Next Article in Journal
Biomedical and Textile Applications of Alternanthera sessilis Leaf Extract Mediated Synthesis of Colloidal Silver Nanoparticle
Next Article in Special Issue
Recent Progress of Metal-Organic Frameworks and Metal-Organic Frameworks-Based Heterostructures as Photocatalysts
Previous Article in Journal
Studies of the Structure and Optical Properties of BaSrMgWO6 Thin Films Deposited by a Spin-Coating Method
Previous Article in Special Issue
Photocatalytic and Electrocatalytic Properties of Cu-Loaded ZIF-67-Derivatized Bean Sprout-Like Co-TiO2/Ti Nanostructures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Interface Engineered V-Zn Hybrids: Electrocatalytic and Photocatalytic CO2 Reductions

1
Department of Chemistry, Chungnam National University, Daejeon 34134, Korea
2
Department of Chemical Engineering and Applied Chemistry, Chungnam National University, Daejeon 34134, Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2022, 12(16), 2758; https://doi.org/10.3390/nano12162758
Submission received: 21 July 2022 / Revised: 3 August 2022 / Accepted: 9 August 2022 / Published: 11 August 2022

Abstract

:
V-Zn hybrids have widely been used as catalyst materials in the environment and as energy. Herein, V-Zn hybrid electrodes were prepared by the hydrothermal and sputter-deposition methods using a Zn foil support. Their electrocatalytic CO2 reduction (EC CO2 RR) performances were tested under various applied potentials, different electrolytes, and concentrations before and after thermal treatment of the demonstrated electrode. Gas and liquid products were confirmed by gas chromatography and nuclear magnetic resonance spectroscopy, respectively. For V-Zn electrode by hydrothermal method produced mainly syngas (CO and H2) with tunable ratio by varying applied potential. Minor products include CH4, C2H4, and C2H6. A liquid product of formate showed a Faradaic efficiency (FE) of 2%. EC CO2 RR efficiency for CO, CH4, and formate was best in 0.2 M KHCO3 electrolyte condition. CO and formate were further increased by photoirradiation and Nafion-treated electrode. Formate and CH4 productions were significantly increased by thermal treatment of the V-Zn electrode. CO production was diminished for the V-Zn electrode by sputter deposition but was recovered by thermal treatment. Photocatalytic CO2 RR was tested to find that RR products include CH3OH, CO, CH4, C2H4, and C2H6. Interestingly long-chain hydrocarbons (CnH2n and CnH2n+2, where n = 3–6) were first observed under mild conditions. The long-chain formation was understood by Fisher-Tropsch (F-T) synthesis. Alkenes were observed to be more produced than alkanes unlike in the conventional F-T synthesis. The present new findings provide useful clues for the development of hybrid electro-and photo-catalysts tested under various experimental conditions in energy and environment.

1. Introduction

Hybrids of different metal elements and their oxides have extensively been employed in catalyst application areas for energy and environments [1,2,3,4,5]. Among them, V-Zn hybrids with diverse compositions and morphologies have been developed and applied to photocatalysts, electrocatalysts, and energy storage materials [6,7,8,9,10,11,12,13,14,15,16,17,18]. In photocatalytic catalyst application of a V-Zn hybrid, it was theoretically found that photogenerated carriers are efficiently separated at the interface of V2O5/ZnV2O6, where the O 2p state is highly contributed near the Fermi level in the valence band structure [13]. Li et al. synthesized V2O5/ZnV2O6 nanosheets by the solvothermal method and the post thermal treatment (300–500 °C) [14]; they tested photocatalytic CO2 reduction activities for V2O5/ZnV2O6 nanosheets, V2O5, and ZnV2O6, and found that the V2O5/ZnV2O6 showed the highest CO production activity; 2.2× and 1.9× higher than those observed in V2O5 and ZnV2O6, respectively. The enhanced activity was attributed to charge transfer from ZnV2O6 to V2O5 at the interface and enhanced adsorption of CO2 on the surface. Similarly, in the photocatalytic CO2 RR over ZnV2O6/g-C3N4, CH3OH, CO, and CH4 were reported to be more produced than in ZnV2O6 and g-C3N4, attributed to the synergistic interfacial hybrid effect [15]. In the case of photocatalytic CO2 RR over ZnO, V2O5, ZnO/V2O5 composite, and ZnV2O6 nanosheets [16], Bafaqueer et al. reported that ZnV2O6 nanosheets showed the highest performance on CH3OH production by CO2 RR. The activity of ZnV2O6 was 3.4× higher than that of the ZnO/V2O5 composite. In this experiment, formic acid and acetic acid were newly reported; they also found that the CO2 RR activity was poorly degraded upon calcination at 550 °C; this indicates that many factors are involved in the mechanism, including surface and interface engineering. Tahir also reported a similar observation that hierarchical 3D ZnV2O4 microspheres showed higher CO2 RR performance than ZnO/V2O5 composite [17]. In this experiment, CO2 RR products include CO, CH3OH, CH4, and C2H6. As discussed above, the interface plays a significant role in CO2 RR efficiency and it needs to be further elucidated in other developed catalyst systems.
Although there are several works of literature on the photocatalytic CO2 RR [11,12,13,14,15,16,17] no studies have been reported on EC CO2 RR over V-Zn hybrids although it shows electrochemical activity on water oxidation and hydrogen production [18,19]. From a practical point of view, EC CO2 RR is more feasible in the production of value-added products from CO2 than photocatalytic CO2 RR [20]. In EC CO2 RR over V-Zn hybrids and their oxides, CO and H2 (called syngas) are expected to be produced because Zn is known to produce CO by EC CO2 RR [3,4,21,22,23] and the Zn-V hybrid has shown EC activity on hydrogen production [19]. Guzman et al. prepared CuZnAl-oxide nanomaterials and demonstrated that H2/CO ratio could be tuned by applying potential [24]. Therefore, the syngas production performance needs to be tested for V-Zn hybrids. CH4, formate/formic acid, alcohols, and multi-carbon (Cn≥2) products are also included in CO2 reduction products [3,4,25,26,27,28,29]. Therefore, EC CO2 RR products and efficiencies are needed to be newly examined by modifying the V-Zn interface.
Motivated by this and the potential application of V-Zn hybrids to EC CO2 RR, we have directly prepared diverse interface-engineered V/Zn electrodes using Zn foil support. The hydrothermal method and sputter-deposition method were both employed, and the prepared electrodes were tested before and after thermal annealing. Zn sputter deposition was also performed on V support to examine the difference between V/Zn and Zn/V before and after thermal treatment. On the basis of CO2 RR products and Faradaic efficiencies tested over the systematically prepared electrodes, the roles of the V/Zn interface were further deeply discussed. The oxidation states of the diverse V-Zn hybrid electrodes were examined by X-ray photoelectron spectroscopy (XPS) before and after EC CO2 RR. The stability of the electrodes was found to be dependent on the preparation methods. Thus, the present unique results provide clues on the interface engineering of V-Zn hybrids for improving CO2 RR performance and stability, and strategy for the development of hybrid catalysts for energy and environment.

