Next Article in Journal
In Situ Conditioning of CO2-Rich Syngas during the Synthesis of Methanol
Previous Article in Journal
In Situ Growth of ZIF-8 Nanocrystals on the Pore Walls of 3D Ordered Macroporous TiO2 for a One-Pot Cascade Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Core-Shell ZnO@Cu2O as Catalyst to Enhance the Electrochemical Reduction of Carbon Dioxide to C2 Products

1
Key Laboratory for Green Chemical Technology of the Ministry of Education, School of Chemical Engineering and Technology, Tianjin University, Tianjin 300072, China
2
Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), School of Chemical Engineering and Technology, Tianjin University, Tianjin 300072, China
*
Author to whom correspondence should be addressed.
Catalysts 2021, 11(5), 535; https://doi.org/10.3390/catal11050535
Submission received: 27 March 2021 / Revised: 19 April 2021 / Accepted: 20 April 2021 / Published: 21 April 2021
(This article belongs to the Section Catalytic Materials)

Abstract

:
The copper-based catalyst is considered to be the only catalyst for electrochemical carbon dioxide reduction to produce a variety of hydrocarbons, but its low selectivity and low current density to C2 products restrict its development. Herein, a core-shell xZnO@yCu2O catalysts for electrochemical CO2 reduction was fabricated via a two-step route. The high selectivity of C2 products of 49.8% on ZnO@4Cu2O (ethylene 33.5%, ethanol 16.3%) with an excellent total current density of 140.1 mA cm−2 was achieved over this core-shell structure catalyst in a flow cell, in which the C2 selectivity was twice that of Cu2O. The high electrochemical activity for ECR to C2 products was attributed to the synergetic effects of the ZnO core and Cu2O shell, which not only enhanced the selectivity of the coordinating electron, improved the HER overpotential, and fastened the electron transfer, but also promoted the multielectron involved kinetics for ethylene and ethanol production. This work provides some new insights into the design of highly efficient Cu-based electrocatalysts for enhancing the selectivity of electrochemical CO2 reduction to produce high-value C2 products.

Graphical Abstract

1. Introduction

In recent years, great progress was made in the utilization of carbon dioxide (CO2), including the thermochemical, photochemical, and electrochemical methods. The electrochemical reduction of CO2 (ECR) is now the main research in the current stage due to its great advantages of simple reaction conditions [1,2,3]. Currently, the products obtained from ECR are mainly C1 products [4,5] (such as CO, HCOOH, and CH4) and C2 products (C2H4, C2H6OH, and C2H6, etc.) [6,7,8,9]. However, high overpotential, low product selectivity and low current density are still the main problems for the electrochemical reduction of CO2. Designing novel and efficient catalysts is now an effective way to solve the above problems.
Copper-based catalysts are the only known catalysts that can electrochemically reduce CO2 to a variety of hydrocarbons and oxygen-containing compounds because of their moderate binding energy with CO* [10,11], but the selectivity to products is low. Therefore, the research of high-selectivity copper-based catalysts for ECR is particularly important. To date, some effective strategies for modifying copper-based catalysts were widely investigated, including doping strategy (metal and non-metal) [12], defect engineering [13], oxide-derived copper [14,15], morphology control [16], and size control [17].
Metal oxides or metal oxide derivatives are effective strategies to improve the selectivity of electrochemical CO2 reduction to C2 products. Many works were carried out on Cu2O because Cu2O can strengthen the C–C coupling process for the formation of C2 products. For example, Cu2O films on a Cu disc catalyst was prepared for ECR to improve the selectivity of C2 products. The optimum selectivity of 45% was obtained by adjusting the thickness of the oxide layer, but its partial current density was only 33 mA cm−2 [18]. Li et al. [19] found that the high surface roughness Cu2O catalyst formed at 500 °C would require 0.5 V less overpotential than polycrystalline Cu, to reduce CO2. Furthermore, interface-induced Cu2O nanocubes with controllable morphology and size also exhibited strong selectivity to ethylene (31.1%) in the H-cell at −1.15 V vs. RHE, but the current density was just 3.72 mA cm−2 [20].
Moreover, the active sites and electron distribution of copper catalysts could be tailored by introducing the second element. The synergistic effect of the two elements could strengthen their activity and stability for CO2 electrochemical reduction [21]. A Cu/CNT catalyst was applied for ECR in a flow cell, the main product of CO was obtained but the total current density was 15 mA cm−2 [22]. Ning et al. [23] designed a new type of N-doped Cu2O catalyst using the synergy between Cu2O and pyridinic-N for boosting C2H4 selectivity and durability of Cu2O. The selectivity of ethylene reached 19.7% (current density reached 12 mA cm−2) in an H-type cell. Designing electrocatalysts with dual metal active sites of metal active sites and Cu2O active sites are also an important strategy for enhancing ECR [24]. Gao et al. [25] prepared a bimetallic catalyst of Au uniformly supported on Cu2O. The presence of Au active (an electrocatalyst for CO2 to CO) helped to increase the coverage of *CO on Cu, thereby strengthening the production of C2 products on Cu–Au catalysts. The total Faraday Efficiency of C2 products reached 61.3% under −1.05 V vs. RHE in an H-type gas-tight electrolytic cell, and the current density of 42.87 mA cm−2 was also nearly doubled, as compared to Cu. Similarly, the AgCu foam catalyst also presented a good activity to produce ethanol with 33.7% FE in a glass beaker [26]. Furthermore, a ZnO deposited on CuO nanowire catalysts were synthesized by the atomic layer deposition method. Operando Raman spectroscopy showed that CO* binding on Cu sites was modified by Zn2+. The key intermediate of ethanol products *COCH4 was easily generated to enhance the product of ethanol (FE = 48.6%) with a current density of 98 mA cm−2 in a flow cell [27]. A Cu oxide/ZnO-based catalyst showed high ethylene selectivity in the gas-phase CO2 electro conversion [28]. Hence, the catalyst with dual active sites is an effective strategy for electrochemical CO2 reduction, and it is a promising approach to achieve high selectivity and high current density for C2 products in a flow cell.
Inspired by these works, we designed and synthesized a catalyst with ZnO core and Cu2O shell through a wet chemical method. Due to the double active sites on the ZnO@4Cu2O catalyst. It is possible to make CO2 first generate CO at the ZnO site, and then CO couples on Cu2O, to generate C2 product. The synergistic effect of Cu and Zn enhances the selectivity of C2 products on the catalyst. This core-shell structure was used to enhance the multielectron reactions in electrochemical CO2 reduction in a flow cell. The thickness of the Cu2O shell was controlled to boost the coupling of CO on the Cu site to produce C2 products. By tailoring the synergy of Cu and Zn oxides, a high Faradic efficiency of 49.8% for C2 products at a high current density of 140.1 mA cm−2 was achieved under the potential of −1.0 V vs. RHE. This work exceeded the selectivity of C2 products with many jobs, while achieving a higher current density.

2. Results and Discussion

2.1. The Morphology and Structure of the Catalysts

2.1.1. ZnO Catalysts

The morphology, structure, and chemical state of the as-synthesized ZnO catalysts were characterized. Figure 1a–c shows the scanning electron microscopy (SEM) images of ZnO particles obtained at different calcination temperatures (400 °C, 600 °C, and 800 °C). All samples showed a similar sphere-like morphology. The particle s’ size became bigger as the calcination temperature increased from 400 °C to 800 °C, probably due to the aggregation of particles. The smallest particle size was obtained at 24.4 ± 5 nm for the ZnO-400 sample. The particle size distribution diagrams of ZnO catalyst are displayed in Figure S7.
The structure of ZnO particles was analyzed by XRD characterization. As shown in Figure 1d, all ZnO samples presented almost the same diffraction peaks at 2θ = 31.8°, 34.5°, 36.3°, 47.7°, 56.7°, 63.0°, and 68.1°, which were attributed to (100), (002), (101), (102), (110), (103), and (112) facets of hexagonal wurtzite ZnO (JCPDS No. 75-0576). Moreover, as shown in Figure 1e, the ZnO-400 sample illustrated the highest proportion of (101) and (002) facets, which would have a great impact on the ECR activity. The ratio of facets was calculated by comparing the intensities of the peak in XRD. The composition and chemical valence states of the ZnO samples were analyzed by X-ray photoelectron spectroscopy (XPS) technology. As shown in Figure 1f, the characteristic peaks of Zn 2p1/2 at 1045.63 eV and Zn 2p3/2 at 1022.47 eV were attributed to Zn2+ [29].

