Next Article in Journal
Silver-Carbonaceous Microsphere Precursor-Derived Nano-Coral Ag Catalyst for Electrochemical Carbon Dioxide Reduction
Next Article in Special Issue
Is Black Titania a Promising Photocatalyst?
Previous Article in Journal
Effect of Different Carbon Supports on the Activity of PtNi Bimetallic Catalysts toward the Oxygen Reduction
Previous Article in Special Issue
Uncovering the Mechanism of the Hydrogen Poisoning on Ru Nanoparticles via Density Functional Theory Calculations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cu Modified TiO2 Catalyst for Electrochemical Reduction of Carbon Dioxide to Methane

1
International Institute for Carbon-Neutral Energy Research (WPI-I2CNER), Kyushu University, 744 Motooka, Nishi-ku, Fukuoka 819-0395, Japan
2
Department of Chemistry, Graduate School of Science, Kyushu University, 744 Motooka, Nishi-ku, Fukuoka 819-0395, Japan
3
RIKEN SPring-8 Center, 1-1-1 Kouto, Sayo-cho, Sayo-gun 679-5148, Japan
4
Research Center for Synchrotron Light Applications, Kyushu University, 6-1 Kasuga-koen, Kasuga 816-8580, Japan
5
Institute for Materials Chemistry and Engineering (IMCE), Kyushu University, 744 Motooka, Nishi-ku, Fukuoka 819-0395, Japan
6
Advanced Institute for Materials Research (WPI-AIMR), Tohoku University, 2-1-1 Katahira, Aoba-ku, Sendai 980-8577, Japan
7
Research Center for Negative Emissions Technologies (K-Nets), Kyushu University, 744 Motooka, Nishi-ku, Fukuoka 819-0395, Japan
*
Author to whom correspondence should be addressed.
Submission received: 11 March 2022 / Revised: 6 April 2022 / Accepted: 21 April 2022 / Published: 23 April 2022
(This article belongs to the Special Issue Role of Defects and Disorder in Catalysis)

Abstract

:
Electrochemical reduction of CO2 (ECO2R) is gaining attention as a promising approach to store excess or intermittent electricity generated from renewable energies in the form of valuable chemicals such as CO, HCOOH, CH4, and so on. Selective ECO2R to CH4 is a challenging target because the rate-determining step of CH4 formation, namely CO* protonation, competes with hydrogen evolution reaction and the C–C coupling toward the production of longer-chain chemicals. Herein, a Cu-TiO2 composite catalyst consisting of CuOx clusters or Cu nanoparticles (CuNPs), which are isolated on the TiO2 grain surface, was synthesized using a one-pot solvothermal method and subsequent thermal treatment. The Cu-TiO2 catalyst exhibited high selectivity for CH4, and the ratio of FE for CH4 to total FE for all products in ECO2R reached 70%.

1. Introduction

Electrochemical reduction of CO2 (ECO2R), which uses renewable electricity to produce fuels and chemical feedstocks from CO2, has attracted attention not only as an eco-friendly material synthesis process but also as a novel method to store intermittent renewable electricity which is generated from renewable energies such as solar, wind power, and so on [1]. Among metal catalysts, Cu is known to exhibit the highest activity in ECO2R [2] and also produces multi-carbon products such as hydrocarbons and alcohols. Meanwhile, the selective production of CH4, which is widely used as a fuel, is a demanding task but has not been achieved to the desired level in ECO2R employing active Cu catalysts.
CH4 formation proceeds through the addition of 8 electrons and 8 protons to CO2 (Equation (1)).
CO2 + 8H+ + 8e → CH4 + 2H2O
In this process, CH4 formation is thought to occur via *CHO formation by protonation of *CO [3]. The *CHO formation competes with both the C–C coupling of two *CO and hydrogen evolution reaction (HER), which significantly lowers selectivity for the production of CH4. Recently, the formation of isolated Cu sites has been found to effectively improve the selectivity for the CH4 production by suppressing the unfavorable C–C coupling [4,5]. However, the selectivity for the production of CH4 from CO2 (CO2 to CH4 selectivity) should be improved. The increase of surface area is found to efficiently suppresses the Cu agglomeration on a support material such as carbons [6] and oxides [4,7,8]. However, the formation relatively large Cu portions where C–C couplings preferentially progress cannot be suppressed in conventional impregnation synthesis.
Titanium dioxide (TiO2) is a multifunctional material with many advantages, being ubiquitous, low-cost, and environmentally friendly. TiO2 is a widely used catalyst material for various application [9] due to its high activity and adequate stability. Recently, we have uncovered relatively high overpotentials for electrochemical H2 evolution and the favorable interactions of oxygen species included in organic acids, oximes, and imines with the surface of TiO2-based catalysts for electrochemical hydrogenation [10,11,12,13,14,15,16,17,18], which could enhance the selectivity for the production of CO2-derived chemicals in ECO2R. Highly dispersed and isolated Cu sites on TiO2 are expected to suppress the generation of multi-carbon products and show high selectivity for CH4 synthesis. Thus, we develop Cu-TiO2 catalysts presenting isolated Cu sites with high dispersion for the selective CH4 production by the electrochemical reduction of CO2.

