catalyst to CuO in the air after the synthesis of Cu2O [38]. When octahedral Cu2O nanocrystals were doped with Al, the peaks of Cu 2p3/2, Cu 2p1/2, and Cu2+ of Cu2O or Cu were shifted positively. These results may be attributed to the introduction of Al, which can induce charge transfer from Al atoms to Cu atoms, thus modulating the electronic structure of Al–Cu2O. The existence of Cu2O was also confirmed in the O 1s XPS spectra of Cu2O and Al–Cu2O (Figure 3c,d). There were three XPS peaks in both catalysts, of which the peak at 530.5 eV corresponded to the Cu-O bond, and 532.11 and 532.77 eV corresponded to Olat and C=O, respectively [39]. In the high-resolution spectrum of Al 2p (Figure 3e), the peaks at 74.55 and 77.35 eV corresponded to the Al 2p1/2 and Al 2p3/2 of metal Al, respectively. The Al atom was 0.41% by XPS analysis, indicating that the Al–Cu2O catalyst has been successfully prepared.

2.2. Electrocatalytic CO2RR Performances

To further analyze the electrochemical performance of the catalyst, the linear sweep voltammetry (LSV) of Cu2O and Al–Cu2O–X in saturated CO2 electrolyte and saturated N2 electrolyte were tested. The analysis of Figure 4a shows that the current density of Al–Cu2O catalyst in CO2 saturated electrolyte is higher than that in N2, indicating that Al–Cu2O catalyst had higher activity to CO2RR. The LSV curve was measured in a CO2-saturated 0.1 M KHCO3 electrolyte (Figure S3a). The current density of the Al–Cu2O catalyst in CO2 saturated electrolyte was higher than that of Cu2O and Al–Cu2O–2 catalysts, indicating that the Al–Cu2O catalyst had better electrocatalytic activity to CO2RR. Figure S3b shows the potentiostatic electrolysis of CO2 at various potentials. The almost constant current signal indicated that the Al–Cu2O catalyst exhibited good electrochemical stability during the CO2RR process. In Figure 4b, the formation rates of three kinds of catalysts were presented for ethylene products. The Al–Cu2O catalyst had a higher current density for ethylene formation than that of Cu2O and the Al–Cu2O–2 catalysts in a wide potential range. The partial current density of 16.7 mA cm−2 was achieved at −1.38 V (vs. RHE). The above results showed that the Al–Cu2O catalyst was more conducive to producing ethylene as the main product and has a better inhibitory effect on competition for hydrogen formation.
To determine the CO2RR selectivity of the Al–Cu2O catalyst, the reduction products were qualitatively and quantitatively analyzed. In this study, the reduction products of each catalyst were determined in the wide potential range from −0.98 V to −1.38 V (vs. RHE). From Figure 4c and Figure 5, the CO2RR products by Cu2O, and Al–Cu2O–X catalysts were C2H4, HCOO, CO, CH4 and by-product H2. Figure 5a shows the FE of the electrochemical CO2RR product catalyzed by an octahedral Cu2O nanocrystal catalyst without the Al doping. The octahedral Cu2O nanocrystal catalyst had a good effect on inhibiting HER at low potential, and the FEC2H4 was 26.1%. As shown in Figure 4b, with the potential increase, the FE value of H2 decreases from 35.1% to 22.1%.
On the contrary, the FE value of C2H4 increases to 44.9% at −1.23 V (vs. RHE). The results showed the catalyst’s good selectivity for ethylene production and inhibition effect on HER. If an appropriate amount of Al (0.02 M) was introduced into the octahedral Cu2O nanocrystal (Figure 4c), the selectivity of the Al–Cu2O catalyst was improved. If more Al3+ was added to the reaction, The result showed that the FE of C2H4 was 32.8%, indicating that the catalyst had a good selectivity for ethylene (Figure 5b). we also studied the effects of catalysts synthetic conditions, including reaction time (Figure 5c,d) and NaOH concentration (Figure 5e,f), on the selectivity of the CO2RR product. The results indicated that optimising reaction time and NaOH concentration could give the catalyst a certain selectivity. Figure 4d compares the selectivity of three kinds of catalysts (Cu2O, Al–Cu2O, and Al–Cu2O–2) for ethylene products. Under different applied potentials, the efficiency of the Al–Cu2O catalyst for CO2RR to C2H4 was higher than that of the other two catalysts. This result suggested that the Al introduced into the catalyst affected the selectivity of the catalyst. This may be because Al-doped Cu2O will cause changes in the electronic structure and the morphology of the catalyst, thus reducing the adsorption energy of the catalyst for ethylene intermediates in the CO2RR process and enhancing the selectivity of the reaction to the products. It was worth noting that the Faradaic efficiencies sometimes do not reach 100%. A small number of liquid products may still be produced in the electrocatalysis process.
The electrochemical surface area (ECSA) is also a key point for the electrocatalyst. According to the formula for calculating ECSA, it is known that this parameter is related to the Cdl and Cds values of their catalysts because the catalysts are coated on hydrophobic carbon paper (model 060). Therefore, the Cds of the three catalysts are the same, and only the Cdl value of the catalyst can be calculated to determine the ECSA of the catalyst. According the cyclic voltammograms (Figure S4a–c) of Cu2O, Al–Cu2O, and Al–Cu2O–2 catalysts at different scanning rates (20, 40, 60, 80,100, 120 mV s−1), the capacitance values of Cu2O, Al–Cu2O, and Al–Cu2O–2 catalysts were 0.109, 0.122, and 0.076 mF cm−2, respectively, as shown in Figure 4e. The largest Cdl of the Al–Cu2O electrocatalyst suggested the high electrochemical activity surface area of the Al–Cu2O–2 catalyst. This high ECSA can offer a lot of catalytic active sites for improving the electrocatalytic performance of CO2RR, which was consistent with the previous research conclusion.
The impedance of Cu2O and Al–Cu2O catalysts under open-circuit voltage was obtained (Figure S5). The EIS arc of the Al–Cu2O catalyst was smaller than that of the Cu2O catalyst. The results indicate that interface charges can be rapidly transferred during the reaction process, and catalytic activity can be improved. To better understand the activity and kinetics of Al–Cu2O materials on CO2RR, the Tafel slope analysis of the local current density of the catalyst product was carried out. As shown in Figure 4f, the Tafel slope of the Al–Cu2O catalyst (74.3 mV dec−1) was lower than that of Cu2O (85.9 mV dec−1) and the Al–Cu2O–2 (110.4 mV dec−1), indicating that the electron transfer rate of the catalyst is faster, which was beneficial to the rapid adsorption and desorption of the important intermediate from the surface of Al–Cu2O catalyst.
The stability of the Al–Cu2O catalyst was investigated in the CO2RR process. As seen in Figure 6a, the current density of the Al–Cu2O catalyst can be kept stable, and the FE of ethylene can be kept above 40% in the first 3600 s. With the change in reaction time, the current density increases gradually. However, the selectivity of the catalyst to ethylene decreased obviously after two hours of electrolysis. This may be because of the shedding of the catalyst in the long-term electrolysis process, resulting in a decrease in the FE of the catalyst. The stability of copper-based catalysts is poor. Therefore, other strategies must be used to improve the stability of copper-based catalysts for a long time [40]. The XRD pattern after long-term electrolysis showed that the Al–Cu2O showed good structure stability (Figure 6b) in the whole CO2RR test. After the electrolysis of the Al–Cu2O catalyst for 10 min, 20 min, 30 min and 7 h (Figure S6a–d), the morphology of the octahedron remains unchanged. With the increase of electrolysis time, some small pores appear on the catalyst’s surface. The appearance of these pores may provide more active sites, resulting in an increase in current density in the electrolysis process. However, it yielded a decrease in the FE of ethylene. The above results show that the catalyst can maintain stability under long-term electrolysis.