2. Materials and Methods

For the preparation of V-Zn hybrids by the hydrothermal method, a Zn foil (30 mm × 5 mm, 2 mm thick) was polished and cleaned by sonication in deionized water. A solution of 1 mM vanadium (V) was prepared by dissolving Na3VO4 (99.9%, Thermo Scientific, Waltham, MA, USA) in deionized water. 60 mL of 1 mM V solution was taken into a 100 mL-size Teflon-lined stainless autoclave and a pre-cleaned Zn foil was dipped in the solution. After that, the autoclave was tightly capped and placed in an oven set at 180 °C for 24 h. After completion of the reaction, and the autoclave was naturally cooled to laboratory temperature the Zn foil was removed, cleaned with deionized water, and dried under an infrared lamp. The V-Zn hybrids prepared by the hydrothermal method were abbreviated as V-Zn(H). The V-Zn(H) sample was thermally treated at 400 °C for 1 h, and abbreviated as V-Zn(H-h). Vanadium (V) was sputter-deposited on a Zn foil support using an SPT-20 ion sputter coater (COXEM Co., Daejeon, Korea) at an ionization current of 5 mA for 600 s and 2400 s, and abbreviated as V(600 s)/Zn(S) and V(2400 s)/Zn(S), respectively. Their thermal treated (at 400 °C for 1 h) samples were abbreviated as V(600 s)/Zn(S-h) and V(2400 s)/Zn(S-h), respectively. Zn was also sputter-deposited on a V foil support at 5 mA for 600 s and 2400 s, and abbreviated as Zn(600 s)/V(S) and Zn(2400 s)/V(S), respectively. Their thermal treated (at 400 °C for 1 h) samples were abbreviated as Zn(600 s)/V(S-h) and Zn(2400 s)/V(S-h), respectively.
The crystal phases of the V-Zn(H) sample were examined using a X-ray (Cu Kα radiation) diffractometer (MiniFlex II, Rigaku Corp., Tokyo, Japan) in CNU Chemistry Core Facility. The surface morphologies of the demonstrated samples were examined before and after EC CO2 RR using a scanning electron microscope (SEM, model S-4800, Hitachi Ltd., Tokyo, Japan) setting at 10.0 keV. For Raman spectra of V-Zn(H) sample before and after EC CO2 RR, a LabRAM HR-800 UV-Visible-NIR Raman spectrometer (Horiba Jobin Yvon, Kyoto, Japan) was used with experimental conditions of 514 nm laser light, a 100× objective, and the monochromator 1800 grating. For UV–visible absorption spectrum for V-Zn(H), a NeoSys-2000 double beam UV–visible spectrophotometer (SCINCO Co., Ltd., Seoul, Korea) was used with a diffuse reflectance mode. The oxidation states and the surface compositions of the demonstrated electrodes before and after EC CO2 RR were examined by taking X-ray photoelectron spectra (XPS) using a K-Alpha+ XPS spectrometer (Thermo-VG Scientific, Waltham, MA, USA) equipped with a hemispherical energy analyzer and a monochromated Al Kα X-ray (1486.6 eV) light.
Electrocatalytic CO2 reduction reaction (EC CO2 RR) experiments were conducted in a conventional three-electrode system; a 1 mm thick Pt counter electrode, an Ag/AgCl (3.0 M KCl) reference electrode, and a V-Zn hybrid electrode (30 mm × 5 mm) working electrode. The electrodes were connected to a ZIVE SP1 compact type Potentiostat/Galvanostat (WonATech Co., Ltd., Seoul, Korea). Electrolytes were KHCO3, K2CO3, NaHCO3, Na2CO3, and KH2PO4/K2HPO4 buffer with the desired concentration. A volume of 50 mL electrolyte was taken in a 100 mL size glass cell. Pure CO2 gas (99.999%) was fully bubbled in the electrolyte and charged in a tightly closed EC cell. Amperometry experiments were conducted at a fixed applied potential (vs. Ag/AgCl) under dark, photoirradiation (365 nm, 3000 mW/cm2), and thermal irradiation (830 nm, 250 mW) conditions.
Photocatalytic CO2 RR experiment was conducted in a closed stainless-steel chamber with a quartz window (47 mm) on top for UVC (200–290 nm) light irradiation (5.94 mW/cm2) to a V-Zn disc inside (45 mm diameter). Before the experiment, 20 μL of deionized water was placed beside the sample and the chamber was fully flushed and charged with CO2 gas (99.999%). After that, UVC light was irradiated for 6 h on the disc sample through the quartz window.
After the EC and photocatalytic CO2 RR experiments, gas and liquid products were examined by gas chromatography (GC) and nuclear magnetic resonance spectroscopy (NMR), respectively. For gas products, 0.5 mL volume of gas was taken from the EC cell (or the photocatalyst closed reactor) and injected into a GC system (YL 6500 GC, Young In Chromass Co., Ltd., Seoul, Korea). Diverse gas products were separated using two different columns of 40/60 Carboxen-1000 (Sigma-Aldrich, St. Louis, MO, USA) and HP-Plot Q PT (Agilent Technologies, Inc., Santa Clara, CA, USA). The GC system was equipped with a thermal conductivity detector, a flame ionization detector, and a Ni catalyst methanizer assembly. For the analysis of liquid products, an NMR spectrometer (600 MHz FT-NMR, AVANCE III, Bruker Corp., Billerica, MA, USA) was employed with 0.5 mL electrolyte and 0.1 mL of DMSO/D2O (v/v = 1:20,000) internal standard.