2.1.2. xZnO@yCu2O Catalyst

The core-shell xZnO@yCu2O catalysts were synthesized with ZnO-400 as the core, through the epitaxial growth method (Scheme 1). Figure 2a–d displays the morphology and structure of xZnO@yCu2O as well as the Cu2O catalysts. The surface composed of accumulated nano-spherical Cu2O was generated after growing the Cu2O shell, as compared to the pristine ZnO core. This coarse structure generally had a larger external surface area and could provide more active sites [30,31]. The Cu2O sample showed a similar morphology. The particle sizes of the xZnO@yCu2O samples were obviously larger than pristine ZnO particle and gradually grew up as the Cu2O shell thickness increased (101 ± 20 nm for 4ZnO@Cu2O, 129 ± 20 nm for ZnO@4Cu2O, and 227.9 ± 25 nm for ZnO@20Cu2O). Under similar conditions, the particle size of the obtained Cu2O particles was about 181.7 nm. The particle size distribution diagrams are displayed in Figure S8. The TEM image in Figure 2e shows that the Cu2O shell covered the ZnO particles uniformly. The surface composition of the ZnO@4Cu2O sample was analyzed by EDX along the arrow direction. The Cu signal was first monitored, and then the Zn signal was detected when close to the core [32], exhibiting an obvious core-shell distribution. In addition, the signal intensity of the Cu element was significantly stronger than that of the Zn element. Both the 4ZnO@Cu2O and ZnO@20Cu2O catalysts proved that a similar core-shell structure was shown in Figure S1. These results indicated that the core-shell xZnO@yCu2O structure was fabricated. Moreover, the size of the ZnO core was about 30 nm, which was consistent with the ZnO-400. The Cu2O shell thickness was approximately twice that of the ZnO core for the ZnO@4Cu2O catalyst. Furthermore, the element mapping showed that the Cu element was distributed uniformly on the surface of the catalyst, and Zn was mainly distributed in the center. The atomic ratio of Cu to Zn in transmission was 43:1, which was very different from ICP-OES (about 4:1, in Table S1) probably because most of the Zn element inside was not monitored. From Figure 2g, it could be found that three different distinct lattice fringes of the shell were observed. The lattice spacing was 0.30 nm, 0.21 nm, and 0.24 nm, assigned to the (110), (200), and (111) facets of Cu2O, respectively [33]. These results indicated that the cuprous oxide shell was successfully formed on the ZnO core.
Figure 3a shows the XRD patterns of the xZnO@yCu2O catalysts. Compared to the patterns of ZnO-400 (JCPDS:75-0576) and Cu2O particles (JCPDS:78-2076) (Figure S2), the core-shell xZnO@yCu2O showed a mixed one with the characteristic peaks at 2θ = 34.5°, 36.3° corresponding to the (002), (101) facets of ZnO and 2θ = 42.3° corresponding to the (200) facets of Cu2O [34,35], respectively. The peak around 36.5° was obviously strengthened as compared to pristine ZnO, due to the diffraction overlapping of the ZnO (101) facet and Cu2O (111) facet. No peaks assigned to the Cu/Zn alloys were found [36,37]. This indicated that the structure of ZnO was inherited in the xZnO@yCu2O catalysts and the Cu2O (110) shell enclosed with Cu2O (111) and the Cu2O (200) facets. As the binding energies of Cu0 and Cu+ species are very close (only 0.1 eV), the Cu LMM spectrum was applied to distinguish the distribution of the Cu0 and Cu1+ species. No characteristic peaks appeared at the binding energy of 568.0 eV (Figure S3), indicating no presence of Cu0 species. The Cu XPS spectra of the xZnO@yCu2O catalysts in Figure 3c showed two characteristic peaks of Cu 2p3/2 at binding energies of 933.4 eV and 935.2 eV mainly attributed to Cu+ and Cu2+, respectively [38]. According to the relative intensity of the Cu 2p peak in the spectrum, it could be inferred that the Cu+ species dominated the Cu oxide shell. This result further proved that the Cu2O shell was formed, consistent with the above characterization. Cu2+ species were also observed probably due to the easy oxidation of Cu2O in the air. The proportion of Cu+ species in different catalysts was calculated; the result is shown in Figure S4. Although the particle size of 4ZnO@Cu2O was much smaller than the Cu2O particle, the core-shell sample with the thinnest shell exhibited almost the same level of Cu+ species. This was probably attributed to the Cu–Zn interaction in 4ZnO@Cu2O, which resulted in part of the copper oxide being reduced to cuprous oxide. The ZnO@4Cu2O catalysts showed a much higher level of Cu+ species, indicating the presence of the stronger Cu–Zn interaction. In order to prove this result, the chemical state of ZnO in xZnO@yCu2O was also characterized by XPS. As shown in Figure 3d, all xZnO@yCu2O samples presented the same characteristic peak of Zn 2p3/2 with ZnO-400. This indicated that the valence state of the Zn species was kept during the growth of the Cu2O shell in the second step. However, by carefully comparing these samples, it could be observed that the binding energy of Zn 2p3/2 peak of 4ZnO@Cu2O and ZnO@4Cu2O shifted toward a higher one by 0.32 eV than that of ZnO (1022.82 eV vs.1022.50 eV). However, for ZnO@20Cu2O (1022.58 eV), it only shifted about 0.08 eV. These results indicated that the redistribution of electrons between the metals occurred due to the interaction between Cu and Zn [39], and the thin Cu2O shell in the xZnO@yCu2O catalysts had a Cu–Zn interaction rather than the thick ones. These were in accordance with the Cu 2p XPS results.

2.2. ECR Performance of ZnO Catalysts on the Catalyst

2.2.1. ZnO Catalysts

The electrochemical CO2 reduction performance of ZnO particles as the core was first measured by chronic amperometry in a flow cell with KOH as the electrolyte. As shown in Figure 4, all ZnO particles showed high CO selectivity. However, the H2 selectivity showed the opposite trend of CO, and almost no methane was produced. The ZnO-400 sample presented the best CO selectivity at −0.8~−1.0 V (vs. RHE), and the highest CO Faraday efficiency of 89.8% was achieved due to its large specific surface area and the highest proportion of ZnO (101) and (002) facets. According to the literature [40], the (101) and (002) facets of ZnO have a strong COOH* binding capacity and weak CO* adsorption capacity, therefore, they play a decisive role in the electroreduction of CO2 to produce CO. As for the ZnO-600 and ZnO-800 samples, the CO selectivity was reduced significantly by 20% because of the increase of the particle size and the decrease of (101) and (002) facets. Among these ZnO samples, the ZnO-400 sample also exhibited the highest current density (Figure 4d). The increase of current density at a high overpotential was probably due to the reinforcement of the hydrogen evolution reaction [41].