2.3. DFT Computations

We used Density functional theory (DFT) to further calculate, simulate and compare the CO2RR reaction path on the surface of Al–Cu2O and Cu2O catalysts to understand the path from CO2 to C2H4. Figure 7 shows the spatial structure (Figure 7a) and energy distribution of Al–Cu2O and Cu2O. Figure 7b shows the energy distribution of ethylene production and by-product H2 of Cu2O and Al–Cu2O catalysts. The Gibbs free energies of each intermediate along ethylene on Cu2O and Al–Cu2O catalysts *CHCOH, *CCH, *CCH, *CCH2, *CHCH2 (intermediates for ethylene production) and *H (intermediates to H2) have been calculated. Because the Gibbs free energy of the Al–Cu2O catalyst was lower than that of the Cu2O catalyst in each reaction step, the path of ethylene production of CO2RR was easier to occur. It can be seen that the strategy of doping Al to octahedral Cu2O was beneficial in improving the selectivity of product C2H4 [41]. At the same time, the analysis of Figure 7c showed that the Al–Cu2O catalyst with Al doping enhanced the adsorption of intermediate *H and further departed from the ideal hydrogen adsorption value (0 eV). It makes the competitive reaction of HER more disadvantageous. To further analyze the potential reason for the selective improvement of this product, the density of states (DOS) of d orbitals on Cu2O (001) and Al–Cu2O (001) surfaces before CHCOH adsorption was compared (Figure 7d,e). Since the electronic states near the Fermi level are mainly contributed by the d electrons of Cu atoms, it is observed that the reaction is mainly caused by the interaction between Cu and C, and the d band center of octahedron Cu2O (001) was −2.087 eV. The −2.027 eV of the Al–Cu2O (001) surface was closer to the Fermi level (0 eV), and the d-band shifts upward on the Abscissa, which makes the center of the d-band closer to the Fermi level and increases the density of electronic states. So, the adsorption of Cu atoms through d electrons and intermediates was facilitated, thus promoting the CO2RR process and improving the selectivity of the catalyst for the C2H4 product.