3. Results and Discussion

Figure 1a shows amperometry i-t curves over V/Zn(H) electrode obtained at various applied potentials of −1.6 V, −1.8 V, −2.0 V, and −2.2 V (vs. Ag/AgCl) for EC CO2 RR. The current density (CD) was increased with increasing the negative potential as expected. The CD was observed to be approximately 1.5 mA/cm2 and 6.5 mA/cm2 at −1.6 V and −2.0 V, respectively. The gas and liquid products were examined by taking GC and NMR profiles shown in Figure 1b,c, respectively. Major gas products were detected to be H2 and CO. A minor product includes CH4. A major liquid product was observed to be mainly formate. Other NMR signals were in impurity levels that were not discussed in detail. In Figure 1d, CO production was higher than H2 production at a lower negative applied potential; 2266 ppm and 28,643 ppm at −1.6 V and −2.0 V, respectively. The CO/H2 (syngas) ratio was estimated to be 2.5 at −1.6 V and decreased with increasing the negative potential; this reflects that the syngas ratio could be controlled from 2.5 to 0.4 by applying applied potential. In Figure 1e, CH4 was very meaningfully produced and increased with applied potential; 1.8 ppm at −1.6 V and 10.3 ppm at −2.0 V. Although their amounts were only below 1 ppm the productions of C2H4 and C2H6 were not negligible. Figure 1f shows the Faradaic efficiency (FE) of the detectable products with applied potentials. The FE of H2 was maximized to be 38.8% at −2.0 V, but the FE of CO was minimized to 15.4% at this applied potential. The maximum FE of CO was 30.9% at −1.8 V. The FE of formate was observed to be between 1.2% (at −1.6 V) and 2.4% (at −1.8 V). The FEs of other products were small; e.g., lower than 0.04% for CH4. For minor liquid products at −1.6 V, acetate and acetone were weakly detectable with FEs of 0.5% and 0.1%, respectively; however, these were not clearly observed at higher applied potentials. In Figure 1h, the image depicts the CO2 reduction process with major products of CO, formate, and CH4 over V-Zn hybrid material.
The V/Zn(H) electrode was further tested in diverse different electrolytes of KHCO3, K2CO3, NaHCO3, Na2CO3, and KH2PO4/K2HPO4 buffer with a concentration of 0.1 M (Figure 2(a,a1,a2)). All the FEs were observed to be highly dependent on the electrolyte. The FEs of H2 and CO were all decreased in K2CO3 (28.9% and 10.4%, respectively) and NaHCO3 (30.9% and 13.8%, respectively) electrolytes, compared with those (38.8% and 15.4%, respectively) in 0.1 M KHCO3 electrolyte (Figure 2a). The FE of H2 was increased in Na2CO3 (69.2%) and KH2PO4/K2HPO4 buffer (41.5%) electrolytes, but the FE of CO was decreased in the electrolytes with FEs of 1.2% and 6.3%, respectively. The FEs of formate were all decreased to 1.2%, 0%, 0%, and 0.3% in K2CO3, NaHCO3, Na2CO3, and KH2PO4/K2HPO4 buffer, respectively (Figure 2(a1)). Formate was not produced in NaHCO3 and Na2CO3 electrolytes; this indicates that Na+ plays a negative role in formate production, due to interaction between solvated cations and the adsorbed species [30]. CH4 production was commonly dependent on H2 production (Figure 2(a2)). The FE of CH4 was maximized to be 0.05% in 0.1 M Na2CO3 electrolyte where the FE of H2 was shown to be maximum. The production of C2H6 was higher than that of C2H4 and was observed to be between 0.002% (corresponding to 0.3 ppm) and 0.01% (corresponding to 0.7 ppm).
The V/Zn(H) electrode was tested in different concentrations of 0.1 M, 0.2 M, and 0.5 M KHCO3 electrolytes (Figure 2(b,b1,b2)). In 0.2 M concentration, CO2 RR performance was significantly improved, compared with the result in 0.1 M concentration. The FEs of CO and formate were dramatically increased to 23.2% by 1.5× and 6.0% by 2.9×, respectively, compared with those in 0.1 M concentration. The FE of H2 was instead decreased from 15.4% to 6.1%. In 0.5 M concentration, the FEs of CO and formate were observed to be only 6.1% and 1.9%, respectively, much smaller than those in 0.1 M and 0.2 M concentrations (Figure 2(b,b1)). The CH4 production amount (ppm) was linearly increased with increasing the concentration; 6.7 ppm, 20.0 ppm, and 30.7 ppm in 0.1 M, 0.2 M, and 0.5 M, respectively (inset in Figure 2(b2)). The FE of CH4 was maximized to be 0.07% in 0.2 M (Figure 2(b2)). C2H4 and C2H6 were not detected at all in these conditions.
The V/Zn(H) electrode was further tested at −2.0 V in 0.1 M KHCO3 electrolyte under other various conditions of dark, photoirradiation (365 nm), thermal irradiation (830 nm), photothermal irradiation (365 nm + 830 nm), Nafion-treated electrode, and H-cell condition (Figure 2(c,c1,c2)). The FE of CO was distinctly increased to 27.6% by 1.8× under photoirradiation (365 nm), compared with that under dark conditions (Figure 2c). The increase in CO under photoirradiation was in good consistent with the literature [3]. The FE was also increased to 23.2% by 1.5× when the electrode was treated with Nafion. The FE in the other conditions showed no dramatic effect and decreased to 9.0% in the H-cell condition. In the H-cell condition, the FE of H2 was instead increased to 42.1%. For the formate production (Figure 2(c1)), the FE was somewhat increased under photoirradiation and thermal irradiation conditions. The FE of formate was more significantly increased to 6.4% by 3.0×, compared with that under dark conditions. For the FE of CH4 production (Figure 2(c2)), CH4 production was enhanced under photoirradiation but decreased in Nafion-treated electrode and H-cell conditions. On the basis of the results, it was concluded that the CO2 RR performance was highly dependent on the experimental conditions.
The V/Zn(H) electrode interface was modified by thermal annealing at 400 °C. The consequent V/Zn(H-h) electrode was tested in 0.1 M KHCO3 electrolyte with various applied potentials of −1.6 V, −1.8 V, −2.0 V, and −2.2 V (vs. Ag/AgCl). In the amperometry i-t curves (Figure 3a), the final CD value at 3600 s was increased with increasing the potential; 1.0 mA/cm2, 2.3 mA/cm2, 4.6 mA/cm2, and 5.7 mA/cm2 at −1.6 V, −1.8 V, −2.0 V, and −2.2 V (vs. Ag/AgCl), respectively; however, for the initial and final CD values the i-t curve showed different behaviour at an applied potential of −2.2 V; this indicates that the V/Zn(H-h) electrode was oxidized after thermal annealing and more significantly altered during the initial EC at the highest applied potential. The gas and liquid products were examined by GC and NMR profiles shown in Figure 3b,c, respectively. Major gas products were detected to be H2 and CO that were increased with applied potential. Minor products include CH4, C2H4, and C2H6 (Figure 3b). A major liquid product was formate and others were unassignable (Figure 3c).
The H2 production was drastically increased with increasing potential; 2250 ppm and 29,115 ppm at −1.6 V and −2.2 V, respectively (Figure 3(d1)). CO production was sharply increased from 2693 ppm to 8952 ppm when the potential was increased from −1.6 V and −1.8 V; however, the CO production showed a sluggish increase above −1.8 V. As a consequence, CO production amount was higher than H2 production at −1.6 V and −1.8 V, but lower at −2.0 V and −2.2 V. In other words, the CO/H2 ratio was estimated to be 1.2 and 1.3 at −1.6 V and −1.8 V, respectively, but 0.84 and 0.41 at −2.0 V and −2.2 V, respectively. The syngas ratio was consequently tuned from 1.3 to 0.4. CH4 was only detected to be 0.4 ppm at −1.6 V, but detected to be 32–35 ppm at −1.8 V and −2.0 V. At a higher potential of −2.2 V, the CH4 amount was decreased to 8.1 ppm. As shown in Figure 3e, the FE of H2 was increased with applied potential and reached 28.0% at −2.2 V. The FE of CO at −1.8 V was maximized and observed to be 22.0%. The FE of CO decreased to 15.6% and 11.6% at higher potentials of −2.0 V and −2.2 V, respectively. The FE of formate was shown to be a maximum of 12.9% at −2.0 V. in Figure 3(e1), the FEs of CH4, C2H4, and C2H6 were observed to be maxima at −1.8 V.
The (FEV/Zn(H-h)−FEV/Zn(H))/FEV/Zn(H) ratios were estimated, where FEV/Zn(H-h) is the FE of the thermal-treated sample and FEV/Zn(H) is the FE of the as-prepared sample at an applied potential. In Figure 3f, the ratios for H2 and CO commonly showed negative values, indicating that the thermal-treated sample showed smaller FEs for H2 and CO productions. For the formate production, the ratios showed high values between 1.6 and 5.2, indicating that the thermal-treated sample showed higher performance for formate production. In Figure 3(f1), the ratios for CH4 production showed high positive values at −1.8 V and −2.0 V, but small negative values at −1.6 V and −2.2 V.
Another V/Zn interface was prepared by sputter-deposition of V on Zn foil support and also inversely sputter deposition of Zn on V foil support. Figure 4(a,a1) shows FEs of detectable gas and liquid products for as-sputtered V(2400 s)/Zn(S) electrode with various applied potentials of −1.6 V, −1.8 V, −2.0 V, and −2.2 V (vs. Ag/AgCl). The FEs of H2, CO, and formate were measured to be 20.9%, 15.5%, and 6.3% at −1.6 V, respectively (Figure 4a). The FE of H2 was increased at higher potentials, but those of CO and formate were decreased at higher potentials. The FE of formate at −1.6 V was approximately three times higher than those at higher applied potentials (Figure 4b). The FEs of CH4 and C2H6 are shown in the inset of Figure 4a. The (FEV/Zn(S)−FEV/Zn(H))/FEV/Zn(H) ratios were estimated (Figure 4(a2)), where FEV/Zn(S) is the FE of the sample by sputter deposition and FEV/Zn(H) is the FE of the sample by hydrothermal method at an applied potential. The ratios for H2, CH4, and formate showed positive value, indicating that the V/Zn electrode by sputter deposition method showed higher performances in H2, CH4, and formate productions at all applied potentials, compared with those over the V/Zn electrode by hydrothermal method. On the other hand, the ratio for CO production showed negative values at all applied potentials, indicating that the V/Zn electrode by sputter deposition showed poor performance than the V/Zn electrode by hydrothermal method.
The thickness of V and thermal treatment (400 °C for 1 h) effects were examined for the V/Zn(S) electrodes prepared by sputter deposition. As shown in Figure 4b, the FE of H2 was increased upon thermal treatment for V(600 s)/Zn electrode. On the other hand, the FE of H2 was increased upon thermal treatment for V(2400 s)/Zn electrode. For the FE of CO (Figure 4(b1)), the thermal treatment showed a negative effect for both V(600 s)/Zn(S) and V(2400 s)/Zn(S) electrodes. The as-sputtered V(600 s)/Zn(S) electrode showed an FE of 9.8% for CO, but decreased to 7.6% upon thermal treatment. The as-sputtered V(2400 s)/Zn(S) electrode showed an FE of 15.5% for CO, but decreased to 4.1% upon thermal treatment. For the FE of formate, the V(600 s)/Zn(S) electrode showed enhancement upon thermal treatment from 0.8% to 1.8%, but the V(2400 s)/Zn(S) electrode showed diminishment upon thermal treatment from 6.3% to 0.8%. Interestingly, the FE of C2H6 showed a large enhancement with a FE of 1.6%, compared with other samples (Figure 4(b2)). In addition, the FE was larger than that of CH4. Although it is not clear whether the experimental observation was meaningful. The FE of CH4 showed the highest FE of 0.14% for the V(600 s)/Zn(S-h) electrode (inset of Figure 4(b2)).
When Zn was used as a support material the EC CO2 RR showed higher FEs for CO and formate, as discussed above. Inversely, the Zn/V interface was prepared by sputter-deposition of Zn on V foil support and tested for EC CO2 RR. As shown in Figure 4c, H2 production was major, but CO production was dramatically diminished. The FE of formate was distinctly enhanced upon thermal treatment at 400 °C for 1 h. The FE of formate for Zn(600 s)/V(S) was increased from 0.6% to 4.7% (Figure 4(c1)). The FE of formate for Zn(2400 s)/V(S) was increased from 0.7% to 2.4%. The FE of CO was much lower than 1% (Figure 4(c2)). The FE of CO for Zn(600 s)/V(S) was increased from 0.1% to 0.3%. The FE of CO for Zn(2400 s)/V(S) was decreased from 0.4% to 0.06%. The FE of CO and CH4 showed the highest for the as-sputtered Zn(2400 s)/V(S) electrode; this indicates that Zn played a more important role in EC CO2 RR.
Generally, the total FE(%) was observed to be less than 100%. Several reasons appeared to be involved; the electrochemical current producing undetectable products, non-Faradaic current occurring at the metal oxide electrode surface, dissolution of gas products in electrolyte, and surface reduction reaction current of V-Zn oxides [3].