2. Results

To enhance their dispersivity, Cu-TiO2 catalysts were prepared via a one-pot solvothermal method (See Section 4). The crystal structures of the prepared TiO2, Cu-TiO2, and Cu-TiO2 treated with H2 (Cu-TiO2-H) were examined by X-ray diffraction (XRD) (Figure 1 and Figure S1). All samples showed a diffraction pattern mostly attributable to an anatase phase of TiO2, but a very small peak from a brookite phase was also observed at 14.7°. The intensity of the brookite peak of Cu-TiO2 seemed slightly large compared to that in the XRD pattern of TiO2. It has been reported that the formation of brookite phase preferentially occurs in alkaline conditions. The usage of DMF as an alkalescent organic solvent probably induces the formation of brookite phase [19,20,21]. No diffraction peaks from copper oxides such as Cu2O or CuO were observed in the Cu-TiO2, despite the introduction of Cu. In contrast, Cu-TiO2-H showed a diffraction peak from Cu. It is notable that no shift in the diffraction peak position of the anatase phase was observed in either Cu-TiO2 or Cu-TiO2-H, although the ionic radii of the 6-coordinated Ti4+ ions (0.605 Å) and Cu2+ ions (0.73 Å) are different and a Cu2+ ion is larger than that of a Ti4+ ion [22], which suggests that most of the Cu species are precipitated over the TiO2 surface as CuOx clusters on Cu-TiO2, where the clusters are too small to be detected by XRD, and as Cu nanoparticles (NPs) on Cu-TiO2-H, rather than being enclosed in the TiO2 lattice. To obtain detailed structural parameters for the catalysts, we conducted Rietveld profile fitting of these XRD patterns (Figure S1). Structural parameters are summarized in Table S1. The lattice constants of anatase phase constituting Cu-TiO2 and Cu-TiO2-H showed a slight increase in the a-axis but a slight decrease in the c-axis compared to those of pure TiO2, suggesting the possibility of incorporation of Cu species into the TiO2 lattice. The weight fraction of Cu species deposited on TiO2 was estimated by the Rietveld analysis to be low (1.9%), even though the initial starting amount was 10%. Considering the slight change in the structure of Cu-TiO2 and Cu-TiO2-H, a relatively large percentage of Cu species (more than 5%) possibly exists as amorphous on the surface of TiO2.
To confirm the oxidation state of Cu on Cu/TiO2, the diffuse reflectance UV-Vis spectra of the catalyst were measured (Figure 2). The inset represents the Tauc plots [23,24] where the (F(R) hv)1/2 is plotted as a function of photon energy. The indirect band gap was estimated from an X-intercept of a linear fit of the Tauc plot, and the determined band gap values are also given in the inset. The band gap energy of TiO2 was estimated to be 3.10 eV, which is slightly small compared to the reported value of anatase TiO2 [25]. Cu-TiO2 showed a large red-shift in optical absorption edge compared to that of TiO2, with an estimated band gap of 1.46 eV, which occurs probably by interaction between CuO and TiO2. Furthermore, absorption bands appeared in the region of 400–600 nm and 600–1200 nm. The former absorption band is assignable to the interfacial charge transfer from a TiO2 O 2p valence band to a Cu(II) ions connected to TiO2 [26], whereas the latter comes from a d–d transition of Cu (II) species [27]. These Cu(II) states may exist either as Cu(II) clusters or amorphous oxide grains of CuO. Cu-TiO2-H exhibited a remarkably larger red-shift of the absorption edge in optical absorption than Cu-TiO2 and with 1.16 eV of band gap, revealing good contact between Cu and TiO2 grains. Additional absorption bands appeared in the region of 400–600 nm and 550–1200 nm. The former absorption band was assigned to the interfacial charge transfer from the TiO2 O 2p valence band to the Cu(II) ions attached to TiO2 [26], whereas the latter band was attributable to the Cu surface plasmon resonance of Cu NPs [28,29,30]. Therefore, we confirmed the formation of Cu NPs in these catalysts from the XRD and DRS results.
To investigate the dispersion of Cu NPs on TiO2, we conducted scanning transmission electron micrography (STEM) for Cu-TiO2-H. A high-angle annular dark-field STEM (HAADF-STEM) image of the Cu-TiO2-H (Figure 3) suggested that Cu NPs with a diameter of 2–3 nm, which appear as white dot-like objects, were well dispersed over a TiO2 grain, indicating that isolated Cu clusters are formed in Cu-TiO2-H. On the other hand, Cu-TiO2 (Figure S2) suggested that Cu NPs with a diameter of 2–5 nm, which appear as white aggregated objects, existed in a TiO2 grain, indicating that larger Cu NPs are formed in Cu-TiO2 compared to Cu-TiO2-H.