2.2.2. ZnO-400, Cu2O, and ZnO@4Cu2O Catalysts

As a comparison, the ECR performance of Cu2O and ZnO@4Cu2O catalysts was carried out in Figure 5. The main product of the ZnO catalyst was CO, which was proved above. At the low overpotential, the CO and H2 dominated the products over the Cu2O catalysts. As the overpotential increased, the selectivity of HCOOH, C2H4, and C2H5OH increased significantly with the inhibition of hydrogen evolution reaction, and HCOOH became the main product. C2H4 and C2H5OH were also first monitored at −0.6 V vs. RHE and then gradually increased with applied potential. The best performance of 15.2% C2H4 and 10.5% C2H5OH on Cu2O was obtained, which was similar to the report [42,43]. Compared to the Cu2O and ZnO catalysts, the selectivity of C2 products at a high overpotential was significantly boosted over the ZnO@4Cu2O catalyst (49.8% at −1.0 V vs. RHE), with the inhibition of HCOOH. These results indicated that the introduction of ZnO with high activity of CO2 to CO enhanced the CO–CO dimerization to generate C2H4, which was probably attributed to the high stability of OCCOH* intermediate on Cu+ derived from the Zn–Cu interaction, and the high selectivity of the coordinating transfer electron to facilitate the C2 products.

2.2.3. xZnO@yCu2O Catalysts

The performance of the electrochemical CO2 reduction over the xZnO@yCu2O catalysts with different Cu2O shell thickness and ZnO–400 as the core was carried out. As shown in Figure 6, besides the H2, CO, CH4, and HCOOH, the C2 products including ethylene and ethanol were obtained. The selectivity of CH4 over all catalysts was maintained at a very low level (almost near zero). As for the 4ZnO@Cu2O catalyst, the main products were still CO and H2. The selectivity of CO increased with the increase of the applied overpotential and reached the maximum (ca.60%) at −1.0 V vs. RHE. This result was similar to that of the ZnO core because the Cu2O shell was too thin. Almost the same selectivity of HCOOH (ca.10%) in different overpotentials was achieved. Moreover, with an overpotential beyond −0.6 V vs. RHE, C2H5OH was detected and increased up to 10.2% at −0.8 V vs. RHE and then gradually decreased to zero at −1.2 V vs. RHE. Simultaneously, the C2H4 product appeared at −0.8 V vs. RHE and reached about 5.1~6.4% at high potential (−1.0 V~−1.2 V vs. RHE). When the thickness of the Cu2O shell increased, the selectivity of HCOOH, C2H4, and C2H5OH products increased significantly. As for the ZnO@4Cu2O and ZnO@20Cu2O catalysts, HCOOH and CO became the main products at a low overpotential. C2H4 and C2H5OH products were initially detected at a low potential of −0.4 V vs. RHE with the selectivity of C2H5OH being a little higher than that of C2H4 over the ZnO@4Cu2O catalyst, which quickly increased as the applied potential increased. With an applied potential over the −0.8 V vs. RHE, the Faradic efficiency of C2H4 increased quickly and was significantly higher than that of C2H5OH. Thus, C2H4 became the main product over the ZnO@4Cu2O and the best faradaic efficiency of 33.5% was obtained at −1.0 V vs. RHE. As for the ZnO@20Cu2O catalyst, the C2H4 and C2H5OH products were first detected at −0.6 V vs. RHE and then gradually increased. When the applied potential was below −0.8 V vs. RHE, the selectivity of the C2H5OH product was a little higher than that of C2H4. At a high potential (−1.0~−1.2 V vs. RHE), the products were also dominated by CO and C2H4, and the faradaic efficiency of C2H4 was also significantly higher than that of C2H5OH. ZnO@20Cu2O exhibited almost the same product distribution with the Cu2O sample, which was due to the thick Cu2O shell, and the ZnO sites lose their function. Obviously, the xZnO@yCu2O catalysts showed different selectivity on C2H4 and C2H5OH at −0.8 V vs. RHE. As the shell thickness increased, the total current also increased, especially at a high overpotential. The highest total current density of 140.1 mA cm−2 with the partial current of C2 products at 70 mA cm−2 on ZnO@4Cu2O was achieved at −1.0 V vs. RHE (Figure 6d). The ZnO@20Cu2O catalyst with a thick Cu2O shell presented a much higher total current density than the Cu2O catalyst at a high overpotential. These results indicated that the distribution of the product depended on the shell thickness, and the suitable shell thickness could significantly enhance the C–C coupling of CO. These could be probably attributed to the interaction and synergetic effect of the ZnO core and Cu2O shell, where ZnO@4Cu2O exhibited the strongest interaction between ZnO and Cu2O.
The selectivity of C2 products (C2H4 and C2H5OH) over different xZnO@yCu2O catalysts is further summarized in Figure 7a. The selectivity of the C2 product showed a clear volcanic curve on all catalysts. Only low selectivity of C2 products was achieved over the 4ZnO@Cu2O catalyst, probably because the Cu2O shell was too thin to provide enough residence time for C–C coupling of CO. The ZnO@4Cu2O catalyst exhibited the highest Faradic efficiency of C2 products (49.8%, FE ethylene = 33.5% and FE ethanol = 16.3%) under −1.0 V vs. RHE, which was almost twice that of the Cu2O catalyst. The ZnO@20Cu2O catalyst showed a similar selectivity of C2 products with the Cu2O catalyst at a low overpotential. However, at a high overpotential, ZnO@20Cu2O had a higher selectivity of C2 products than the Cu2O catalyst (30% for ZnO@20Cu2O and 25.7% for Cu2O at −1.0 V vs. RHE). The C2 products selectivity on ZnO@20Cu2O was lower than ZnO@4Cu2O at high potential (30% vs. 49.8%). ZnO@20Cu2O had too thick a Cu2O shell and the function of ZnO core might be reduced, accounting for less intermediate CO* and inhibition of the C–C coupling process. The selectivity of C2 products corresponded to the ratio of the Cu2O (200)/(111) facet analyzed earlier. The interface of the Cu2O (200) facet and (111) could strengthen the CO–CO coupling, due to the strong *CO binding capacity and weak C2H4 adsorption capacity. The FE values of the C2 products and the total current over ZnO@4Cu2O were among the best reported in flow-cell and surpassed most reports in the H-cell (Figure 7b, Table S4) [18,37,44,45,46,47].