For direct comparison with the EC process, photocatalytic CO2 RR was tested for a selected sample of V/Zn(H) disc (Figure 5). The experiment was conducted in the gas (CO2 and H2O)-solid (catalyst) mode as depicted in the inset of Figure 5a [2,4]. Photocatalytic CO2 RR experiments over V-Zn hybrids in the literature have commonly been conducted in the liquid (bulk H2O with dissolved CO2)-solid (dispersed powder) mode [14,15,16,17]. Gas products were mainly detected because of the gas-phase mode. CO, CH4, C2H4, and C2H6 were clearly detected in the GC profiles separated by the Carboxene-1000 column (Figure 5(a,a1)). More interestingly, although the amounts were small Hydrocarbons (CnH2n and CnH2n+2, where n = 3–6) were detected in the GC profile separated by HP Plot Q PT column (Figure 5b), including CH3OH (MeOH). The production amounts of CO, CH4, and MeOH were measured to be 45.2 ppm, 9.4 ppm, and 15.6 ppm (Figure 5c). CO was more produced than CH4, in good with the literature using a similar catalyst system of V2O5/ZnV2O6 nanosheets [14]. The total amounts of C2-5 reached to approximately 3 ppm. In the photocatalytic CO2 RR, alkene (CnH2n) was produced more than alkanes (CnH2n+2). The C2H4/C2H6 ratio was observed to be 2.0. In the EC CO2 RR, the C2H6 production was more dominant than C2H4 (e.g., 6.7 ppm vs. 0.2 ppm at −2.0 V over V/Zn(H) electrode). The alkene/alkane ratios were estimated to be 1.7, 3.6 and 1.2 for C3, C4, and C5 hydrocarbons, respectively. The long-chain formation was understood by C-C coupling in the conventional Fisher-Tropsch synthesis [25,26], further discussed below.
Figure 6 displays SEM images (a, a1, a2, a3, and a4) of V/Zn(H) electrodes before and after EC at various applied potentials of −1.6 V, −1.8 V, −2.0 V, and −2.2 V (vs. Ag/AgCl); it was commonly observed that the surface morphology became changed and the surface became more drastically changed with increasing the applied potential. The surface reconstruction has commonly been reported during EC CO2 reduction at negative applied potentials, especially for oxide materials due to the reduction of surface oxide species [21,31]. XRD profiles (Figure 6b) were obtained for V/Zn(H) electrode and a ZnO/Zn reference electrode. XRD profiles of the two different samples were quite similar. Strong XRD patterns (closed squares, ■) were commonly observed at 2θ = 36.7°, 39.4°, 43.7°, and 54.7°, attributed to the (002), (010), (011), and (012) crystal planes of metallic Zn [32,33,34]; these signals were from the Zn support material. The other XRD signals (closed circles, ●) were also commonly observed at 2θ = 32.0°, 34.7°, 36.5°, 47.7°, and 56.9°, attributed to the (010), (002), (011), (012), and (110) crystal planes of hexagonal phase wurtzite ZnO [21,27,32]; this indicates that ZnO phase was commonly formed during the hydrothermal synthesis using metallic Zn support. No significant XRD profiles of V oxide and Zn-V oxides were observed, indicating that these species were ultrathin and/or amorphous; however, V species were clearly observed by XPS, discussed below. In Figure 6c, the Raman profile of the V/Zn(H) electrode showed mainly characteristics of ZnO [31]. The peaks around 435 cm−1 and 570 cm−1 were assigned to E2 high and E1(LO) modes, respectively [31]. The peaks were weakened after EC, due to surface reduction of V-Zn oxide species. SEM images were obtained for other demonstrated samples of V/Zn(H-h), V(2400 s)/Zn(S), V(2400 s)/Zn(S-h), and Zn(2400 s)/V(S-h) before and after EC in Figure 6(d,d1),(e,e1),(f,f1),(g,g1), respectively. The morphologies were different from those of V/Zn(H) electrodes as expected for different electrode materials.
XPS spectra were obtained to examine oxidation states, stability before and after EC, and differences in surface electronic structures before and after thermal treatment for the interface engineered V/Zn electrodes. Figure 7(a–a3) display Zn 2p, V 2p, O 1s, and VB XPS profiles, respectively for V/Zn(H) electrodes before and after EC CO2 RR at −1.6 V, −1.8 V, −2.0 V, and −2.2 V. For the Zn 2p XPS of as-prepared V/Zn(H) electrode before EC (Figure 7a), Zn 2p3/2 and Zn 2p1/2 peaks were observed at binding energies (BEs) of 1021.9 eV and 1044.9 eV, respectively with a spin-orbit (S-O) splitting energy of 23.0 eV; this is plausibly due to Zn(II) oxide species [14,35,36,37,38]. In Figure 7(a1), V 2p3/2 peaks could be resolved into two peaks at 517.1 eV (major) and 515.8 eV (shoulder), attributed to V5+ and V4+, respectively [6,7,14,37]. The V 2p3/2 peak at 517.1 eV was predominant. For the O 1s XPS of the as-prepared V/Zn(H) electrode sample (Figure 7(a2)), two broad peaks were observed at 530.5 eV and 532.6 eV, commonly attributed to lattice oxygen (OL) of Zn-V oxide species and surface oxygen species (e.g., Oad: OH/H2O, OH, and defects), respectively [7,14,15,39,40]. For the corresponding valence band (VB) profile (Figure 7(a3)), the VB edge was positioned at 3.0 eV below the Fermi level. The corresponding band gap was observed to be 3.1 eV (measured by UV-visible absorption spectrometer), indicating that the conduction band (CB) edge is positioned near the Fermi level. Two broad features were seen around 5.0 eV (A) and 8.0 eV, attributed to hybridizations of O 2p/Zn 3d/partially Zn 4p states and O 2p/Zn 3d/Zn 4s states, respectively [41]; it appears that O 2p state is highly contributed in the VB structure near the Femi level [13].
For the Zn 2p XPS after EC CO2 RR at −1.6 V, −1.8 V, and −2.0 V, the Zn 2p3/2 peak position was observed at a higher BE position of 1022.8 eV. V 2p peaks were somewhat weakened, but their BE positions were slightly altered (Figure 7(a1)). More clearly, the O 1s peak for OL was weakened while the O 1s for Oad became dominant. In the VB region, A and B peaks appeared to be merged one peak centred around 6 eV; it appears that the surface state changed to more like Zn(II)/V-OH species [35]. Upon applying the highest potential of −2.2 V, the Zn 2p3/2 peak was again shifted to a lower BE position of 1022.1 eV. The corresponding V 2p peak disappeared, indicating that the surface V species were dissolved and diffused into the bulk electrolyte during EC. The O 1s and VB XPS peaks became critically different from others, due to mainly surface states of V-free reduced ZnO/Zn(OH)2 [35,36,37,38].
For the XPS of thermal treated electrode of V/Zn(H-h) (Figure 7(b–b3)), Zn 2p3/2 and Zn 2p1/2 peaks were observed at binding energies (BEs) of 1022.1 eV and 1045.2 eV, respectively with a S-O splitting energy of 23.1 eV (Figure 7b); this is plausibly due to Zn(II)/V oxide species [35,36,37]. For the Zn 2p XPS after EC CO2 RR at −1.6 V, −1.8 V, −2.0 V, and −2.2 V, the Zn 2p3/2 peak position was observed at a higher BE position of 1023.1 eV, plausibly due to Zn(II)/V-OH species [35]. V 2p peaks were not seen (Figure 7(b1)), due to diffusion of V into the bulk state upon thermal treatment. For the O 1s XPS of V/Zn(H-h), two broad peaks were seen at 531.0 eV and 532.7 eV, attributed to lattice oxygen (OL) of Zn-V oxide species and surface oxygen species (e.g., Oad: OH/H2O, OH, and defects), respectively, as mentioned above [7,39,40]. After EC CO2 RR at −1.6 V, −1.8 V, −2.0 V, and −2.2 V, a broad peak around 533.2 eV became dominant, attributed to an increase in surface oxygen species such as OH. In the VB XPS profiles (Figure 7(b3)), two broad features (A and B) became more distinct and appeared to be merged one after EC CO2 RR, as mentioned above.
For the XPS profiles of sputtered V on Zn foil support before and after thermal treatment (Figure 7(c–c3)), a Zn 2p3/2 peak was observed at 1021.8 eV, due to the more metallic nature of Zn. The BE position showed no critical change after thermal treatment, due to the protection of overlayer V. After EC CO2 RR at −1.6 V, the BE position of the Zn 2p3/2 peak was commonly shifted to 1021.3 eV. The Zn 2p XPS profiles of V(600 s)/Zn(S) and V(2400 s)/Zn(S) were qualitatively similar before and after EC. A V 2p3/2 peak was dominantly seen at 516.9 eV, attributed to V5+ of V2O5 [6,7,37]. The V 2p of V(2400 s)/Zn(S) was stronger than that of V(600 s)/Zn(S) as expected. V 2p peaks became disappeared (Figure 7(c1)), due to diffusion of V into bulk state upon thermal treatment. For the V 2p XPS after EC, the peak intensity was quite weakened, especially for V(600 s)/Zn(S); however, the V 2p signal of V(2400 s)/Zn(S) was quite stable and broadened after EC. The BE distribution around 515.5 eV was due to V4+. The O 1s XPS profiles of V(600 s)/Zn(S) and V(2400 s)/Zn(S) were also qualitatively similar before and after EC (Figure 7(c2)). The O 1s XPS peaks of V(600 s)/Zn(S) and V(2400 s)/Zn(S) before thermal treatment were observed around 530.2 eV (530.0 eV) and 531.6 eV, due to lattice oxygen (OL) and surface oxygen species, respectively. After thermal treatment for V(600 s)/Zn(S-h) and V(2400 s)/Zn(S-h), two O 1s peaks were observed around 530.8 eV and 532.0 eV, attributed to lattice oxygen (OL) and surface oxygen species, respectively. After EC CO2 RR at −1.6 V, the lattice O 1s peak was weakened while the surface O 1s peak became stronger, as discussed above. For the VB profiles, a broad feature was commonly seen around 6 eV, attributed to hybridized O 2p states with Zn/V. The thermally treated sample of V(600 s)/Zn(S-h) and V(2400 s)/Zn(S-h) showed more clearly two features around 5.0 (A) and 8.0 eV (B), as mentioned above; these two merged one after EC, also discussed above.
On the basis of the products tested over diverse electrode samples, the CO2 RR mechanism was discussed below and depicted in Figure 8. In an electrolyte, H+, electrolyte cations/anions, and CO2 are commonly present. In CO2 RR, H+ is consumed for producing organic compounds, and negatively and competitively used for H2 production via H+ + e → Had followed by Had + Had → H2 or Had + H+ + e → H2 [3,4,5,39,40]. CO2 is adsorbed on an electrode surface commonly with two different bindings via CO2 + H+ + e → HCOOad or HOOCad [3,4]. The O or C of CO2 is hybridized with surface metal orbitals as depicted in Figure 8. When HCOOad is formed it may release into bulk electrolyte as formate. The hybridization forming HCOOad was observed to be enhanced by thermal annealing of both V/Zn(H) and Zn/V(S). In other words, formate production was further improved by a more V-Zn hybridized state. The fate HOOCad is transformed into OCad via HOOCad + H+ + e → OCad + H2O [3,4]. The OCad is then released into free CO or converted into other surface species. For the productions of minor CH4, C2H4, and C2H6, it requires a surface reaction process of OCad + H+ + eadCHO. The adCHO species proceed surface adCH2 and adCH3 via adCHO + 3H+ + 3eadCH2 + H2O and adCHO + 4H+ + 4eadCH3 + H2O, respectively [3,4]. The CH3 associates with H to form CH4 via CH3 + H → CH4 or CH3 + H+ + e → CH4 [42,43]. The surface CHx (x = 2, 3) may couple to generate free C2H4 and C2H6. The productions of CH4, C2H4, and C2H6 were significantly increased when V-Zn hybridization was increased by thermal annealing treatment. When the desorption of CO is easy the productions of CH4, C2H4, and C2H6 are expected to be reduced, or vice versa.
In electrochemical CO2 reduction, it is not easy to determine different active sites involved because of diverse surface species, electrode surface reconstruction, and the topmost element diffusion into a bulk electrolyte during electrochemical reaction [44,45]. For the different active sites, it needs to determine hydrogen adsorption, CO adsorption by the stripping method, and formate adsorption by the oxidation peak analysis [44,45]. For these V-Zn hybrids, the surface electronic structures are altered by different V/Zn ratios, the amounts and active site species become different, and thereby CO2 reduction performances are expected to be different. In the present study, two different d-block transition metal elements of V and Zn were used to examine EC CO2 reduction performances. To increase selectivity for a desired product, it is also necessary to employ different synthesis methods and other metal elements for hybrid systems; for example, Cu for C-C compounds, Au for CO, and p-block elements for formate [46,47,48,49].
For the CO2 RR products by photocatalysis, CH3OH and meaningful amounts of C3-5 compounds were newly detected. CO, CH4, C2H4, and C2H6 were common products as observed in electrocatalysis. In addition, the alkenes were more produced than the alkanes. For example, as discussed above, C2H4 was more produced than C2H6, unlike in electrocatalysis and conventional F-T synthesis. The alkene/alkane ratio has commonly been determined by the nature of a catalyst and the surface H/CO ratio at an experimental condition [50]. To confirm the alkene preference, we have also tested photocatalytic CO2 RR experiments over other transition metal elements and found similar results (not shown here). On the basis of these results, intuitively, because the neighbouring surface H in photocatalysis is much less than in electrocatalysis (and conventional high-pressurized conventional F-T synthesis) unsaturated alkenes are expected to be more produced in photocatalysis. Otherwise, the V-Zn surface may not be active for hydrogenation reaction; it needs further study to confirm the mechanism.
CH3OH is expected to be produced via adCHO + 3Had → CH3OH (g). For the production of long-chain C3-5 compounds, C-C coupling can be understood by (1) coupling of surface CHx and (2) CO insertion mechanism proposed in the conventional Fischer–Tropsch (F-T) synthesis [25,26]. In the coupling of surface CHx, the coupling of CHx occurs via CH3(CH2)xCadH2 + CHx → CH3(CH2)x+1CadH2 followed by association with surface H to liberate alkenes and alkanes. The CO insertion mechanism occurs via CH3(CH2)xCadH2 + CO + Had → CH3(CH2)x+1CadH2 + HOad. As a consequence, longer chain hydrocarbons were formed.