Figure 4 provides Cu 2p3/2 XPS spectra of Cu-TiO2 and Cu-TiO2-H. A relatively broad Cu 2p3/2 XPS peak centered at 932.5 eV and the shoulder peak near 934.3 eV in the high binding energy side were observed on Cu-TiO2. The main and the shoulder peaks could be attributed to Copper oxide containing Cu(I) and Cu(II), and Cu(OH)2 [31] species, respectively, which implies that Cu clusters in Cu-TiO2 are composed of Cu(I) and Cu(II). Furthermore, the presence of a well known shake-up satellite at higher binding energies than that for the main peak strongly indicates the presence of Cu(II) species in Cu-TiO2 [32]. Therefore, Cu clusters in Cu-TiO2 may exist mainly as CuO-like and Cu(OH)2-like species. Cu-TiO2-H showed the sharp Cu 2p3/2 peak centered at 932.6 eV, which is attributed to the formation of Cu(0). The absence of the shake-up satellite, as observed in Cu-TiO2, suggests that Cu NPs on Cu-TiO2-H are mainly composed of Cu(0) species, which is consistent with the observation in the XRD measurement for Cu-TiO2-H. Figure S3A shows Ti 2p XPS spectra of TiO2, Cu-TiO2, and Cu-TiO2-H. All catalysts exhibited a symmetrical peak centered around 458.6 eV, which is a typical peak with characteristic binding energy value for Ti4+ ions contained in anatase TiO2 [33], although the spectrum of TiO2 had a slightly extended tail which possibly comes from the formation of Ti3+. There was also no obvious peak shift in these catalysts. Therefore, Ti ions near the surface of all catalysts would have an analogous chemical environment. Figure S3B represents O1s peaks at around 529.7 eV in XPS spectra of TiO2, Cu-TiO2, and Cu-TiO2-H, which are assigned to lattice oxygen of TiO2 [33]. In addition, each spectrum contained two other weak shoulder peaks on the higher binding energy side of the main O 1s peak. The former peak observed at 530.8 eV can be attributed to the hydroxide group or water molecules that are present at the surface [34], and the other peak at 531.9 eV is originated from organic contaminants containing oxygen species [35]. Cu-TiO2-H showed a relatively large peak at 530.8 eV, which is attributed to the existence of hydroxide groups or water molecules. Chalastara et al. reported that a brookite-rich sample has a larger amount of surface-bound OH/H2O groups than anatase-rich samples [36]. The increase of intensity at 530.8 eV indicates the increase of a brookite phase in Cu-TiO2-H. The number of defects such as Ti3+ and oxygen vacancy on the surface of these catalysts did not change so much by the introduction of Cu or by the change of the atmosphere during heating. The all-measured core level positions for the samples are summarized in Table S2.
The difference in Cu state between Cu-TiO2 and Cu-TiO2-H was further examined by the measurement of Cu-K edge X-ray absorption near edge structure (XANES) spectra using a conversion electron yield (CEY) method, which reflects information on the species near surface region due to the shallow escape depth of Auger electrons from an atom irradiated by X-rays within a sample [37]. Figure 5 shows XANES spectra of Cu-TiO2 and Cu-TiO2-H, and standard samples such as Cu foil, Cu2O, and CuO. Small pre-edge bumps at 8978 eV were observed on Cu-TiO2 and CuO, which is attributed to dipole-forbidden 1s → 3d transition and is indicative of existence of Cu2+ ions, implying that the surface of Cu-TiO2 contains Cu(II) ions. Cu-TiO2-H showed an XANES spectrum similar to that of Cu foil, suggesting that the Cu species on the surface of Cu-TiO2-H are Cu(0) species. XANES spectra were further analyzed in detail. The first derivatives of the XANES spectrum of CuO gave a small pre-edge bump at 8978 eV for the 1s → 3d transition [38], peak for the 1s → 4p transition [39,40] (Figure 5). The spectrum also showed a small pre-edge bump at 8978 eV attributed to the 1s → 3d transition, but the spectral features in the region related to the 1s → 4p transition were significantly different from those of CuO, meaning that the oxidation state of Cu on Cu-TiO2 is different from that of CuO, although it includes Cu(II) species. The red shift and increase in intensity of the 1s → 4p peak was interpreted as an increase of covalency in the ligand−copper bond [40]. Cu(II) states on Cu-TiO2 may emerge as Cu(OH)2-like species, which is consistent with the observation in the XPS measurement for Cu-TiO2. The first derivative spectra of Cu-TiO2-H and Cu foil look similar, indicating that Cu(0) is the main oxidation state of Cu species in Cu-TiO2-H, which is consistent with all observations discussed above.