2.3. Electrochemical Characterization on xZnO@yCu2O Catalysts

To gain insights into the role of the ZnO core on the ECR performance of xZnO@yCu2O and enhancement of the C2 products (C2H4 and C2H5OH), the electrochemical active surface area (ECSA), electrochemical impedance, and electrokinetic were further determined. As shown in Figure 8a, the ZnO@4Cu2O catalyst showed the highest ECSA value (0.47 mF cm−2), indicating that more active sites were generated. The electrochemical impedance spectroscopy in Figure 8b showed that the ZnO@4Cu2O catalyst had the smallest charge transfer resistance of 2.84 Ω. This result suggested that the introduction of the ZnO core facilitated the electron transfer between the electrolyte and the electrocatalysts. Moreover, the LSV curves in Figure 8c illustrate that the ZnO@4Cu2O catalyst had the lowest initial overpotential for ECR. Additionally, the lower overpotential and higher current density were obtained in the KHCO3 electrolyte saturated with CO2 than in the electrolyte saturated with Ar. This indicated that the core-shell ZnO@4Cu2O catalyst could inhibit the hydrogen evolution reaction better than Cu2O, facilitating the improvement of C2 products selectivity and current density. As shown in Figure 8d, the ratio of C2 products/HCOOH on the ZnO@4Cu2O catalyst was almost two times higher than that on Cu2O (2.07 vs. 0.73). This indicated that the introduction of the ZnO core could significantly promote the selectivity of the coordinating transfer electron to generate the C2 products [48]. As previously reported, the Tafel slopes could provide some insights into the reaction mechanism. Tafel analyses of ethylene and ethanol production from electrochemical CO2 reduction on the xZnO@yCu2O catalysts were also carried out. As shown in Figure 8e, the ZnO@4Cu2O catalysts gave the smallest slopes of 46.9 mV dec−1 for C2H4 production, indicating that it had a faster electrode kinetic than other samples. Similar slopes for C2H4 production were also observed on the ZnO@20Cu2O, Cu2O, and 4ZnO@Cu2O catalysts, suggesting that the thickness of the Cu2O shell changed the overall rate-determining step. The same trend of Tafel slopes for ethanol production on these catalysts were also observed. The ZnO@4Cu2O catalyst still gave the smallest slope of 46.5 mV dec−1 for C2H6O. Additionally, the Tafel slopes of 46.9 and 46.5 mV dec−1 also suggested that the electron transfer to CO2 was not the rate-determining step for CO2 reduction [49]. The higher C2H4 and C2H5O production rate on the ZnO@4Cu2O catalyst indicated that C–C coupling was fast once the electrocatalytic CO2 reduction initiated.
Based on the above discussion, the high ECR activity of ZnO@4Cu2O could be attributed to several factors. As shown in Figure 3b, the ZnO@4Cu2O catalyst exposed a more enhanced joint interface between the (200) and (111) facets, due to the highest ratio of (200)/(111) facets, which could facilitate the C–C coupling to produce C2+ products [50]. The introduction of the ZnO core in the core-shell ZnO@4Cu2O catalyst enlarged the electrochemically active surface, lowered the electron transfer resistance between the electrolyte and the electrocatalysts, and enabled the ZnO@4Cu2O catalyst to have a higher overpotential of HER to inhibit H2O reduction and improve the efficiency of CO2 electroreduction to C2 products. Additionally, it also significantly promoted the selectivity of the coordinating transfer electron to generate C2+ products, probably due to the high CO production activity of the ZnO core. Moreover, the core-shell ZnO@4Cu2O catalysts had the faster electrode kinetic for C2 production. Therefore, the high electrochemical activity of ZnO@4Cu2O for ECR to C2 products most likely originated from the synergetic effects of the ZnO core and the Cu2O shell, which could not only generate a large electrochemical surface area, improve the HER overpotential, and fasten electron transfer, but could also promote the multielectron involved kinetics for ethylene and ethanol production.

3. Materials and Methods

3.1. Materials

Zinc acetate (C4H6O4Zn, 99.99%, Mackin), deuterium oxide (D2O, 99.9%, Macklin), potassium hydrogen carbonate (KHCO3, 99.7–100% (dried basis), Alfa), and dimethyl sulfoxide (DMSO, AR > 99.9%, Alfa) were used as received. Methoxy polyethylene glycol (CH3O(C2H4O)n H, Mn = 5000), ammonium citrate (C6H5O7(NH4)3, AR 98.5%), ammonia solution (NH3·H2O, AR, 25%~28%), copper nitrate trihydrate (Cu(CH3COO)2, 99.5%,), potassium hydroxide (KOH, 99.99%,), isopropyl alcohol (C3H8O, AR ≥ 99.5, Aladdin), and potassium chloride (KCl, AR, 99.5%) were purchased from Aladdin. Ethanol (C2H6O, 99.9%, Tianjin Yuanli Chemical Co., Ltd.), hydrazine hydrate (N2H4·H2O, 35% of volume fraction), Nafion 117 solution (5%, Sigma-Aldrich), and carbon-based gas diffusion layer (GDL, China, Wuhan, Gaoshiruilian Technology Co., Ltd.) were purchased from the specified suppliers. Ultrapure water (18.2 MΩ cm−1) was homemade by a Millipore system (Milli-Q Advantage A10).

3.2. Preparation of Catalysts

3.2.1. Synthesis of ZnO Catalysts

The ZnO core was prepared by a sol–gel method according to the reference [51]. A total of 0.015 mol of zinc acetate powder, 0.0076 mol of ammonium citrate, and 15 mL of ultrapure water were dissolved into a 100 mL beaker. These mixtures were mixed at room temperature. Then, 2 mL of ammonia solution and 25 mL of ethanol was added to the mixture and stirred at 75 °C for 120 min, and then the mixture was cooled to room temperature and aged for an additional 120 min. Finally, ammonia aqueous was dropped into the obtained solution under stirring, until the precipitate disappeared completely to form a wet gel. Then, the gel was dried at 90 °C for 720 min and calcined at different temperatures for 120 min in a muffle furnace. The size of ZnO was controlled by calcination at different temperatures (400 °C, 600 °C, and 800 °C), the heating rate was 5 °C/min. The final products were denoted as ZnO-400, ZnO-600, and ZnO-800, respectively.

3.2.2. Preparation of the ZnO@Cu2O Catalyst

The cuprous oxide shell was prepared using the epitaxial shell growth approach [52]. A total of 120 mL (2 wt.%) of polyethylene glycol methyl ether, 0.0002 mol as-synthesized ZnO core, and 2 mL 0.1 M Cu(NO3)2 were added into a 250 mL beaker, under stirring, for 30 min, and then 0.1 mL hydrazine hydrate (35 wt.%) was added. After stirring for an additional 5 min, the catalyst was separated by centrifugation at 8000 rpm for 20 min. The obtained ZnO@Cu2O particles were washed twice with water and then washed twice with ethanol. The products were separated at 12,000 rpm for 15 min for each wash. Finally, the ZnO@Cu2O particles were dried at 50 °C in a vacuum drying oven for 1200 min. The different shell thickness of the catalyst was tailored by controlling the different amounts of Cu(NO3)2 (0.5 mL, 8 mL, 40 mL), the obtained samples were named xZnO@yCu2O (4ZnO@Cu2O, ZnO@4Cu2O, and ZnO@20Cu2O), based on the ICP results.

3.2.3. Synthesis of Cu2O Catalyst

A reference Cu2O sample was also prepared in the same method with the Cu2O shell in ZnO@Cu2O. A total of 120 mL of polyethylene glycol methyl ether (2 wt.%) and 2 mL of 0.1 M Cu(NO3)2 were mixed under stirring at room temperature. A total of 0.1 mL of hydrazine hydrate (35 wt.%) was then added under stirring for an additional 5 min. Then, after the same post-treatment with ZnO@Cu2O, the Cu2O samples were obtained.

3.2.4. Preparation of Gas Diffusion Electrode

The gas diffusion electrode (GDE) was prepared according to the reference [53]. A total of 3 mL of isopropanol, 0.066 mL of Nafion solution, and 50 mg of the above as-synthesized powder was added to the beaker and well mixed. The mixture was handled by ultrasonication for 60 min to obtain the catalyst ink. Then, the ink was dropped on a gas diffusion layer (GDL, 3.5 cm × 1.5 cm) with a catalyst loading of 1 mg cm−2. The prepared gas diffusion electrode (GDE) was dried overnight for use.

3.3. Characterizations

The morphology and structure of the catalysts were characterized by scanning electron microscope (SEM, S-4800), transmission electron microscope (TEM, JEM-F200), and high-resolution TEM (HRTEM). The element distribution was analyzed by energy dispersive spectroscopy (EDS). The X-ray diffraction (XRD) pattern was collected on a D/MAX-2500 equipped with a Cu Kα X-ray source (λ = 0.1541 nm) at 40 kV and 40 mA, the scanning range was 20–90°, and the scanning speed was 5°/min. The atomic ratio of Zn and Cu was measured by inductively coupled plasma optical emission spectrometry (ICP-OES, VISTA-MPX). The X-ray photoelectron spectroscopy (XPS) was determined by ESCALAB 250Xi, using Al Kα X-ray source (h v = 1486.6 eV) and the binding energy was calibrated with C 1s spectra (binding energy = 284.8 eV).