3. Materials and Methods

3.1. Preparation of Al–Cu2O Nanocrystals

The Al–Cu2O nanocrystals were synthesized with an improved method according to the literature [42]. The specific step was as follows: 10 mL of 0.6 M NaOH aqueous solution was first added to the sample bottle. Subsequently, a certain amount of CuCl2·2H2O, Al(NO3)3·9H2O and glucose were added to the sample bottle successively. The concentrations of CuCl2·2H2O, Al(NO3)3·9H2O, and glucose were 0.10 M, 0.02 M, and 0.07 M, respectively. After continuous agitation for 5 min, the sample bottle was placed in a 70 °C water bath and vigorously stirred for 4 min. The precipitation obtained by centrifugal collection was rinsed with deionized water and dried under vacuum at room temperature for 12 h to obtain an Al–Cu2O catalyst. At the same time, the effects of the amount of Al3+, NaOH concentration, and the sythiestic reaction time on ethylene products were also investigated in this chapter, and the optimum preparation conditions were obtained, as shown in the following Table S1.

3.2. Preparation of Al–Cu2O Coated Carbon Paper Electrode

5 mg of the prepared catalyst was added to 25 μL of Nafion. Then, 300 μL of distilled water and 175 μL of ethanol to prepare 500 μL of reagent was added and mixed by sonication for 2 h. Subsequently, 100 μL was uniformly applied with a pipette to a carbon paper (type 060) with a total surface area of 1 cm2. The loading on the carbon paper was calculated to be 1 mg cm−2 and dried in a vacuum oven to obtain the Al–Cu2O electrode for the next test.

3.3. Electrochemical Measurements

The electrocatalytic CO2RR was carried out in an H-type electrolytic cell with a proton exchange membrane (Nafion 117) separation. A platinum sheet (1 cm2) as the counter electrode and Ag/AgCl (saturated KCl) as the reference electrode, respectively. Before conducting the experimental test, CO2 (99.999% purity) or N2 gas was introduced into the electrolytic cell, which was saturated with 0.1 M KHCO3 (pH = 6.8) electrolyte after approximately 30 min. In this work, all electrochemical performance was measured on the electrochemical workstation (CHI760E, Shanghai Chenhua, Shanghai, China). All electrode potentials were converted into electrode potentials relative to RHE through the Nernst equation: E   ( vs .   RHE ) =   E   ( vs .   Ag / AgCl ) + 0.0591 × pH + 0.197   V . The electrochemical active surface area was tested by the cyclic voltammetry curves of the bilayer capacitance values at different scanning rates (20, 40, 60, 80, 100 and 120 mV s−1). The gaseous products were collected by electrolysis of the four catalysts in a 0.1 M KHCO3 electrolyte saturated with CO2 for 10 min at different measurement potentials and then analyzed using gas chromatography (8890, Agilent, Santa Clara, CA, USA). The liquid products of the four catalysts were collected by electrolysis in an aqueous 0.1 M KHCO3 solution saturated with CO2 for 30 min at each measurement potential, followed by qualitative and quantitative analysis using ion chromatography (AS-DV, Thermo Scientific, Waltham, MA, USA).

3.4. Product Analysis

The gas products are detected by gas chromatography (GC, Agilent 8890) directly from the gas outlet. The carbonaceous gas products from the cathode chamber are analyzed by a methane reformer and flame ionization detector (FID). A thermal conductivity detector (TCD) was used to detect the eCO2RR by-product H2. When the current stabilizes, the gas product is detected. Quantification of the gaseous products was determined by comparison with the standard curve. the Faraday efficiency (FE) of C2H4, H2 and CO was calculated as follows:
FE = N × n × v × F V m × j × 100 %
where   v is the CO2 flow rate (v = 20 mL min−1), n is the total molar fraction of C2H4, H2 or CO of the gas measured in the GC, N is the number of electrons required to form a molecule of H2 or CO (N = 2), F is Faraday’s constant (96,485 C mol−1), and Vm is the molar volume of the gas at 298 K and j is current at each potential (A).
Liquid products Faraday efficiency test method: A saturated solution of electrocatalytic CO2 was electro-catalyzed by the Coulomb method using a controlled potential, and the electrolytic reduction product was analyzed and calculated after 0.5 h. The CO2 flow rate during electrolysis was controlled at 20 mL min−1, and the liquid product was determined by ion chromatography (AS-DV, Thermo Scientific, Waltham, MA, USA). The FE of the liquid phase product was calculated as follows:
FE = N n F Q × 100 %
where N is the number of electrons transferred, n is the amount of formate in the cathode chamber, F is Faraday’s constant (96,485 C mol−1) and Q is the total charge passing through the electrode.