4. Conclusions

In summary, V-Zn hybrids were prepared by hydrothermal (H) and sputter-deposition (S) methods using a Zn foil support. For EC CO2 RR over V/Zn(H) electrode in 0.1 M KHCO3, CO (FE = 15–31%), H2 (FE = 11–39%), and formate (FE = 1.1–2.4%) were mainly produced with minor CH4, C2H4, and C2H6. Syngas (CO/H2) ratio was tuned from 2.5 to 0.4 by applying applied potential from −1.6 V to −2.2 V (vs. Ag/AgCl). The maximum FE of CO was 30.9% at −1.8 V in 0.1 M KHCO3 electrolyte. The KHCO3 electrolyte showed the best performance in CO and formate production, compared with other demonstrated electrolytes. In 0.2 M KHCO3 electrolyte, the FEs of CO, CH4, and formate were increased by 1.5×, 2.3×, and 2.9×, respectively, compared with those in 0.1 M KHCO3 electrolyte. CO and formate were further increased under photoirradiation conditions and by using a Nafion-treated electrode. For EC CO2 RR over V/Zn(H-h) electrode, formate production was dramatically increased by 160–520%, compared with those over the V/Zn(H) electrode, understood by an increased hybridization of V and Zn. CH4 production was increased by 450–1200%.
For EC CO2 RR over V/Zn(S) electrode in 0.1 M KHCO3, the FEs of CO and formate at −1.6 V were much higher than those at higher applied potentials. CH4 and formate were produced more in the V/Zn(S) electrode, compared with those in V/Zn(H) electrode. CO production was negated in the V/Zn(S) electrode by sputter deposition. CO production was negated by thermal treatment of the V/Zn(S) electrode. For Zn/V(S) electrodes, when V support was used CO production was drastically diminished and the FE of CO was much lower than 1%. The FE of formate was also lower than 1%. When Zn/V(S) electrode was thermally treated the FE of formate was increased to 2–5%, due to an increased hybridization of V and Zn; this is in good consistency with the V/Zn(H) electrode. The EC CO2 RR mechanism was proposed by two different initial bindings via CO2 + H+ + e → HCOOad or HOOCad. The hybridization between surface V/Zn and C and O may determine the production of CO, formate, and C2 compounds.
In photocatalysis over V/Zn(H), CH3OH, CO, and CH4 were mainly produced. Meaningful amounts of C2-6 hydrocarbons were first observed to be produced. The long chain formation was understood by the coupling of surface CHx and CO insertion mechanism in the F-T synthesis process. Alkenes were more produced than alkanes unlike in electrocatalysis and conventional F-T synthesis.
Overall, the present unique results on electrocatalysis and photocatalysis over interface engineered V-Zn hybrid materials provide deeper information on the development of V-Zn hybrid materials producing value-added carbon products from CO2 and H2O by electrocatalyst as well as photocatalysis for energy and the environment.