Figure 6A shows linear sweep voltammetry (LSV) curves under CO2 flow for TiO2, Cu-TiO2, and Cu-TiO2-H. Cu-TiO2-H exhibited a higher current density than the other two catalysts, meaning that the formation of Cu NPs on TiO2 increases ECO2R activity. All catalysts produced H2, CO, CH4, and C2H4 as gaseous products, which were detected by online gas chromatography. We could not observe other gaseous product such as HCOOH. Only HCOO was detected in the liquid phase by HPLC analysis, as shown in Figures S4–S6. TiO2 mainly produced H2 via the hydrogen evolution reaction (HER), and a small amount of HCOO, CO, and CH4 were also produced, suggesting that TiO2 does not show high activity for ECO2R. On Cu-TiO2, the percentage of CH4 in the products increased and a tiny amount of C2H4 was also produced. This suggests that Cu(II) species are reduced to Cu(0) in Cu NPs under the potential and enhances ECO2R activity. The measured total Faradaic efficiencies on the TiO2 and Cu-TiO2 sometimes became slightly higher than 100%, which has been explained by the experimental errors introduced by GC detection or inconsistencies in the flow rate, as shown in some reports [41,42,43]. Cu-TiO2-H exhibited higher ECO2R activity than TiO2 and Cu-TiO2, as shown in the LSV results. Faradaic efficiency for the production of CH4 reached 18% with 36 mA cm−2 of partial current density at −1.8 V vs. RHE (Figure 6F), where CH4 partial current density was defined as a product of the average total current density and the Faradaic efficiency for the production of CH4 in ECO2R at each potential. Notably, Faradaic efficiency for CH4 production much increased at potentials more negative than −1.4 V, indicating that CO2 is selectively converted to CH4 with the applied potentials. To evaluate the selectivity for CH4 formation, we further calculated the ratios of FE for CH4 to total FE for all products in ECO2R (FECH4/FEC1+C2) on the catalysts at each potential (Figure 6E). Cu-TiO2 and Cu-TiO2-H showed larger FECH4/FEC1+C2 values than TiO2. This result is probably attributable to Cu sites homogeneously dispersed on these catalysts. We achieved 70% of FECH4/FEC1+C2 at −1.8 V vs. RHE, which compares with prior reports related to highly dispersed or single-site copper catalysts (Table S3). On the other hand, Cu-TiO2 and Cu-TiO2-H showed different CH4 partial current densities (Figure 6F (b) and (c)), although the selectivity for CH4 in the products from CO2, i.e., FECH4/FEC1+C2 over these catalysts looked similar (Figure 6E (b) and (c)). DRS results represented that the band gap of Cu-TiO2-H is narrower than that of Cu-TiO2 (Figure 2), which implies that Cu species on Cu-TiO2-H have better contact with the TiO2 support than those on Cu-TiO2. Such favorable interactions between Cu catalysts and the TiO2 support in Cu-TiO2-H may enhance the electrical conductivity, resulting in its high CH4 partial current density on Cu-TiO2-H. These results indicate that Cu(0) loading on TiO2 is indispensable and effective to enhance activity in the ECO2R to CH4. However, the state of catalysts under operating conditions are not well clarified and need to be further studied.
It has been reported that composite materials of Cu and CeO2 exhibit high CH4 selectivity [4,7,44,45], where abundant oxygen vacancy (VO) sites on CeO2 play an important role in both CO2 adsorption and activation. Wang et al. suggested that Cu site surrounded by 3VO effectively promoted CH4 formation, where the 2VO neighboring Cu are filled with the two oxygen atoms of CO2 and should enhance the CO2 adsorption and activation [4]. In this study, the surface of the prepared TiO2 samples may consist of low concentration of VO. CO2 dissociation unfavorably occur on stoichiometric anatase surface compared to Vo-rich surface due to its large CO2 dissociation energy [46]. HER mainly proceeded on the TiO2 catalyst, although small amounts of C1 products, such as CO, HCOO, and CH4, were generated (Figure 6B). In contrast, Cu-TiO2 and Cu-TiO2-H, where Cu(0) species were well dispersed over a TiO2 grain, showed high selectively for the production of CH4. Hence, well dispersed Cu species on TiO2 would play a key role in the formation of CH4. Therefore, it is reasonable that homogeneity of Cu NPs in their size and dispersity on TiO2 can be a key for the high CO2-to-CH4 selectivity.