4. Conclusions

In summary, the Al-doped octahedral Cu2O nanocrystal was successfully prepared and used as an efficient CO2RR electrocatalyst. The Al–Cu2O exhibited high activity and selectivity for ethylene production. The Al–Cu2O catalyst demonstrates a high % faradaic efficiency of 44.9% at −1.23 V (vs. RHE) for C2H4 production. The high catalytic activity for CO2 electrochemical reduction is due to the optimized electronic state by Al doping in octahedral Cu2O nanocrystals. The DFT simulation suggested the C–C coupling mechanism in the electrochemical CO2RR process. The Al–Cu2O doped Cu2O octahedron can significantly reduce the free energy in the coupling process of *CHCOH intermediate, promote the formation of C2H4, and inhibit the occurrence of HER side effect. Furthermore, our work demonstrates a simple doping strategy for preparing copper-based catalysts, which can be extended to the design and study of other highly efficient electrocatalysts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms241612680/s1. References [43,44,45,46] are cited in the supplementary materials.

Author Contributions

Conceptualization, J.P.; methodology, S.L. and X.S.; formal analysis, S.L., investigation, S.L., X.G. and X.S.; resources, J.P.; data curation, S.L. and X.S.; writing—original draft preparation, S.L., X.S. and J.P.; writing—review and editing, S.L., X.S. and J.P.; visualization, S.L., X.G. and X.S.; project administration, J.P.; funding acquisition, J.P. 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 (No. 22262027, 22132003), Ningxia leading scientific and technological innovation talents projects (No. KJT2018002) and Ningxia Natural Science Foundation (No. 2022AAC03103).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data in this study can be found in public data bases and Supplementary Information, as described in the Material and Methods section (Section 3).

Conflicts of Interest

The authors declared no competing financial interest.