3.4. Electrochemical CO2 Reduction

The electrochemical CO2 reduction was carried out in a flow electrolyzer. The cathode compartment and anode compartment of the flow cell were separated by ion-exchange membranes. The prepared GDE was used as the working electrode. GDL (3.5 cm × 1.5 cm) loaded with IrO2 and Ag/AgCl electrode were used as the counter electrode and reference electrode, respectively. All experiments were performed on the AutoLab (Model PGSTAT302 N). The electrolyte, 1 M KOH, was pumped through the cell at the flow rate of 1.5 mL min−1. CO2 was introduced at a flow rate of 20 mL min−1 to exhaust the air and saturated the electrolyte solution. The Faraday efficiency of the product on all samples was calculated using Equations (1–3) in the supplementary materials. Equipment should be checked for leaks of gas and liquid before starting the experiment.
During the electrochemical CO2 reduction process, the gas product was quantitatively analyzed by online gas chromatography (GC, GC-2010) equipped with a BID detector. After the reaction, the cathode liquid products were quantitatively analyzed with a liquid nuclear magnetic resonance spectrometer (400 MHz). The electrochemical analysis was carried out in an H-Cell. The cathode chamber and anode chamber were separated by a proton exchange membrane. Platinum mesh and the Ag/AgCl electrode were used as the counter electrode and the reference electrode. The 0.1 M KHCO3 solution was used as the electrolyte. Cyclic voltammetry in the range of 0.45~0.84 V vs. RHE was carried out to measure the double-layer capacitance of the catalyst for the electrochemical CO2 reduction at the scanning rate of 50 mV s−1, 100 mV s−1, 150 mV s−1, 200 mV s−1, 250 mV s−1, and 300 mV s−1. The redox peaks of the catalysts were explored by cyclic voltammetry at −1.31 V~0.69 V vs. RHE, with a scan rate of 100 mV s−1. Linear sweep voltammetry was carried out at −1.31 V~0.69 V vs. RHE with a sweep rate of 10 mV s−1 in the electrolyte saturated with CO2 and Ar, respectively. The electrochemical impedance spectroscopy of the catalyst was performed in the 0.1 M KHCO3 electrolyte saturated with CO2 at −1.0 V vs. RHE from 0.01 Hz to 105 Hz.

4. Conclusions

In summary, we synthesized a new type of catalyst with zinc core-copper shell by the sol–gel method and epitaxial shell growth method. Due to the double active sites on the ZnO@4Cu2O catalyst, it was possible to get the CO2 to first generate CO at the ZnO site, and then the CO coupled on Cu2O to generate the C2 product. The synergistic effect of Cu and Zn enhanced the selectivity of C2 products on the catalyst. The core-shell ZnO@4Cu2O enclosed with the enhanced joint interface of Cu2O (111) and Cu2O (200) facets and high ECSA, exhibited the best C2 product selectivity of 49.8% (ethylene 33.5%, ethanol 16.3%) and a high total current density of 140.1 mA cm−2. This core-shell structure could strengthen the synergy between Cu and Zn oxides to simultaneously improve the selectivity of C2 products and the total current. The product distribution depended on the shell thickness, and the suitable shell thickness could significantly enhance the C–C coupling of CO. The thickness of the Cu2O shell in xZnO@yCu2O could change the overall rate-determining step. Our work might provide some new insights for developing novel high-efficiency catalysts and enhancing the selectivity of electrochemical CO2 reduction to produce high-value C2 products.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/catal11050535/s1. Figure S1: TEM image of (a) 4ZnO@Cu2O and (b) ZnO@20Cu2O. The line EDX analysis along the arrow (c) 4ZnO@Cu2O and (d) ZnO@20Cu2O. Figure S2: XRD pattern of Cu2O and its PDF card. Figure S3: Cu LMM Auger spectra of xZnO@yCu2O catalysts with different shell thickness. Figure S4: The ratio of Cu1+/Cu total with different shell thickness catalysts. Figure S5: The current density of C2 products on xZnO@Cu2O catalysts with different shell thickness. Figure S6: Cyclic voltammograms of different catalysts in 0.1 M KHCO3 (CO2-saturated) at scan sates of 50 mV s−1, 100 mV s−1, 150 mV s−1, 200 mV s−1, 250 mV s−1, and 300 mV s−1. (a) Cu2O; (b) ZnO@20Cu2O; (c) ZnO@4Cu2O; (d) 4ZnO@Cu2O; (e) ZnO; and (f) double-layer capacitance measured by cyclic voltammetry in 0.1 M KHCO3 saturated by CO2 on the catalysts with different shell thickness. Figure S7: The particle size distribution of (a) ZnO-400, (b) ZnO-600, and (c) ZnO-800. Figure S8: The particle size distribution of (a) Cu2O, (b) 4ZnO@Cu2O, (c) ZnO@4Cu2O, and (d) ZnO@20Cu2O. Figure S9: The flow cell used for electrochemical reduction of CO2. Table S1: Proportion of elements in the sample measured by ICP-OES. Table S2: The intensity and proportion of different facets of ZnO in the xZnO@yCu2O catalyst. Table S3: The intensity and proportion of significant facets of Cu2O in xZnO@yCu2O catalyst. Table S4: Comparison of multicarbon product formation from various Cu-based catalysts. Table S5: The distribution of products at different applied potentials on ZnO@4Cu2O catalyst. Table S6: The distribution of products at different applied potentials on Cu2O catalyst. Table S7: The resistance of different catalysts.

Author Contributions

Investigation, Methodology, Formal analysis, Data curation, Writing—original draft, S.Z.; Validation, Investigation, Formal analysis, X.R.; Validation, Formal analysis, X.L.; Formal analysis, X.N.; Visualization, Investigation, M.W.; Resources, Formal analysis, S.X.; Resources, Z.W.; Formal analysis, Y.H.; Supervision, Conceptualization, Data curation, Methodology, Writing—review and editing, Q.W. All authors have read and agreed to the published version of the manuscript.

Funding

We got funding support from the National Natural Science Foundation of China (Grant Numbers 21476169).