Author Contributions

Conceptualization, Y.S., S.Y.H. and H.J.J.; methodology, S.Y.H. and H.J.J.; validation, S.Y.H. and H.J.J.; formal analysis, S.Y.H. and H.J.J.; investigation, S.Y.H., H.J.J., Y.J.K., J.Y.M., G.E.P. and S.Y.Y.; data curation, S.Y.H.; writing—original draft preparation, Y.S. and S.Y.H.; writing—review and editing, Y.S. and C.K.R.; visualization, S.Y.H.; supervision, Y.S. and C.K.R.; project administration, Y.S.; funding acquisition, Y.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by a National Research Foundation of Korea (NRF) grant funded by the Korean government (MEST) (2021R1A2C2003929).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Usman, M.; Humayun, M.; Garba, M.D.; Ullah, L.; Zeb, Z.; Helal, A.; Suliman, M.H.; Alfaifi, B.Y.; Iqbal, N.; Abdinejad, M.; et al. Electrochemical Reduction of CO2: A Review of Cobalt Based Catalysts for Carbon Dioxide Conversion to Fuels. Nanomaterials 2021, 11, 2029. [Google Scholar] [CrossRef]
  2. Yang, A.Q.; Luo, J.L.Y.; Xie, Z. Synthesis and Applications of ZnV2O6 Nanomaterials. Ferroelectrics 2021, 581, 125–143. [Google Scholar] [CrossRef]
  3. Jang, H.J.; Yang, J.H.; Maeng, J.Y.; Joo, M.H.; Kim, Y.J.; Rhee, C.K.; Sohn, Y. Photoelectrochemical CO2 Reduction Products over Sandwiched Hybrid Ga2O3: ZnO/Indium/ZnO Nanorods. Front. Chem. 2022, 10, 814766. [Google Scholar] [CrossRef]
  4. Jang, H.J.; Yang, J.H.; Maeng, J.Y.; Joo, M.H.; Kim, Y.J.; Hong, S.M.; Rhee, C.K.; Sohn, Y. CO2 Reduction by Photocatalytic and Photoelectrocatalytic Approaches over Eu(III)-ZnGa2O4 Nanoparticles and Eu(III)-ZnGa2O4/ZnO Nanorods. J. CO2 Util. 2022, 60, 101994. [Google Scholar] [CrossRef]
  5. Sohn, Y.; Huang, W.; Taghipour, F. Recent Progress and Perspectives in the Photocatalytic CO2 Reduction of Ti-Oxide-Based Nanomaterials. Appl. Surf. Sci. 2017, 396, 1696–1711. [Google Scholar] [CrossRef]
  6. Li, S.; Qin, L.; Li, L.; Cheng, H.; Fang, G.; Liang, S.; Zhu, Q.; Chen, M. Porous Structure ZnV2O4/C-N Composite Activating Vanadium-based Cathode in Aqueous Zinc-ion Batteries. Mater. Today Commun. 2021, 27, 102271. [Google Scholar] [CrossRef]
  7. Luo, P.; Tang, W.; Cai, W.; Yang, J.; Zhang, W.; Zuo, C.; Liu, G.; Xiao, Y.; Dong, S. Charged-optimized ZnO/ZnV2O4 Composite Hollow Microspheres Robust Zinc-ion Storage Capacity. J. Solid State Chem. 2021, 301, 122371. [Google Scholar] [CrossRef]
  8. Ravindra, Y.S.; Puttaiah, S.H.; Yadav, S.; Prabagar, J.S. Evaluation of Polymeric g-C3N4 Contained Nonhierarchical ZnV2O6 Composite for Energy-efficient LED Assisted Photocatalytic Mineralization of Organic Pollutant. J. Mater. Sci. Mater. Electron. 2020, 31, 16806–16818. [Google Scholar] [CrossRef]
  9. Sameie, H.; Alvani, A.A.S.; Naseri, N.; Rosei, F.; Mul, G.; Mei, B.T. Photocatalytic Activity of ZnV2O6/Reduced Graphene Oxide Nanocomposite: From Theory to Experiment. J. Electrochem. Soc. 2018, 165, 353–359. [Google Scholar] [CrossRef]
  10. Ghodrati, M.; Mousavi-Kamazani, M.; Zinatloo-Ajabshir, S. Zn3V3O8 Nanostructures: Facile Hydrothermal/Solvothermal Synthesis, Characterization, and Electrochemical Hydrogen Storage. Ceram. Int. 2020, 46, 28894–28902. [Google Scholar] [CrossRef]
  11. Sajid, M.M.; Shad, N.A.; Khan, S.B.; Zhang, Z.J.; Amin, N. Facile Synthesis of Zinc Vanadate Zn3(VO4)2 for Highly Efficient Visible Light Assisted Photocatalytic Activity. J. Alloys Compd. 2019, 775, 281–289. [Google Scholar] [CrossRef]
  12. Pei, L.Z.; Lin, N.; Wei, T.; Liu, H.D.; Yu, H.Y. Zinc Vanadate Nanorods and their Visible light Photocatalytic Activity. J. Alloys Compd. 2015, 631, 90–98. [Google Scholar] [CrossRef]
  13. Yang, A.; Luo, J.; Xie, Z.; Chen, Q. Photocatalytic Activity of V2O5/ZnV2O6 Catalysts and its Origin: Insights into Enhanced Photocatalytic Mechanisms via DFT study. Appl. Surf. Sci. 2022, 599, 153894. [Google Scholar] [CrossRef]
  14. Li, S.P.; Hasan, N.; Ma, H.X.; Zhu, G.Q.; Pan, L.K.; Zhang, F.C.; Son, N.Y.; Kang, M.; Liu, C.L. Hierarchical V2O5/ZnV2O6 Nanosheets Photocatalyst for CO2 Reduction to Solar Fuels. Chem. Eng. J. 2022, 430, 132863. [Google Scholar] [CrossRef]
  15. Bafaqeer, A.; Tahir, M.; Amin, N.A.S. Well-Designed ZnV2O6/g-C3N4 2D/2D Nanosheets Heterojunction with Faster Charges Separation via pCN as Mediator towards Enhanced Photocatalytic Reduction of CO2 to Fuels. Appl. Catal. B Environ. 2019, 242, 312–326. [Google Scholar] [CrossRef]
  16. Bafaqeer, A.; Tahir, M.; Amin, N.A.S. Synthesis of Hierarchical ZnV2O6 Nanosheets with Enhanced Activity and Stability for Visible Light Driven CO2 Reduction to Solar Fuels. Appl. Surf. Sci. 2018, 435, 953–962. [Google Scholar] [CrossRef]
  17. Tahir, M. Hierarchical 3D VO2/ZnV2O4 Microspheres as an Excellent Visible Light Photocatalyst for CO2 Reduction to Solar Fuels. Appl. Surf. Sci. 2019, 467–468, 1170–1180. [Google Scholar] [CrossRef]
  18. Sameie, H.; Sabbagh Alvani, A.A.; Naseri, N.; Du, S.; Rosei, F. First-Principles Study on ZnV2O6 and Zn2V2O7: Two Nnew Photoanode Candidates for Photoelectrochemical Water Oxidation. Ceram. Int. 2018, 44, 6607–6613. [Google Scholar] [CrossRef]
  19. Sameiea, H.; Sabbagh Alvani, A.A.; Mei, B.T.; Salimi, R.; Poelman, D.; Rosei, F. Mo-doped ZnV2O6/Reduced Graphene Oxide Photoanodes for Solar Hydrogen Production. Electrochim. Acta 2021, 382, 138333. [Google Scholar] [CrossRef]
  20. Ješić, D.; Jurković, D.L.; Pohar, A.; Suhadolnik, L.; Likozar, B. Engineering Photocatalytic and Photoelectrocatalytic CO2 Reduction Reactions: Mechanisms, Intrinsic Kinetics, Mass Transfer Resistances, Reactors and Multi-Scale Modelling Simulations. Chem. Eng. J. 2021, 407, 126799. [Google Scholar] [CrossRef]
  21. Xue, L.; Zhang, A.; Wu, J.; Wang, Q.; Liu, Y.; Zhao, Y.; Liu, S.; Liu, Z.; Li, P.; Zeng, S. Surface Modification and Reconstruction of ZnO Hollow Microspheres for Selective Electroreduction of CO2 to CO. J. Alloys Compd. 2021, 882, 160703. [Google Scholar] [CrossRef]
  22. Kim, J.; Kim, H.; Han, G.H.; Ahn, S.H. Solution-Phase-Reconstructed Zn-based Nanowire Electrocatalysts for Electrochemical Reduction of Carbon Dioxide to Carbon Monoxide. Int. J. Energy Res. 2021, 45, 7987–7997. [Google Scholar] [CrossRef]
  23. Wang, H.; Yang, D.; Yang, J.; Ma, X.; Li, H.; Dong, W.; Zhang, R.; Feng, C. Efficient Electroreduction of CO2 to CO on Porous ZnO Nanosheets with Hydroxyl Groups in Ionic Liquid-based Electrolytes. ChemCatChem 2021, 13, 2570–2576. [Google Scholar] [CrossRef]
  24. Guzmán, H.; Roldán, D.; Sacco, A.; Castellino, M.; Fontana, M.; Russo, N.; Hernández, S. CuZnAl-Oxide Nanopyramidal Mesoporous Materials for the Electrocatalytic CO2 Reduction to Syngas: Tuning of H2/CO Ratio. Nanomaterials 2021, 11, 3052. [Google Scholar] [CrossRef]
  25. Teimouri, Z.; Abatzoglou, N.; Dalai, A.K. Kinetics and Selectivity Study of Fischer–Tropsch Synthesis to C5+ Hydrocarbons: A Review. Catalysts 2021, 11, 330. [Google Scholar] [CrossRef]