References

  1. Alli, Y.A.; Oladoye, P.O.; Ejeromedoghene, O.; Bankole, O.M.; Alimi, O.A.; Omotola, E.O.; Olanrewaju, C.A.; Philippot, K.; Adeleye, A.S.; Ogunlaja, A.S. Nanomaterials as catalysts for CO2 transformation into value-added products: A review. Sci. Total Environ. 2023, 868, 161547. [Google Scholar] [CrossRef] [PubMed]
  2. Zhao, M.; Gu, Y.; Gao, W.; Cui, P.; Tang, H.; Wei, X.; Zhu, H.; Li, G.; Yan, S.; Zhang, X.; et al. Atom vacancies induced electron-rich surface of ultrathin Bi nanosheet for efficient electrochemical CO2 reduction. Appl. Catal. B Environ. 2020, 266, 118625. [Google Scholar] [CrossRef]
  3. Feng, X.; Zou, H.; Zheng, R.; Wei, W.; Wang, R.; Zou, W.; Lim, G.; Hong, J.; Duan, L.; Chen, H. Bi2O3/BiO2 Nanoheterojunction for Highly Efficient Electrocatalytic CO2 Reduction to Formate. Nano Lett. 2022, 22, 1656–1664. [Google Scholar] [CrossRef] [PubMed]
  4. Yang, Y.; He, A.; Yang, M.; Zou, Q.; Li, H.; Liu, Z.; Tao, C.; Du, J. Selective electroreduction of CO2 to ethanol over a highly stable catalyst derived from polyaniline/CuBi2O4. Catal. Sci. Technol. 2021, 11, 5908–5916. [Google Scholar] [CrossRef]
  5. Sakamoto, N.; Nishimura, Y.F.; Nonaka, T.; Ohashi, M.; Ishida, N.; Kitazumi, K.; Kato, Y.; Sekizawa, K.; Morikawa, T.; Arai, T. Self-assembled Cuprous Coordination Polymer as a Catalyst for CO2 Electrochemical Reduction into C2 Products. ACS Catal. 2020, 10, 10412–10419. [Google Scholar] [CrossRef]
  6. Liu, B.; Yao, X.; Zhang, Z.; Li, C.; Zhang, J.; Wang, P.; Zhao, J.; Guo, Y.; Sun, J.; Zhao, C. Synthesis of Cu2O Nanostructures with Tunable Crystal Facets for Electrochemical CO2 Reduction to Alcohols. ACS Appl. Mater. Interfaces 2021, 13, 39165–39177. [Google Scholar] [CrossRef]
  7. Zhou, X.; Shan, J.; Chen, L.; Xia, B.Y.; Ling, T.; Duan, J.; Jiao, Y.; Zheng, Y.; Qiao, S.Z. Stabilizing Cu2+ Ions by Solid Solutions to Promote CO2 Electroreduction to Methane. J. Am. Chem. Soc. 2022, 144, 2079–2084. [Google Scholar] [CrossRef]
  8. Jiang, R.; Gao, M.; Mao, X.; Wang, D. Advancements and potentials of molten salt CO2 capture and electrochemical transformation (MSCC-ET) process. Curr. Opin. Electrochem. 2019, 17, 38–46. [Google Scholar] [CrossRef]
  9. Küngas, R. Review—Electrochemical CO2 Reduction for CO Production: Comparison of Low- and High-Temperature Electrolysis Technologies. J. Electrochem. Soc. 2020, 167, 044508. [Google Scholar] [CrossRef]
  10. Lacarbonara, G.; Chini, S.; Ratso, S.; Kruusenberg, I.; Arbizzani, C. A MnOx–graphitic carbon composite from CO2 for sustainable Li-ion battery anodes. Mater. Adv. 2022, 3, 7087–7097. [Google Scholar] [CrossRef]
  11. Remmel, A.-L.; Ratso, S.; Divitini, G.; Danilson, M.; Mikli, V.; Uibu, M.; Aruväli, J.; Kruusenberg, I. Nickel and Nitrogen-Doped Bifunctional ORR and HER Electrocatalysts Derived from CO2. ACS Sustain. Chem. Eng. 2021, 10, 134–145. [Google Scholar] [CrossRef]
  12. Dong, W.; Zhang, N.; Li, S.; Min, S.; Peng, J.; Liu, W.; Zhan, D.; Bai, H. A Mn single atom catalyst with Mn–N2O2 sites integrated into carbon nanosheets for efficient electrocatalytic CO2 reduction. J. Mater. Chem. A 2022, 10, 10892–10901. [Google Scholar] [CrossRef]
  13. Clark, E.L.; Ringe, S.; Tang, M.; Walton, A.; Hahn, C.; Jaramillo, T.F.; Chan, K.; Bell, A.T. Influence of Atomic Surface Structure on the Activity of Ag for the Electrochemical Reduction of CO2 to CO. ACS Catal. 2019, 9, 4006–4014. [Google Scholar] [CrossRef] [Green Version]
  14. Wang, Q.; Liu, K.; Hu, K.; Cai, C.; Li, H.; Li, H.; Herran, M.; Lu, Y.-R.; Chan, T.-S.; Ma, C. Attenuating metal-substrate conjugation in atomically dispersed nickel catalysts for electroreduction of CO2 to CO. Nat. Commun. 2022, 13, 6082. [Google Scholar] [CrossRef] [PubMed]
  15. Duan, Y.X.; Zhou, Y.T.; Yu, Z.; Liu, D.X.; Wen, Z.; Yan, J.M.; Jiang, Q. Boosting production of HCOOH from CO2 electroreduction via Bi/CeOx. Angew. Chem. Int. Ed. 2021, 60, 8798–8802. [Google Scholar] [CrossRef]