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Goeppert, A.; Czaun, M.; Jones, J.P.; Prakash, G.K.S.; Olah, G.A. Recycling of carbon dioxide to methanol and derived products-closing the loop. Chem. Soc. Rev. 2014, 43, 7995–8048. [Google Scholar]
  2. Zhang, H.C.; Chang, X.X.; Chen, J.G.G.; Goddard, W.A.; Xu, B.J.; Cheng, M.J.; Lu, Q. Computational and experimental demonstrations of one-pot tandem catalysis for electrochemical carbon dioxide reduction to methane. Nat. Commun. 2019, 10, 3340. [Google Scholar] [CrossRef] [Green Version]
  3. Zhu, D.D.; Liu, J.L.; Qiao, S.Z. Recent advances in inorganic heterogeneous electrocatalysts for reduction of carbon dioxide. Adv. Mater. 2016, 28, 3423–3452. [Google Scholar] [CrossRef]
  4. Liu, Y.M.; Fan, X.F.; Nayak, A.; Wang, Y.; Shan, B.; Quan, X.; Meyer, T.J. Steering CO2 electroreduction toward ethanol production by a surface-bound Ru polypyridyl carbene catalyst on N-doped porous carbon. Proc. Natl. Acad. Sci. USA 2019, 116, 26353–26358. [Google Scholar] [CrossRef]
  5. Li, D.; Liu, T.T.; Huang, L.L.; Wu, J.J.; Li, N.; Zhen, L.; Feng, Y.J. Selective CO2–to–formate electrochemical conversion with core–shell structured Cu2O/Cu@C composites immobilized on nitrogen-doped graphene sheets. J. Mater. Chem. A 2020, 8, 18302–18309. [Google Scholar] [CrossRef]
  6. Li, J.C.; Kuang, Y.; Meng, Y.T.; Tian, X.; Hung, W.H.; Zhang, X.; Li, A.W.; Xu, M.Q.; Zhou, W.; Ku, C.S.; et al. Electroreduction of CO2 to formate on a copper-based electrocatalyst at high pressures with high energy conversion efficiency. J. Am. Chem. Soc. 2020, 142, 7276–7282. [Google Scholar] [CrossRef]
  7. Tan, Z.H.; Peng, T.Y.; Tian, X.J.; Wang, W.H.; Wang, X.S.; Yang, Z.X.; Ning, H.; Zhao, Q.S.; Wu, M.B. Controllable Synthesis of Leaf-Like CuO Nanosheets for Selective CO2 Electroreduction to Ethylene. ChemElectroChem 2020, 7, 2020–2025. [Google Scholar] [CrossRef]
  8. Raciti, D.; Livi, K.J.; Wang, C. Highly dense Cu nanowires for low-overpotential CO2 reduction. Nano Lett. 2015, 15, 6829–6835. [Google Scholar] [CrossRef] [PubMed]
  9. Chen, C.J.; Sun, X.F.; Lu, L.; Yang, D.X.; Ma, J.; Zhu, Q.G.; Qian, Q.L.; Han, B.X. Efficient electroreduction of CO2 to C2 products over B-doped oxide-derived copper. Green Chem. 2018, 20, 4579–4583. [Google Scholar] [CrossRef]
  10. Wang, M.; Ren, X.N.; Yuan, G.; Niu, X.P.; Xu, Q.L.; Gao, W.L.; Zhu, S.K.; Wang, Q.F. Selective electroreduction of CO2 to CO over co-electrodeposited dendritic core-shell indium-doped Cu@Cu2O catalyst. J. CO2 Util. 2020, 37, 204–212. [Google Scholar] [CrossRef]
  11. Bai, X.W.; Li, Q.; Shi, L.; Niu, X.H.; Ling, C.Y.; Wang, J.L. Hybrid Cu0 and Cu x+ as atomic interfaces promote high-selectivity conversion of CO2 to C2H5OH at low potential. Small 2020, 16, 1901981. [Google Scholar] [CrossRef] [PubMed]
  12. Han, H.; Noh, Y.; Kim, Y.; Park, S.; Yoon, W.; Jang, D.; Choi, S.M.; Kim, W.B. Selective electrochemical CO2 conversion to multicarbon alcohols on highly efficient N-doped porous carbon-suported Cu catalysts. Green Chem. 2020, 22, 71–84. [Google Scholar] [CrossRef]
  13. Zhuang, T.T.; Liang, Z.Q.; Seifitokaldani, A.; Li, Y.; De Luna, P.; Burdyny, T.; Che, F.L.; Meng, F.; Min, Y.M.; Quintero-Bermudez, R. Steering post-C-C coupling selectivity enables high efficiency electroreduction of carbon dioxide to multi-carbon alcohols. Nat. Catal. 2018, 1, 421–428. [Google Scholar] [CrossRef]
  14. Kim, D.; Lee, S.; Ocon, J.D.; Jeong, B.; Lee, J.K.; Lee, J. Insights into an autonomously formed oxygen-evacuated Cu2O electrode for the selective production of C2H4 from CO2. Phys. Chem. Chem. Phys. 2015, 17, 824–830. [Google Scholar] [CrossRef] [PubMed]
  15. Lee, S.Y.; Jung, H.; Kim, N.K.; Oh, H.S.; Min, B.K.; Hwang, Y.J. Mixed copper states in anodized Cu electrocatalyst for stable and selective ethylene production from CO2 reduction. J. Am. Chem. Soc. 2018, 140, 8681–8689. [Google Scholar] [CrossRef] [PubMed]
  16. Wang, H.; Matios, E.; Wang, C.L.; Luo, J.M.; Lu, X.; Hu, X.F.; Li, W.Y. Rapid and scalable synthesis of cuprous halide-derived copper nano-architectures for selective electrochemical reduction of carbon dioxide. Nano Lett. 2019, 19, 3925–3932. [Google Scholar] [CrossRef] [PubMed]
  17. Loiudice, A.; Lobaccaro, P.; Kamali, E.A.; Thao, T.; Huang, B.H.; Ager, J.W.; Buonsanti, R. Tailoring copper nanocrystals towards C-2 Products in electrochemical CO2 Reduction. Angew. Chem. Int. Ed. 2016, 55, 5789–5792. [Google Scholar] [CrossRef] [Green Version]
  18. Ren, D.; Deng, Y.L.; 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]
  19. Li, C.W.; Kanan, M.W. CO2 Reduction at low overpotential on Cu electrodes resulting from the reduction of thick Cu2O films. J. Am. Chem. Soc. 2012, 134, 7231–7234. [Google Scholar] [CrossRef]
  20. Wang, W.H.; Ning, H.; Yang, Z.X.; Feng, Z.X.; Wang, J.L.; Wang, X.S.; Mao, Q.H.; Wu, W.T.; Zhao, Q.S.; Hu, H.; et al. Interface-induced controllable synthesis of Cu2O nanocubes for electroreduction CO2 to C2H4. Electrochim. Acta. 2019, 306, 360–365. [Google Scholar] [CrossRef]
  21. Su, X.S.; Sun, Y.M.; Jin, L.; Zhang, L.; Yang, Y.; Kerns, P.; Liu, B.; Li, S.Z.; He, J. Hierarchically porous Cu/Zn bimetallic catalysts for highly selective CO2 electroreduction to liquid C2 products. Appl. Catal. B Environ. 2020, 269, 118800. [Google Scholar] [CrossRef]
  22. Jimenez, C.; Garcia, J.; Martinez, F.; Camarillo, R.; Rincon, J. Cu nanoparticles deposited on CNT by supercritical fluid deposition for electrochemical reduction of CO2 in a gas phase GDE cell. Electrochim. Acta 2020, 337, 135663. [Google Scholar] [CrossRef]
  23. Ning, H.; Mao, Q.H.; Wang, W.H.; Yang, Z.X.; Wang, X.S.; Zhao, Q.S.; Song, Y.; Wu, M.B. N-doped reduced graphene oxide supported Cu2O nanocubes as high active catalyst for CO2 electroreduction to C2H4. J. Alloys Compd. 2019, 785, 7–12. [Google Scholar] [CrossRef]
  24. Feng, Y.H.; An, W.; Wang, Z.M.; Wang, Y.Q.; Men, Y.; Du, Y.Y. Electrochemical CO2 reduction reaction on M@Cu (211) bimetallic single-atom surface alloys: Mechanism, kinetics, and catalyst screening. ACS Sustain. Chem. Eng. 2020, 8, 210–222. [Google Scholar] [CrossRef]