3. Materials and Methods

3.1. Materials

Titanium tetrabutoxide monomer (95.0%), anhydrous copper (II) acetate (97%), N,N-dimethylformamide (99.5%), 2-propanol (99.7%), ethanol (99.5%), acetone (95.5%), hexane (96.0%), and potassium hydroxide were purchased from FUJIFILM Wako Pure Chemical Corporation. Nafion® perfluorinated resin solution (5 wt% in mixture of lower aliphatic alcohols with a water content of 45%) was purchased from Sigma-Aldrich Corporation. All chemicals were used without further purification. All the solutions were prepared with deionized water.

3.2. Preparation of Cu-TiO2 Composite Catalysts

For the preparation of Cu-TiO2 composite catalyst, including 10 wt% of Cu species, titanium tetrabutoxide (2.8 mmol, 1 mL) was quickly added to a mixture of 30 mL of N,N-dimethylformamide (99.5%), 0.215 mL of 2-propanol, and 73 mg of anhydrous copper(II) acetate in a 50 mL Teflon-lined stainless-steel autoclave to avoid exposing the sample to the air, and the mixture was sonicated for 2 h at room temperature. The container was sealed, heated from room temperature to 200 °C over 30 min in an electrical oven, and maintained for 20 h. The product was separated by centrifugation 7500 rpm for 10 min and washed several times with ethanol, acetone, and hexane. After that, it was dried under vacuum at room temperature. Finally, the precursor was calcined at 450 °C for 30 min under air or flowing H2 (60 mL min−1) to obtain TiO2 and Cu-TiO2-x samples, where X is the calcination atmosphere, air or H2; heating rate of 15 °C min−1.

3.3. Catalyst Characterization

The metal composition of prepared oxides was determined using an energy dispersive X-ray spectroscope (EDS, JED-2300, JEOL) equipped with the SEM instrument. Powder X-ray diffraction (XRD) patterns were obtained using synchrotron radiation (λ = 0.740040 Å) at RIKEN Materials Science beamline BL44B2, SPring-8 [47]. Data were acquired using the high-resolution Debye–Scherrer camera equipped with an imaging plate as an X-ray detector. Rietveld analyses were performed using a Topas software package (Bruker AXS Inc., Billerica, MA, USA, version 5). The diffuse reflectance spectra of samples were recorded using a V-670 spectrometer (JASCO, Japan) equipped with an integrating sphere. The diffuse reflection spectra were converted into reflectance spectra using the Kubelka-Munk function. X-ray photoelectron spectroscopy (XPS) studies were performed on a VersaProbeII (ULVAC-PHI) using nonmonochromatic Al Ka radiation. Binding energies in XPS spectra were corrected by referring a C 1s binding energy of the carbon atoms of the ligand in the specimens at 284.6 eV. Scanning transmission electron microscopy (STEM) image was obtained using a JEM-ARM200F (JEOL Co., Tokyo, Japan) at Kyushu University operated at 200 kV. Sample grids for the STEM observations were prepared by dropping ethanol dispersions of the specimens onto a carbon-supported nickel grid. Cu K-edge X-ray absorption fine structure (XAFS spectra of the samples) was measured at Kyushu University beamline BL06 of Kyushu Synchrotron Light Research Center (SAGA-LS, Japan) with an electron storage ring operating at the energy of 1.4 GeV. The energy range of this light source (bending magnet) is 2.1–23 keV. A silicon (111) double-crystal monochromator was used to obtain the incident X-ray beam. The typical photon flux is 1010 photons per second. The spectra of standard samples such as Cu foil, Cu2O, and CuO were recorded in the transmission mode at 20 °C using a Si(111) double-crystal monochromator. The spectra of Cu-TiO2 samples were measured using the conversion electron yield mode. Data processing was carried out by Athena and Artemis included in the Ifeffit package [48].