  16. Koh, J.H.; Won, D.H.; Eom, T.; Kim, N.-K.; Jung, K.D.; Kim, H.; Hwang, Y.J.; Min, B.K. Facile CO2 electro-reduction to formate via oxygen bidentate intermediate stabilized by high-index planes of Bi dendrite catalyst. ACS Catal. 2017, 7, 5071–5077. [Google Scholar] [CrossRef]
  17. Li, D.; Huang, L.; Tian, Y.; Liu, T.; Zhen, L.; Feng, Y. Facile synthesis of porous Cu-Sn alloy electrode with prior selectivity of formate in a wide potential range for CO2 electrochemical reduction. Appl. Catal. B Environ. 2021, 292, 120119. [Google Scholar] [CrossRef]
  18. Grace, A.N.; Choi, S.Y.; Vinoba, M.; Bhagiyalakshmi, M.; Chu, D.H.; Yoon, Y.; Nam, S.C.; Jeong, S.K. Electrochemical reduction of carbon dioxide at low overpotential on a polyaniline/Cu2O nanocomposite based electrode. Appl. Energy 2014, 120, 85–94. [Google Scholar] [CrossRef]
  19. Boutin, E.; Wang, M.; Lin, J.C.; Mesnage, M.; Mendoza, D.; Lassalle-Kaiser, B.; Hahn, C.; Jaramillo, T.F.; Robert, M. Aqueous electrochemical reduction of carbon dioxide and carbon monoxide into methanol with cobalt phthalocyanine. Angew. Chem. Int. Ed. 2019, 58, 16172–16176. [Google Scholar] [CrossRef]
  20. Ren, D.; Deng, Y.; Handoko, A.D.; Chen, C.S.; Malkhandi, S.; Yeo, B.S. Selective electrochemical reduction of carbon dioxide to ethylene and ethanol on copper (I) oxide catalysts. ACS Catal. 2015, 5, 2814–2821. [Google Scholar] [CrossRef]
  21. Yuan, J.; Yang, M.-P.; Zhi, W.-Y.; Wang, H.; Wang, H.; Lu, J.-X. Efficient electrochemical reduction of CO2 to ethanol on Cu nanoparticles decorated on N-doped graphene oxide catalysts. J. CO2 Util. 2019, 33, 452–460. [Google Scholar] [CrossRef]
  22. Iyengar, P.; Huang, J.; De Gregorio, G.L.; Gadiyar, C.; Buonsanti, R. Size dependent selectivity of Cu nano-octahedra catalysts for the electrochemical reduction of CO2 to CH4. Chem. Commun. 2019, 55, 8796–8799. [Google Scholar] [CrossRef]
  23. Chu, S.; Kang, C.; Park, W.; Han, Y.; Hong, S.; Hao, L.; Zhang, H.; Lo, T.W.B.; Robertson, A.W.; Jung, Y. Single atom and defect engineering of CuO for efficient electrochemical reduction of CO2 to C2H4. SmartMat 2022, 3, 194–205. [Google Scholar] [CrossRef]
  24. Luo, H.; Li, B.; Ma, J.G.; Cheng, P. Surface Modification of Nano-Cu2O for Controlling CO2 Electrochemical Reduction to Ethylene and Syngas. Angew. Chem. Int. Ed. 2022, 61, e202116736. [Google Scholar] [CrossRef]
  25. Tan, X.; Yu, C.; Zhao, C.; Huang, H.; Yao, X.; Han, X.; Guo, W.; Cui, S.; Huang, H.; Qiu, J. Interfaces, Restructuring of Cu2O to Cu2O@Cu-metal–organic frameworks for selective electrochemical reduction of CO2. ACS Appl. Mater. Interfaces 2019, 11, 9904–9910. [Google Scholar] [CrossRef]
  26. Cao, S.; Chen, H.; Han, T.; Zhao, C.; Peng, L. Rose-like Cu2O nanoflowers via hydrothermal synthesis and their gas sensing properties. Mater. Lett. 2016, 180, 135–139. [Google Scholar] [CrossRef]
  27. Gao, Y.; Wu, Q.; Liang, X.; Wang, Z.; Zheng, Z.; Wang, P.; Liu, Y.; Dai, Y.; Whangbo, M.H.; Huang, B. Cu2O Nanoparticles with Both 100 and 111 Facets for Enhancing the Selectivity and Activity of CO2 Electroreduction to Ethylene. Adv. Sci. 2020, 7, 1902820. [Google Scholar] [CrossRef] [Green Version]
  28. Fu, W.; Liu, Z.; Wang, T.; Liang, J.; Duan, S.; Xie, L.; Han, J.; Li, Q. Engineering, Promoting C2+ production from electrochemical CO2 reduction on shape-controlled cuprous oxide nanocrystals with high-index facets. ACS Sustain. Chem. Eng. 2020, 8, 15223–15229. [Google Scholar] [CrossRef]
  29. Luo, T.; Liu, K.; Fu, J.; Chen, S.; Li, H.; Hu, J.; Liu, M. Tandem catalysis on adjacent active motifs of copper grain boundary for efficient CO2 electroreduction toward C2 products. J. Energy Chem. 2022, 70, 219–223. [Google Scholar] [CrossRef]