  25. Gao, J.; Ren, D.; Guo, X.Y.; Zakeeruddin, S.M.; Gratzel, M. Sequential catalysis enables enhanced C–C coupling towards multi-carbon alkenes and alcohols in carbon dioxide reduction: A study on bifunctional Cu/Au electrocatalysts. Faraday Discuss 2019, 215, 282–296. [Google Scholar] [CrossRef] [PubMed]
  26. Gao, J.; Ren, D.; Guo, X.Y.; Zakeeruddin, S.M.; Gratzel, M. Activation of bimetallic AgCu foam electrocatalysts for ethanol formation from CO2 by selective Cu oxidation/reduction. Faraday Discuss 2019, 215, 282–296. [Google Scholar] [CrossRef]
  27. Ren, D.; Gao, J.; Pan, L.F.; Wang, Z.W.; Luo, J.S.; Zakeeruddin, S.M.; Hagfeldt, A.; Gratzel, M. Atomic layer deposition of ZnO on CuO enables selective and efficient electroreduction of carbon dioxide to liquid fuels. Angew. Chem. Int. Ed. 2019, 58, 15036–15040. [Google Scholar] [CrossRef]
  28. Merino-Garcia, I.; Albo, J.; Solla-Gullon, J.; Montiel, V.; Lrabien, A. Cu oxide/ZnO-based surfaces for a selective ethylene production from gas-phase CO2 electroconversion. J. CO2 Util. 2019, 31, 135–142. [Google Scholar] [CrossRef]
  29. Payra, S.; Shenoy, S.; Chakraborty, C.; Tarafder, K.; Roy, S. Structure-sensitive electrocatalytic reduction of CO2 to methanol over carbon-supported intermetallic PtZn nano-alloys. ACS Appl. Mater. Inter. 2020, 12, 19402–19414. [Google Scholar] [CrossRef]
  30. Han, J.; Li, S.S.; Chen, J.Y.; Liu, Y.Q.; Geng, D.S.; Wang, D.W.; Zhang, L.C. Dendritic Ag/Pd Alloy Nanostructure Arrays for Electrochemical CO2 Reduction. ChemElectroChem 2020, 7, 2608–2613. [Google Scholar] [CrossRef]
  31. Zhi, X.; Jiao, Y.; Zheng, Y.; Vasileff, A.; Qiao, S.Z. Selectivity roadmap for electrochemical CO2 reduction on copper-based alloy catalysts. Nano Today 2020, 71, 104601. [Google Scholar] [CrossRef]
  32. Tsuji, M.; Yamaguchi, D.; Mstsunaga, M.; Alam, M.J. Epitaxial growth of Au@Cu core-shell nanocrystals prepared using the PVP-assisted polyol reduction method. Cryst. Growth DES 2010, 10, 5129–5135. [Google Scholar] [CrossRef]
  33. Tsai, Y.H.; Chanda, K.; Chu, Y.T.; Chiu, C.Y.; Huang, M.H. Direct formation of small Cu2O nanocubes, octahedra, and octapods for efficient synthesis of triazoles. Nanoscale 2014, 6, 8704–8709. [Google Scholar] [CrossRef] [PubMed]
  34. Chen, C.J.; Yan, X.P.; Liu, S.J.; Wu, Y.H.; Wan, Q.; Sun, X.F.; Zhu, Q.G.; Liu, H.Z.; Ma, J.; Zheng, L.R.; et al. Highly efficient electroreduction of CO2 to C2+ alcohols on heterogeneous dual active sites. Angew. Chem. Int. Ed. 2020, 59, 16459–16464. [Google Scholar] [CrossRef] [PubMed]
  35. Schouten, K.J.P.; Qin, Z.S.; Gallent, E.P.; Koper, M.T.M. Two Pathways for the Formation of Ethylene in CO Reduction on Single-Crystal Copper Electrodes. J. Am. Chem. Soc. 2012, 134, 9864–9867. [Google Scholar] [CrossRef]
  36. Xu, H.P.; Rebollar, D.; He, H.Y.; Chong, L.N.; Liu, Y.Z.; Liu, C.; Sun, C.J.; Li, T.; Muntean, J.V.; Winans, R.E.; et al. Highly selective electrocatalytic CO(2) reduction to ethanol by metallic clusters dynamically formed from atomically dispersed copper. Nat. Energy 2020, 5, 623–632. [Google Scholar] [CrossRef]
  37. Song, H.; Song, J.T.; Kim, B.; Tan, Y.C.; Oh, J. Activation of C2H4 reaction pathways in electrochemical CO2 reduction under low CO2 partial pressure. Appl. Catal. B Environ. 2020, 272, 119049. [Google Scholar] [CrossRef]
  38. Shao, Q.; Wang, P.T.; Liu, S.H.; Huang, X.Q. Advanced engineering of core/shell nanostructures for electrochemical carbon dioxide reduction. J. Mater. Chem. A. 2019, 7, 20478–20493. [Google Scholar] [CrossRef]
  39. Liu, K.H.; Wang, J.Z.; Shi, M.M.; Yan, J.M.; Jiang, Q. Simultaneous Achieving of High Faradaic Efficiency and CO Partial Current Density for CO2 Reduction via Robust, Noble-Metal-Free Zn Nanosheets with Favorable Adsorption Energy. Adv. Energy Mater. 2019, 9, 1900276. [Google Scholar] [CrossRef]
  40. Birdja, Y.Y.; Perez-Gallent, E.; Figueiredo, M.C.; Gottle, A.J.; Calle-Vallejo, F.; Koper, M.T.M. Advances and challenges in understanding the electrocatalytic conversion of carbon dioxide to fuels. Nat. Energy. 2019, 4, 732–745. [Google Scholar] [CrossRef]
  41. Tan, X.Y.; Yu, C.; Zhao, C.T.; Huang, H.W.; Yao, X.C.; Han, X.T.; Guo, W.; Cui, S.; Huang, H.L.; Qiu, J.S. Restructuring of Cu2O to Cu2O@Cu-Metal-Organic Frameworks for Selective Electrochemical Reduction of CO2. ACS Appl. Mater. Inter. 2019, 11, 9904–9910. [Google Scholar] [CrossRef]
  42. Suen, N.T.; Kong, Z.R.; Hsu, C.S.; Chen, H.C.; Tung, C.W.; Lu, Y.R.; Dong, C.L.; Shen, C.C.; Chung, J.C.; Chen, H.M. Morphology Manipulation of Copper Nanocrystals and Product Selectivity in the Electrocatalytic Reduction of Carbon Dioxide. ACS Catal. 2019, 9, 5217–5222. [Google Scholar] [CrossRef]
  43. Zhang, J.; Luo, W.; Zuttel, A. Self-supported copper-based gas diffusion electrodes for CO2 electrochemical reduction. J. Mater. Chem. A 2019, 7, 26285–26292. [Google Scholar] [CrossRef]
  44. Reller, C.; Krause, R.; Volkova, E.; Schmid, B.; Neubauer, S.; Rucki, A.; Schuster, M.; Schmid, G. Selective Electroreduction of CO2 toward Ethylene on Nano Dendritic Copper Catalysts at High Current Density. Adv. Energy Mater. 2017, 7, 1602114. [Google Scholar] [CrossRef]
  45. Shen, S.B.; Peng, X.Y.; Song, L.D.; Qin, Y.; Li, C.; Zhuo, L.C.; He, J.; Ren, J.Q.; Liu, X.J.; Luo, J. AuCu Alloy Nanoparticle Embedded Cu Submicrocone Arrays for Selective Conversion of CO2 to Ethanol. Small 2019, 15, 1902229. [Google Scholar] [CrossRef] [PubMed]
  46. Song, Y.; Peng, R.; Hensley, D.K.; Bonnesen, P.V.; Liang, L.B.; Wu, Z.L.; Meyer, H.M.; Chi, M.F.; Ma, C.; Sumpter, B.G.; et al. High-Selectivity Electrochemical Conversion of CO2 to Ethanol using a Copper Nanoparticle/N-Doped Graphene Electrode. Chemistryselect 2016, 1, 6055–6061. [Google Scholar] [CrossRef] [Green Version]
  47. Balamurugan, M.; Jeong, H.Y.; Choutipalli, V.S.K.; Hong, J.S.; Sao, H.; Saravanan, N.; Jang, J.H.; Lee, K.G.; Lee, Y.H.; Im, S.W.; et al. Electrocatalytic Reduction of CO2 to Ethylene by Molecular Cu-Complex Immobilized on Graphitized Mesoporous Carbon. Small 2020, 16, 2000955. [Google Scholar] [CrossRef]
  48. Wang, S.; Kou, T.; Baker, S.E.; Duoss, E.B.; Li, Y. Recent progress in electrochemical reduction of CO2 by oxide-derived copper catalysts. Mater. Today Nano 2020, 12, 100096. [Google Scholar] [CrossRef]