3.4. Electrode Preparation

The cathode GDE was prepared by airbrushing catalyst inks onto a gas diffusion carbon paper (Fuel Cell Store Sigracet 22 BB, with a microporous layer) with a carrier gas of air. The catalyst ink was prepared with 200 µL of 2-propanol, 200 µL of water, 10 μL of Nafion® solution, and 1 mg of catalyst powder. The catalyst ink mixtures were sonicated in a 4 mL screw neck glass vial for 15 min, and then sprayed onto the gas diffusion carbon paper.

3.5. Electrochemical Reduction of CO2

The electrochemical measurements were conducted in an electrochemical flow cell setup configuration with the three-electrode system. The geometric area of the cathode in the flow cell is 1 cm2, which is used for all current density calculations. 1 M KOH aqueous solution was introduced into the cathode chamber at the rate of 7 mL min−1 and the anode chamber at the rate of 1 mL min−1 by two pumps, respectively. A Nafion 117 cation exchange membrane (Chemours ®) was used to separate the cathode chamber and anode chamber. Pure CO2 gas (Linde, 99.99%) was continuously supplied to the gas chamber of the flow cell at a flow rate of 15 mL min−1. The CO2RR performance was tested using constant-current electrolysis, i.e., chronopotentiometry while purging CO2 into the catholyte during the whole electrochemical test. The potentials vs. Hg/HgO reference electrode were converted to values vs. reversible hydrogen electrode (RHE) using the following equation [49].
E (vs. RHE) = E (vs. Hg/HgO) + 0.098 V + 0.0591 V × pH
All voltages reported are without iR compensation.
Gas products were analyzed by on-line gas chromatography (Micro GC Fusion®, Inficon, Bad Ragaz, Switzerland) with a Molsieve 5A column and a Plot Q column coupled with thermal conductivity detector (TCD). Liquid products were analyzed high performance liquid chromatograph (HPLC, LC-20AD, Shimadzu) equipped with a refractive-index detector (RID-10A, Shimadzu). The Faradaic efficiency (FE) of products in the electroreduction experiments is defined by the following equation:
F E i = n i × z i × F Q × 100
where n i is the number of moles of product i, and z i represents the number of electrons required for the formation of product i ( z i = 2 for CO, formic acid, and H2; z i = 8 for CH4; z i = 12 for C2H4; z i = 14 for C2H6). F is the Faraday constant (96,485 C mol−1 of electrons). Q is the amount of charge passed during the electrolysis. For the gas products, n i was calcurated as follows:
n i , gas = P 0 × x i × v × t R × T
where x i is the volume fraction of gas product i; P 0 is atmospheric pressure (1 atm); v is the CO2 flow rate (0.015 L min−1); t is electrolysis time; R is the ideal gas constant (0.08205 L atm mol−1 K−1); T is 298 K.

4. Conclusions

We successfully prepared Cu-TiO2 composite catalysts, where Cu clusters or NPs were well dispersed by a one-pot solvothermal method and subsequent thermal treatment for electrochemical reduction for CO2. For Cu-TiO2 sample obtained by calcination of the precursor in air, CuOx cluster were dispersed on TiO2 surface and Cu NPs were formed on Cu-TiO2-H obtained by hydrogen treatment of the precursor. Cu-TiO2-H was found to exhibit high selectivity for CH4 in ECO2R. Faradaic efficiency for the CH4 production reached 18% with a CH4 partial current density of 36 mA cm−2 at −1.8 V vs. RHE. Furthermore, 70% of FECH4/FEC1+C2 at −1.8 V vs. RHE. was achieved. We conclude that homogeneity of the Cu NPs formed on TiO2 is one of the necessary factors to maximize CH4 selectivity in the ECO2R.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal12050478/s1, Figure S1: Rietveld analysis results for XRD pattern for (a) TiO2, (b) Cu-TiO2, (c) Cu-TiO2-H. The observed diffraction intensities, calculated patterns, and the difference between the observed and calculated intensity are denoted by red plus signs, a green solid line, and a blue solid line, respectively; Table S1: Structural parameters determined by Rietveld profile fitting for an XRD pattern of TiO2, Cu-TiO2 and Cu-TiO2-H; Figure S2: High-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) image and EDS mapping images of Cu-TiO2; Figure S3: Deconvoluted Ti 2p (A) and O 1s (B) XPS spectra of (a) TiO2, (b) Cu-TiO2, and (c) Cu-TiO2-H. The observed spectra, fitting curves and calculated patterns, and deconvoluted curves are denoted by circle, solid line, and dashed line, respectively; Table S2: XPS peak positions and phase assignment of the TiO2, Cu-TiO2, and Cu-TiO2-H samples; Figure S4: Example of a gas chromatogram (below: enlarged view in the region for CH4 and CO) of H2, CH4, and CO obtained by electrochemical reduction of CO2 using Cu-TiO2-H catalyst on a Molsieve 5A column channel after chronoamperometry operation of 10 min at 1.8 V vs. RHE; Figure S5: Example of a gas chromatogram (below: enlarged view in the region for C2H4) of C2H4 obtained by electrochemical reduction of CO2 using Cu-TiO2-H catalyst on a Plot Q column channel after chronoamperometry operation of 10 min at 1.8 V vs. RHE; Figure S6: Example of a High Performance Liquid Chromatography (HPLC) of liquid products obtained by electrochemical reduction of CO2 using Cu-TiO2-H catalyst after chronoamperometry operation of 10 min at 1.8 V vs. RHE; Table S3: Electrochemical CO2-to-CH4 performance for studies related to highly dispersed or single-site copper catalysts [4,50,51,52,53].