  30. Wang, Y.; Lou, Z.; Niu, W.; Ye, Z.; Zhu, L. Optimization of photoelectrochemical performance in Pt-modified p-Cu2O/n-Cu2O nanocomposite. Nanotechnology 2018, 29, 145402. [Google Scholar] [CrossRef]
  31. Feng, Y.; Li, Z.; Liu, H.; Dong, C.; Wang, J.; Kulinich, S.A.; Du, X. Laser-prepared CuZn alloy catalyst for selective electrochemical reduction of CO2 to ethylene. Langmuir 2018, 34, 13544–13549. [Google Scholar] [CrossRef]
  32. Chen, C.; Sun, X.; Yan, X.; Wu, Y.; Liu, M.; Liu, S.; Zhao, Z.; Han, B. A strategy to control the grain boundary density and Cu+/Cu0 ratio of Cu-based catalysts for efficient electroreduction of CO2 to C2 products. Green Chem. 2020, 22, 1572–1576. [Google Scholar] [CrossRef]
  33. Shang, L.; Lv, X.; Shen, H.; Shao, Z.; Zheng, G. Selective carbon dioxide electroreduction to ethylene and ethanol by core-shell copper/cuprous oxide. J. Colloid Interface Sci. 2019, 552, 426–431. [Google Scholar] [CrossRef] [PubMed]
  34. Ning, H.; Wang, X.; Wang, W.; Mao, Q.; Yang, Z.; Zhao, Q.; Song, Y.; Wu, M. Cubic Cu2O on nitrogen-doped carbon shells for electrocatalytic CO2 reduction to C2H4. Carbon 2019, 146, 218–223. [Google Scholar] [CrossRef]
  35. Chang, X.; Wang, T.; Zhao, Z.J.; Yang, P.; Greeley, J.; Mu, R.; Zhang, G.; Gong, Z.; Luo, Z.; Chen, J. Tuning Cu/Cu2O interfaces for the reduction of carbon dioxide to methanol in aqueous solutions. Angew. Chem. Int. Ed. 2018, 57, 15415–15419. [Google Scholar] [CrossRef]
  36. Qiu, Y.-L.; Zhong, H.-X.; Zhang, T.-T.; Xu, W.-B.; Li, X.-F.; Zhang, H.-M. Copper Electrode Fabricated via Pulse Electrodeposition: Toward High Methane Selectivity and Activity for CO2 Electroreduction. ACS Catal. 2017, 7, 6302–6310. [Google Scholar] [CrossRef]
  37. Zheng, Y.; Zhang, L.; Guan, J.; Qian, S.; Zhang, Z.; Ngaw, C.K.; Wan, S.; Wang, S.; Lin, J.; Wang, Y.; et al. Engineering, Controlled synthesis of Cu0/Cu2O for efficient photothermal catalytic conversion of CO2 and H2O. ACS Sustain. Chem. Eng. 2021, 9, 1754–1761. [Google Scholar] [CrossRef]
  38. Wang, S.; Kou, T.; Varley, J.B.; Akhade, S.A.; Weitzner, S.E.; Baker, S.E.; Duoss, E.B.; Li, Y. Cu2O/CuS nanocomposites show excellent selectivity and stability for formate generation via electrochemical reduction of carbon dioxide. ACS Mater. Lett. 2020, 3, 100–109. [Google Scholar] [CrossRef]
  39. Lv, X.-W.; Liu, Y.; Hao, R.; Tian, W.; Yuan, Z.-Y. Urchin-like Al-doped Co3O4 nanospheres rich in surface oxygen vacancies enable efficient ammonia electrosynthesis. ACS Appl. Mater. Interfaces 2020, 12, 17502–17508. [Google Scholar] [CrossRef]
  40. Sun, B.; Dai, M.; Cai, S.; Cheng, H.; Song, K.; Yu, Y.; Hu, H. Challenges and strategies towards copper-based catalysts for enhanced electrochemical CO2 reduction to multi-carbon products. Fuel 2023, 332, 126114. [Google Scholar] [CrossRef]
  41. Li, Y.C.; Wang, Z.; Yuan, T.; Nam, D.-H.; Luo, M.; Wicks, J.; Chen, B.; Li, J.; Li, F.; de Arquer, F.P.G.; et al. Binding Site Diversity Promotes CO2 Electroreduction to Ethanol. J. Am. Chem. Soc. 2019, 141, 8584–8591. [Google Scholar] [CrossRef]
  42. Yang, R.; Wu, Z.; Tang, D.; Xing, Y.; Ren, Y.; Li, F.; Li, X. Synthesis of Cu2O crystals with negative surface curvature at various positions via Al3+ ions inducing. Mater. Lett. 2014, 117, 211–213. [Google Scholar] [CrossRef]
  43. Kresse, G.; Furthmüller, J.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]
  44. Perdew, J.P.; Burke, K.; Ernzerhof, M.J. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [Green Version]
  45. Hammer, B.; Hansen, L.B.; Nørskov, J.K. Improved adsorption energetics within density-functional theory using revised Perdew-Burke-Ernzerhof functionals. Phys. Rev. B 1999, 59, 7413. [Google Scholar] [CrossRef] [Green Version]
  46. Grimme, S.J. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the fabrication process for Al–Cu2O–X (X = 2, 3, 4, 5, 6).