  26. Mahmoudi, H.; Mahmoudi, M.; Doustdar, O.; Jahangiri, H.; Tsolakis, A.; Gu, S.; Lechwyszynski, M. A review of Fischer Tropsch synthesis process, mechanism, surface chemistry and catalyst formulation. BFUEL 2017, 2, 11–31. [Google Scholar] [CrossRef]
  27. Du, C.; Wang, X.; Chen, W.; Feng, S.; Wen, J.; Wu, Y.A. CO2 Transformation to Multicarbon Products by Photocatalysis and Electrocatalysis. Mater. Today Adv. 2020, 6, 100071. [Google Scholar] [CrossRef]
  28. Xie, S.; Ma, W.; Wu, X.; Zhang, H.; Zhang, Q.; Wang, Y.; Wang, Y. Photocatalytic and Electrocatalytic Transformations of C1 Molecules involving C–C Coupling. Energy Environ. Sci. 2021, 14, 37–89. [Google Scholar] [CrossRef]
  29. Ye, W.; Guo, X.; Ma, T. A Review on Electrochemical Synthesized Copper-based Catalysts for Electrochemical Reduction of CO2 to C2+ Products. Chem. Eng. J. 2021, 414, 128825. [Google Scholar] [CrossRef]
  30. Resasco, J.; Chen, L.D.; Clark, E.; Tsai, C.; Hahn, C.; Jaramillo, T.F.; Chan, K.; Bell, A.T. Promoter Effects of Alkali Metal Cations on the Electrochemical Reduction of Carbon Dioxide. J. Am. Chem. Soc. 2017, 139, 11277–11287. [Google Scholar] [CrossRef]
  31. Luo, W.; Zhang, Q.; Zhang, J.; Moioli, E.; Zhao, K.; Züttel, A. Electrochemical Reconstruction of ZnO for Selective Rreduction of CO2 to CO. Appl. Catal. B Environ. 2020, 273, 119060. [Google Scholar] [CrossRef]
  32. Zhang, T.; Li, X.; Qiu, Y.; Su, P.; Xu, W.; Zhong, H.; Zhang, H. Multilayered Zn Nanosheets as an Electrocatalyst for Efficient Electrochemical Reduction of CO2. J. Catal. 2018, 357, 154–162. [Google Scholar] [CrossRef]
  33. Lu, Y.; Han, B.; Tian, C.; Wu, J.; Geng, D.; Wang, D. Efficient Electrocatalytic Eeduction of CO2 to CO on an Electrodeposited Zn Porous Network. Electrochem. Commun. 2018, 97, 87–90. [Google Scholar] [CrossRef]
  34. Qin, B.; Li, Y.; Fu, H.; Wang, H.; Chen, S.; Liu, Z.; Peng, F. Electrochemical Reduction of CO2 into Tunable Syngas Production by Regulating the Crystal Facets of Earth-abundant Zn Catalyst. ACS Appl. Mater. Interfaces 2018, 10, 20530–20539. [Google Scholar] [CrossRef]
  35. Pradhan, D.; Leung, K.T. Controlled Growth of Two-dimensional and One-dimensional ZnO Nanostructures on Indium Tin Oxide coated Glass by Direct Electrodeposition. Langmuir 2008, 24, 9707–9716. [Google Scholar] [CrossRef]
  36. Rahman, A.; Tan, A.L.; Harunsani, M.H.; Ahmad, N.; Hojamberdiev, M.; Khan, M.M. Visible Light Induced Antibacterial and Antioxidant Studies of ZnO and Cu-doped ZnO Fabricated using Aqueous Leaf Extract of Ziziphus mauritiana Lam. J. Environ. Chem. Eng. 2021, 9, 105481. [Google Scholar] [CrossRef]
  37. Aliaga, J.; Cifuentes, N.; González, G.; Sotomayor-Torres, C.; Benavente, E. Enhancement Photocatalytic Activity of the Heterojunction of Two-Dimensional Hybrid Semiconductors ZnO/V2O5. Catalysts 2018, 8, 374. [Google Scholar] [CrossRef]
  38. Naumkin, A.V.; Kraut-Vass, A.; Gaarenstroom, S.W.; Powell, C.J. NIST Standard Reference Database 20, Version 4.1. 2012. Available online: https://srdata.nist.gov/xps/ (accessed on 1 June 2022).
  39. Yoon, H.J.; Yang, J.H.; Park, S.J.; Rhee, C.K.; Sohn, Y. Photocatalytic CO2 Reduction and Hydrogen Production over Pt/Zn-Embedded β- Ga2O3 Nanorods. Appl. Surf. Sci. 2021, 536, 147753. [Google Scholar] [CrossRef]
  40. Yang, J.H.; Park, S.J.; Rhee, C.K.; Sohn, Y. Photocatalytic CO2 Reduction and Electrocatalytic H2 Evolution over Pt (0,II,IV)-Loaded Oxidized Ti Sheets. Nanomaterials 2020, 10, 1909. [Google Scholar] [CrossRef]
  41. Hung-Chun Lai, H.; Basheer, T.; Kuznetsov, V.L.; Egdell, R.G.; Jacobs, R.M.; Pepper, M.; Edwards, P.P. Dopant-Induced Bandgap Shift in Al-doped ZnO Thin Films Prepared by Spray Pyrolysis. J. Appl. Phys. 2012, 112, 083708. [Google Scholar] [CrossRef]
  42. Yang, J.H.; Park, S.J.; Hong, S.-M.; Kim, J.; Yee, K.-J.; Jung, H.; Sohn, Y. Ultraviolet and Infrared Light Decontamination and the Secondary Pollution Products of G-series Nerve Agent Simulant Model Molecules Contaminating TiO2/Ti Surfaces. J. Ind. Eng. Chem. 2021, 100, 75–91. [Google Scholar] [CrossRef]
  43. Yang, J.H.; Kim, T.; Park, S.J.; Hong, S.-M.; Kim, J.; Yee, K.-J.; Jung, H.; Sohn, Y. UV Light Decontamination and Secondary Pollution Products of G-series Chemical Warfare Model Molecules Contaminated on Mo and MoOx Surfaces. Appl. Surf. Sci. 2021, 565, 150502. [Google Scholar] [CrossRef]
  44. Ling, Y.; Luo, F.; Zhang, Q.; Qu, K.; Guo, L.; Hu, H.; Yang, Z.; Cai, W.; Cheng, H. Tungsten Carbide Hollow Microspheres with Robust and Stable Electrocatalytic Activity toward Hydrogen Evolution Reaction. ACS Omega 2019, 4, 4185–4191. [Google Scholar] [CrossRef]
  45. Huang, W.; Wang, Y.; Liu, J.; Wang, Y.; Liu, D.; Dong, J.; Jia, N.; Liu, C.; Liu, Z.; Liu, B.; et al. Efficient and Selective CO2 Reduction to Formate on Pd-doped Pb3(CO3)2(OH)2: Dynamic Catalyst Reconstruction and Accelerated CO2 Protonation. Small 2022, 18, 2107885. [Google Scholar] [CrossRef]
  46. Zhao, S.; Liu, A.; Li, Y.; Wen, Y.; Gao, X.; Chen, Q. Boosting the Electrocatalytic CO2 Reduction Reaction by Nanostructured Metal Materials via Defects Engineering. Nanomaterials 2022, 12, 2389. [Google Scholar] [CrossRef]
  47. Fan, L.; Xia, C.; Yang, F.; Wang, J.; Wang, H.; Lu, Y. Strategies in catalysts and electrolyzer design for electrochemical CO2 reduction toward C2+ products. Sci. Adv. 2020, 6, eaay3111. [Google Scholar] [CrossRef]
  48. Yang, Z.; Oropeza, F.E.; Zhang, K.H.L. P-block Metal-based (Sn, In, Bi, Pb) Electrocatalysts for Selective Reduction of CO2 to Formate. APL Mater. 2020, 8, 060901. [Google Scholar] [CrossRef]
  49. Alfonso, D. High Performance Ligand-Protected Metal Nanocluster Catalysts for CO2 Conversion through the Exposure of Undercoordinated Sites. Catalysts 2022, 12, 505. [Google Scholar] [CrossRef]
  50. Dry, M.E.; Steynberg, A.P. Chapter 5 Commercial FT process applications. Fischer-Tropsch Technology. In Studies in Surface Science and Catalysis; Steynberg, A., Dry, M., Eds.; Elsevier: Amsterdam, The Netherlands, 2004; Volume 152, pp. 406–481. [Google Scholar]
Figure 1. Amperometry i-t curves over V-Zn(H) hybrid electrodes at applied potentials of −1.6 V, −1.8 V, −2.0 V, and −2.2 V (vs. Ag/AgCl) (a). GC profiles after the amperometry tests (b). NMR spectra for liquid products after the amperometry tests (c). CO and H2 production amount with applied potentials (d). CH4, C2H4, and C2H6 amounts with applied potentials (e). Corresponding FEs including liquid products ((f) and (g), respectively). Scheme for CO2 reduction process over V-Zn hybrid material (h).