Author Contributions

Methodology, validation, formal analysis, investigation, data curation, writing—original draft preparation, writing—review and editing, A.A.; Validation, investigation, data curation, M.-H.L.; Methodology, validation, formal analysis, investigation, data curation, K.U.; Validation, investigation, data curation, T.G.N.; Validation, investigation, data curation, A.Y.; Methodology, validation, formal analysis, investigation, data curation, K.K.; Methodology, validation, formal analysis, investigation, data curation, T.S.; Conceptualization, Methodology, resources, writing—review and editing, supervision, project administration, funding acquisition, M.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by JSPS KAKENHI (JP12852953 and JP18H05517), JST-CREST (15656567), and a project, “Moonshot Research and Development Program” (JPNP18016), commissioned by the New Energy and Industrial Technology Development Organization (NEDO).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are contained within the article or Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kibria, M.G.; Edwards, J.P.; Gabardo, C.M.; Dinh, C.T.; Seifitokaldani, A.; Sinton, D.; Sargent, E.H. Electrochemical CO2 Reduction into Chemical Feedstocks: From Mechanistic Electrocatalysis Models to System Design. Adv. Mater. 2019, 31, e1807166. [Google Scholar] [CrossRef] [PubMed]
  2. Nitopi, S.; Bertheussen, E.; Scott, S.B.; Liu, X.; Engstfeld, A.K.; Horch, S.; Seger, B.; Stephens, I.E.L.; Chan, K.; Hahn, C.; et al. Progress and Perspectives of Electrochemical CO2 Reduction on Copper in Aqueous Electrolyte. Chem. Rev. 2019, 119, 7610–7672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Peterson, A.A.; Abild-Pedersen, F.; Studt, F.; Rossmeisl, J.; Nørskov, J.K. How copper catalyzes the electroreduction of carbon dioxide into hydrocarbon fuels. Energy Environ. Sci. 2010, 3, 1311–1315. [Google Scholar] [CrossRef]
  4. Wang, Y.F.; Chen, Z.; Han, P.; Du, Y.H.; Gu, Z.X.; Xu, X.; Zheng, G.F. Single-Atomic Cu with Multiple Oxygen Vacancies on Ceria for Electrocatalytic CO2 Reduction to CH4. ACS Catal. 2018, 8, 7113–7119. [Google Scholar] [CrossRef]
  5. Xu, Y.; Li, F.W.; Xu, A.N.; Edwards, J.P.; Hung, S.F.; Gabardo, C.M.; O’Brien, C.P.; Liu, S.J.; Wang, X.; Li, Y.H.; et al. Low coordination number copper catalysts for electrochemical CO2 methanation in a membrane electrode assembly. Nat. Commun. 2021, 12, 4–10. [Google Scholar] [CrossRef] [PubMed]
  6. Varela, A.S.; Ju, W.; Strasser, P. Molecular Nitrogen–Carbon Catalysts, Solid Metal Organic Framework Catalysts, and Solid Metal/Nitrogen-Doped Carbon (MNC) Catalysts for the Electrochemical CO2 Reduction. Adv. Energy Mater. 2018, 8, 1703614. [Google Scholar] [CrossRef] [Green Version]
  7. Yin, J.L.; Gao, Z.Y.; Wei, F.Y.; Liu, C.; Gong, J.; Li, J.M.; Li, W.Z.; Li, X.; Wang, G.W.; Lu, J.T.; et al. Customizable CO2 Electroreduction to C1 or C2+ Products through Cuy/CeO2 Interface Engineering. ACS Catal. 2022, 12, 1004–1011. [Google Scholar] [CrossRef]