Figure 1. Schematic illustration of the fabrication process for Al–Cu2O–X (X = 2, 3, 4, 5, 6).
Ijms 24 12680 g001
Figure 2. Characterization of Al–Cu2O: (a) XRD, (b) SEM, (c) TEM, (d) HRTEM, (e) HAADF–TEM, and (f) elemental mapping (blue, green and red represents Cu, O and Al element, respectively).
Figure 2. Characterization of Al–Cu2O: (a) XRD, (b) SEM, (c) TEM, (d) HRTEM, (e) HAADF–TEM, and (f) elemental mapping (blue, green and red represents Cu, O and Al element, respectively).
Ijms 24 12680 g002
Figure 3. XPS spectrum of Cu 2p: (a) Cu2O and (b) Al–Cu2O; O 1s spectrum of (c) Cu2O and (d) Al–Cu2O, (e) Al 2p spectrum of the Al–Cu2O.
Figure 3. XPS spectrum of Cu 2p: (a) Cu2O and (b) Al–Cu2O; O 1s spectrum of (c) Cu2O and (d) Al–Cu2O, (e) Al 2p spectrum of the Al–Cu2O.
Ijms 24 12680 g003
Figure 4. (a) LSV curves of Cu2O, Al–Cu2O and Al–Cu2O–2 catalysts in 0.1 M KHCO3 aqueous solutions saturated CO2, (b) partial current density of Cu2O, Al–Cu2O and Al–Cu2O–2 catalysts, sweeping speed of 5 mV s−1, (c) FE values of Al–Cu2O catalyst in 0.1 M KHCO3 aqueous solutions with saturated CO2, (d) The FEC2H4 values of Cu2O, Al–Cu2O and Al–Cu2O–2 catalysts, (e) The linear relationship between ΔJ and scanning rates, (f) Tafel plots of Cu2O, Al–Cu2O and Al–Cu2O–2.
Figure 4. (a) LSV curves of Cu2O, Al–Cu2O and Al–Cu2O–2 catalysts in 0.1 M KHCO3 aqueous solutions saturated CO2, (b) partial current density of Cu2O, Al–Cu2O and Al–Cu2O–2 catalysts, sweeping speed of 5 mV s−1, (c) FE values of Al–Cu2O catalyst in 0.1 M KHCO3 aqueous solutions with saturated CO2, (d) The FEC2H4 values of Cu2O, Al–Cu2O and Al–Cu2O–2 catalysts, (e) The linear relationship between ΔJ and scanning rates, (f) Tafel plots of Cu2O, Al–Cu2O and Al–Cu2O–2.
Ijms 24 12680 g004
Figure 5. (af) FE values of Cu2O and Al–Cu2O–X (=2, 3, 4, 5, 6) catalysts in 0.1 M KHCO3 aqueous solutions with saturated gase CO2.
Figure 5. (af) FE values of Cu2O and Al–Cu2O–X (=2, 3, 4, 5, 6) catalysts in 0.1 M KHCO3 aqueous solutions with saturated gase CO2.
Ijms 24 12680 g005
Figure 6. Al–Cu2O catalyst in 0.1 M KHCO3 electrolyte (a) electrochemical stability test pattern and (b) the XRD of Al–Cu2O catalyst after long–term stability test.
Figure 6. Al–Cu2O catalyst in 0.1 M KHCO3 electrolyte (a) electrochemical stability test pattern and (b) the XRD of Al–Cu2O catalyst after long–term stability test.
Ijms 24 12680 g006
Figure 7. Free energy diagram of ethylene and hydrogen produced by CO2RR on the surface of (a) side and top views of Cu2O (001) and Al–Cu2O (001) configurations, (b) Cu2O (001) and (c) Al–Cu2O (001) catalysts, * reprents the active site; DOS of d orbitals on (d) Cu2O (001) and (e) Al–Cu2O (001) surfaces before *CHCOH adsorption.
Figure 7. Free energy diagram of ethylene and hydrogen produced by CO2RR on the surface of (a) side and top views of Cu2O (001) and Al–Cu2O (001) configurations, (b) Cu2O (001) and (c) Al–Cu2O (001) catalysts, * reprents the active site; DOS of d orbitals on (d) Cu2O (001) and (e) Al–Cu2O (001) surfaces before *CHCOH adsorption.
Ijms 24 12680 g007
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

Li, S.; Sha, X.; Gao, X.; Peng, J. Al-Doped Octahedral Cu2O Nanocrystal for Electrocatalytic CO2 Reduction to Produce Ethylene. Int. J. Mol. Sci. 2023, 24, 12680. https://doi.org/10.3390/ijms241612680

AMA Style

Li S, Sha X, Gao X, Peng J. Al-Doped Octahedral Cu2O Nanocrystal for Electrocatalytic CO2 Reduction to Produce Ethylene. International Journal of Molecular Sciences. 2023; 24(16):12680. https://doi.org/10.3390/ijms241612680

Chicago/Turabian Style

Li, Sanxiu, Xuelan Sha, Xiafei Gao, and Juan Peng. 2023. "Al-Doped Octahedral Cu2O Nanocrystal for Electrocatalytic CO2 Reduction to Produce Ethylene" International Journal of Molecular Sciences 24, no. 16: 12680. https://doi.org/10.3390/ijms241612680

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