Figure 1. Amperometry i-t curves over V-Zn(H) hybrid electrodes at applied potentials of −1.6 V, −1.8 V, −2.0 V, and −2.2 V (vs. Ag/AgCl) (a). GC profiles after the amperometry tests (b). NMR spectra for liquid products after the amperometry tests (c). CO and H2 production amount with applied potentials (d). CH4, C2H4, and C2H6 amounts with applied potentials (e). Corresponding FEs including liquid products ((f) and (g), respectively). Scheme for CO2 reduction process over V-Zn hybrid material (h).
Nanomaterials 12 02758 g001
Figure 2. FEs of CO2 RR products over V-Zn(H) electrode tested in various electrolytes (a,a1,a2). FEs tested in KHCO3 electrolyte with different concentrations of 0.1 M, 0.2 M, and 0.5 M (b,b1,b2). FEs in 0.1 M KHCO3 electrolyte tested under dark, photoirradiation, thermal irradiation, photothermal irradiation, Nafion-treated electrode, and H-cell condition (c,c1,c2).
Figure 2. FEs of CO2 RR products over V-Zn(H) electrode tested in various electrolytes (a,a1,a2). FEs tested in KHCO3 electrolyte with different concentrations of 0.1 M, 0.2 M, and 0.5 M (b,b1,b2). FEs in 0.1 M KHCO3 electrolyte tested under dark, photoirradiation, thermal irradiation, photothermal irradiation, Nafion-treated electrode, and H-cell condition (c,c1,c2).
Nanomaterials 12 02758 g002
Figure 3. Amperometry i-t curves over V-Zn(H-h) electrode at various applied potentials of −1.6 V, −1.8 V, −2.0 V, and −2.2 V (vs. Ag/AgCl) in 0.1 M KHCO3 electrolyte (a). Corresponding GC (b) and NMR (c) profiles for gas and liquid products, respectively. Measured H2 and CO product amounts (d1). CH4, C2H4, and C2H6 amounts (d2). Corresponding FEs of gas and liquid products (e,e1). (FEV/Zn(H-h)−FEV/Zn(H))/FEV/Zn(H) ratios for H2, CO, and formate productions (f). FEV/Zn(H-h)−FEV/Zn(H))/FEV/Zn(H) ratios for CH4 production (f1).
Figure 3. Amperometry i-t curves over V-Zn(H-h) electrode at various applied potentials of −1.6 V, −1.8 V, −2.0 V, and −2.2 V (vs. Ag/AgCl) in 0.1 M KHCO3 electrolyte (a). Corresponding GC (b) and NMR (c) profiles for gas and liquid products, respectively. Measured H2 and CO product amounts (d1). CH4, C2H4, and C2H6 amounts (d2). Corresponding FEs of gas and liquid products (e,e1). (FEV/Zn(H-h)−FEV/Zn(H))/FEV/Zn(H) ratios for H2, CO, and formate productions (f). FEV/Zn(H-h)−FEV/Zn(H))/FEV/Zn(H) ratios for CH4 production (f1).
Nanomaterials 12 02758 g003
Figure 4. FEs of gas and liquid products over V/Zn(S) electrode at various applied potentials of −1.6 V, −1.8 V, −2.0 V, and −2.2 V (vs. Ag/AgCl) in 0.1 M KHCO3 electrolyte (a,a1). (FEV/Zn(S)-FEV/Zn(H))/FEV/Zn(H) ratios for H2, CO, CH4, and formate productions (a2). FEs of gas and liquid products over V(600 s)/Zn(S), V(600 s)/Zn(S-h), V(2400 s)/Zn(S), V(2400 s)/Zn(S-h) at −1.6 V in 0.1 M KHCO3 electrolyte, and over V(2400 s)/Zn(S) at −1.8 V (b,b1,b2). FEs of gas and liquid products over Zn(600 s)/V(S), Zn(600 s)/V(S-h), Zn(2400 s)/V(S), Zn(2400 s)/V(S-h) at −1.8 V in 0.1 M KHCO3 electrolyte (c,c1,c2).
Figure 4. FEs of gas and liquid products over V/Zn(S) electrode at various applied potentials of −1.6 V, −1.8 V, −2.0 V, and −2.2 V (vs. Ag/AgCl) in 0.1 M KHCO3 electrolyte (a,a1). (FEV/Zn(S)-FEV/Zn(H))/FEV/Zn(H) ratios for H2, CO, CH4, and formate productions (a2). FEs of gas and liquid products over V(600 s)/Zn(S), V(600 s)/Zn(S-h), V(2400 s)/Zn(S), V(2400 s)/Zn(S-h) at −1.6 V in 0.1 M KHCO3 electrolyte, and over V(2400 s)/Zn(S) at −1.8 V (b,b1,b2). FEs of gas and liquid products over Zn(600 s)/V(S), Zn(600 s)/V(S-h), Zn(2400 s)/V(S), Zn(2400 s)/V(S-h) at −1.8 V in 0.1 M KHCO3 electrolyte (c,c1,c2).
Nanomaterials 12 02758 g004
Figure 5. GC profiles separated by Carboxene-1000 column (a,a1) after photocatalytic CO2 RR for 6 h under UVC irradiation. Inset is a schematic of a closed stainless-steel reactor. GC profile separated by HP Plot Q PT column (b). Consequent photocatalytic CO2 RR product amounts (c).
Figure 5. GC profiles separated by Carboxene-1000 column (a,a1) after photocatalytic CO2 RR for 6 h under UVC irradiation. Inset is a schematic of a closed stainless-steel reactor. GC profile separated by HP Plot Q PT column (b). Consequent photocatalytic CO2 RR product amounts (c).
Nanomaterials 12 02758 g005
Figure 6. SEM images of V/Zn(H) electrode before (a) and after EC at various applied potentials (a1a4). XRD profiles (b) of V/Zn(H) and ZnO/Zn (reference). Raman spectra of V/Zn(H) electrode before and after EC at −2.0 V (c). V/Zn(H-h) (d,d1), V(2400 s)/Zn(S) (e,e1), V(2400 s)/Zn(S-h) (f,f1), and Zn(2400 s)/V(S-h) (g,g1) electrodes before and after EC.
Figure 6. SEM images of V/Zn(H) electrode before (a) and after EC at various applied potentials (a1a4). XRD profiles (b) of V/Zn(H) and ZnO/Zn (reference). Raman spectra of V/Zn(H) electrode before and after EC at −2.0 V (c). V/Zn(H-h) (d,d1), V(2400 s)/Zn(S) (e,e1), V(2400 s)/Zn(S-h) (f,f1), and Zn(2400 s)/V(S-h) (g,g1) electrodes before and after EC.
Nanomaterials 12 02758 g006
Figure 7. Zn 2p, V 2p, O 1s, and VB XPS profiles for V/Zn(H) electrode before and after tested at various applied potentials ((aa3), respectively), V/Zn(H-h) electrode before and after tested at various applied potentials ((bb3), respectively), and V(600 s)/Zn(S), V(600 s)/Zn(S-h), V(2400 s)/Zn(S), V(2400 s)/Zn(S-h) electrodes before and after EC (shaded area) at −1.6 V in 0.1 M KHCO3 electrolyte ((cc3), respectively).
Figure 7. Zn 2p, V 2p, O 1s, and VB XPS profiles for V/Zn(H) electrode before and after tested at various applied potentials ((aa3), respectively), V/Zn(H-h) electrode before and after tested at various applied potentials ((bb3), respectively), and V(600 s)/Zn(S), V(600 s)/Zn(S-h), V(2400 s)/Zn(S), V(2400 s)/Zn(S-h) electrodes before and after EC (shaded area) at −1.6 V in 0.1 M KHCO3 electrolyte ((cc3), respectively).
Nanomaterials 12 02758 g007
Figure 8. Proposed CO2 RR mechanism on the basis of the products.
Figure 8. Proposed CO2 RR mechanism on the basis of the products.
Nanomaterials 12 02758 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

Hwang, S.Y.; Jang, H.J.; Kim, Y.J.; Maeng, J.Y.; Park, G.E.; Yang, S.Y.; Rhee, C.K.; Sohn, Y. Interface Engineered V-Zn Hybrids: Electrocatalytic and Photocatalytic CO2 Reductions. Nanomaterials 2022, 12, 2758. https://doi.org/10.3390/nano12162758

AMA Style

Hwang SY, Jang HJ, Kim YJ, Maeng JY, Park GE, Yang SY, Rhee CK, Sohn Y. Interface Engineered V-Zn Hybrids: Electrocatalytic and Photocatalytic CO2 Reductions. Nanomaterials. 2022; 12(16):2758. https://doi.org/10.3390/nano12162758

Chicago/Turabian Style

Hwang, Seon Young, Hye Ji Jang, Young Jun Kim, Ju Young Maeng, Go Eun Park, Seo Young Yang, Choong Kyun Rhee, and Youngku Sohn. 2022. "Interface Engineered V-Zn Hybrids: Electrocatalytic and Photocatalytic CO2 Reductions" Nanomaterials 12, no. 16: 2758. https://doi.org/10.3390/nano12162758

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