  49. Liu, Y.M.; Chen, S.; Quan, X.; Yu, H.T. Efficient Electrochemical Reduction of Carbon Dioxide to Acetate on Nitrogen-Doped Nanodiamond. J. Am. Chem. Soc. 2015, 137, 11631–11636. [Google Scholar] [CrossRef]
  50. Gao, Y.G.; Wu, Q.; Liang, X.Z.; Wang, Z.Y.; Zheng, Z.K.; Wang, P.; Liu, Y.Y.; Dai, Y.; Whangbo, M.H.; Huang, B.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]
  51. Liu, Z.F.; Jin, Z.G.; Li, W.; Qiu, J.J. Preparation of ZnO porous thin films by sol-gel method using PEG template. Mater. Lett. 2005, 59, 3620–3625. [Google Scholar] [CrossRef]
  52. Zhang, L.; Jing, H.; Boisvert, G.; He, J.Z.; Wang, H. Geometry Control and Optical Tunability of Metal-Cuprous Oxide Core-Shell Nanoparticles. ACS Nano 2012, 6, 3514–3527. [Google Scholar] [CrossRef] [PubMed]
  53. Ma, S.C.; Sadakiyo, M.; Luo, R.; Heima, M.; Yamauchi, M.; Kenis, P.J.A. One-step electrosynthesis of ethylene and ethanol from CO2 in an alkaline electrolyzer. J. Power Sources 2016, 301, 219–228. [Google Scholar] [CrossRef] [Green Version]
Figure 1. SEM images of (a) ZnO-400, (b) ZnO-600, and (c)ZnO-800. (d) XRD patterns and (e) (101) and (002) facets proportion of different ZnO catalysts. (f) Zn XPS spectrum of ZnO-400.
Figure 1. SEM images of (a) ZnO-400, (b) ZnO-600, and (c)ZnO-800. (d) XRD patterns and (e) (101) and (002) facets proportion of different ZnO catalysts. (f) Zn XPS spectrum of ZnO-400.
Catalysts 11 00535 g001
Scheme 1. Illustration of formation of xZnO@yCu2O.
Scheme 1. Illustration of formation of xZnO@yCu2O.
Catalysts 11 00535 sch001
Figure 2. SEM images of (a) Cu2O, (b) 4ZnO@Cu2O, (c) ZnO@4Cu2O, and (d) ZnO@20Cu2O. (e) TEM image of ZnO@4Cu2O and (f) line EDX analysis along the arrow. (g) HR-TEM image of ZnO@4Cu2O. (h) Cu K and (i) Zn K mappings of ZnO@4Cu2O.
Figure 2. SEM images of (a) Cu2O, (b) 4ZnO@Cu2O, (c) ZnO@4Cu2O, and (d) ZnO@20Cu2O. (e) TEM image of ZnO@4Cu2O and (f) line EDX analysis along the arrow. (g) HR-TEM image of ZnO@4Cu2O. (h) Cu K and (i) Zn K mappings of ZnO@4Cu2O.
Catalysts 11 00535 g002
Figure 3. (a) XRD patterns of xZnO@yCu2O catalyst. (b) The ratio of Cu2O (200) facet/(111) facet on different catalysts. (c) Cu 2p XPS diagrams and (d) Zn 2p XPS diagrams of different shell thickness catalysts.
Figure 3. (a) XRD patterns of xZnO@yCu2O catalyst. (b) The ratio of Cu2O (200) facet/(111) facet on different catalysts. (c) Cu 2p XPS diagrams and (d) Zn 2p XPS diagrams of different shell thickness catalysts.
Catalysts 11 00535 g003
Figure 4. Faradaic efficiency on (a) ZnO-400, (b) Zn-600, and (c) ZnO-800. (d) The total current density of products on different ZnO catalysts.
Figure 4. Faradaic efficiency on (a) ZnO-400, (b) Zn-600, and (c) ZnO-800. (d) The total current density of products on different ZnO catalysts.
Catalysts 11 00535 g004
Figure 5. Faradaic efficiency of products on different catalysts (a) ZnO-400, (b) Cu2O, and (c) ZnO@4Cu2O. (d) The total current density of products on different catalysts.
Figure 5. Faradaic efficiency of products on different catalysts (a) ZnO-400, (b) Cu2O, and (c) ZnO@4Cu2O. (d) The total current density of products on different catalysts.
Catalysts 11 00535 g005
Figure 6. Faradaic efficiency of products on xZnO@uCu2O catalysts (a) 4ZnO@Cu2O, (b) ZnO@4Cu2O, and (c) ZnO@20Cu2O. (d) The total current density of products on xZnO@yCu2O catalysts.
Figure 6. Faradaic efficiency of products on xZnO@uCu2O catalysts (a) 4ZnO@Cu2O, (b) ZnO@4Cu2O, and (c) ZnO@20Cu2O. (d) The total current density of products on xZnO@yCu2O catalysts.
Catalysts 11 00535 g006
Figure 7. (a) Selectivity towards C2 products on xZnO@Cu2O catalysts with different shell thickness. (b) The comparison of total current density and selectivity of C2 products between this work with others.
Figure 7. (a) Selectivity towards C2 products on xZnO@Cu2O catalysts with different shell thickness. (b) The comparison of total current density and selectivity of C2 products between this work with others.
Catalysts 11 00535 g007
Figure 8. Electrochemical characterizations of different catalysts. (a) Double-layer capacitance measured by cyclic voltammetry in 0.1 M KHCO3 saturated by CO2. (b) Electrochemical impedance spectroscopy (EIS) plots were carried at −1.0 V vs. RHE in 0.1 M KHCO3 saturated with CO2. (c) Linear sweep voltammetry (LSV) curves measured in 0.1 M KHCO3 solution saturated by CO2 and Ar with a scan rate of 10 mV s−1. (d) The ratio of Faraday efficiency of C2H4 and C2H5OH to Faraday efficiency of HCOOH on catalysts with different shell thickness. Tafel Kinetics over different catalysts of (e) C2H4 and (f) C2H5OH.
Figure 8. Electrochemical characterizations of different catalysts. (a) Double-layer capacitance measured by cyclic voltammetry in 0.1 M KHCO3 saturated by CO2. (b) Electrochemical impedance spectroscopy (EIS) plots were carried at −1.0 V vs. RHE in 0.1 M KHCO3 saturated with CO2. (c) Linear sweep voltammetry (LSV) curves measured in 0.1 M KHCO3 solution saturated by CO2 and Ar with a scan rate of 10 mV s−1. (d) The ratio of Faraday efficiency of C2H4 and C2H5OH to Faraday efficiency of HCOOH on catalysts with different shell thickness. Tafel Kinetics over different catalysts of (e) C2H4 and (f) C2H5OH.
Catalysts 11 00535 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhu, S.; Ren, X.; Li, X.; Niu, X.; Wang, M.; Xu, S.; Wang, Z.; Han, Y.; Wang, Q. Core-Shell ZnO@Cu2O as Catalyst to Enhance the Electrochemical Reduction of Carbon Dioxide to C2 Products. Catalysts 2021, 11, 535. https://doi.org/10.3390/catal11050535

AMA Style

Zhu S, Ren X, Li X, Niu X, Wang M, Xu S, Wang Z, Han Y, Wang Q. Core-Shell ZnO@Cu2O as Catalyst to Enhance the Electrochemical Reduction of Carbon Dioxide to C2 Products. Catalysts. 2021; 11(5):535. https://doi.org/10.3390/catal11050535

Chicago/Turabian Style

Zhu, Shuaikang, Xiaona Ren, Xiaoxue Li, Xiaopo Niu, Miao Wang, Shuang Xu, Zheyuan Wang, Yunxi Han, and Qingfa Wang. 2021. "Core-Shell ZnO@Cu2O as Catalyst to Enhance the Electrochemical Reduction of Carbon Dioxide to C2 Products" Catalysts 11, no. 5: 535. https://doi.org/10.3390/catal11050535

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