  8. Gao, D.F.; Zhang, Y.; Zhou, Z.W.; Cai, F.; Zhao, X.F.; Huang, W.; Li, Y.S.; Zhu, J.F.; Liu, P.; Yang, F.; et al. Enhancing CO2 Electroreduction with the Metal–Oxide Interface. J. Am. Chem. Soc. 2017, 139, 5652–5655. [Google Scholar] [CrossRef]
  9. Nakata, K.; Fujishima, A. TiO2 photocatalysis: Design and applications. J. Photochem. Photobiol. C Photochem. Rev. 2012, 13, 169–189. [Google Scholar] [CrossRef]
  10. Watanabe, R.; Yamauchi, M.; Sadakiyo, M.; Abe, R.; Takeguchi, T. CO2-free electric power circulation via direct charge and discharge using the glycolic acid/oxalic acid redox couple. Energy Environ. Sci. 2015, 8, 1456–1462. [Google Scholar] [CrossRef]
  11. Sadakiyo, M.; Hata, S.; Cui, X.; Yamauchi, M. Electrochemical Production of Glycolic Acid from Oxalic Acid Using a Polymer Electrolyte Alcohol Electrosynthesis Cell Containing a Porous TiO2 Catalyst. Sci. Rep. 2017, 7, 17032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Sadakiyo, M.; Hata, S.; Fukushima, T.; Juhász, G.; Yamauchi, M. Electrochemical hydrogenation of non-aromatic carboxylic acid derivatives as a sustainable synthesis process: From catalyst design to device construction. Phys. Chem. Chem. Phys. 2019, 21, 5882–5889. [Google Scholar] [CrossRef] [PubMed]
  13. Yamauchi, M.; Hata, S.; Eguchi, H.; Kitano, S.; Fukushima, T.; Higashi, M.; Sadakiyo, M.; Kato, K. Catalytic enhancement on Ti-Zr complex oxide particles for electrochemical hydrogenation of oxalic acid to produce an alcoholic compound by controlling electronic states and oxide structures. Catal. Sci. Technol. 2019, 9, 6561. [Google Scholar] [CrossRef]
  14. Fukushima, T.; Yamauchi, M. Electrosynthesis of amino acids from biomass-derivable acids on titanium dioxide. Chem. Commun. 2019, 55, 14721–14724. [Google Scholar] [CrossRef] [PubMed]
  15. Yamauchi, M. Inorganic Nanocatalysts for Hydrogenation Reactions Contributable to a Sustainable Material Supply. Chem. Lett. 2021, 50, 1901–1908. [Google Scholar] [CrossRef]
  16. Eguchi, H.; Kato, K.; Juhasz, G.; Yamauchi, M. Selectivity enhancement in the electrochemical reduction of oxalic acid over titanium dioxide nanoparticles achieved by shape and energy-state control. Catal. Sci. Technol. 2021, 11, 7592–7597. [Google Scholar] [CrossRef]
  17. Isegawa, M.; Staykov, A.; Yamauchi, M. Proton-Coupled Electron Transfer in Electrochemical Alanine Formation from Pyruvic Acid: Mechanism of Catalytic Reaction at the Interface between TiO2(101) and Water. J. Phys. Chem. C 2021, 125, 12603–12613. [Google Scholar] [CrossRef]
  18. Fukutani, K.; Yoshinobu, J.; Yamauchi, M.; Shima, T.; Orimo, S. Hydrogenomics: Efficient and Selective Hydrogenation of Stable Molecules Utilizing Three Aspects of Hydrogen. Catal. Lett. 2021. [Google Scholar] [CrossRef]
  19. Paola, A.D.; Bellardita, M.; Palmisano, L. Brookite, the Least Known TiO2 Photocatalyst. Catalysts 2013, 3, 36–73. [Google Scholar] [CrossRef] [Green Version]
  20. Zhao, B.; Chen, F.; Jiao, Y.; Zhang, J. Phase transition and morphological evolution of titania/titanate nanomaterials under alkalescent hydrothermal treatment. J. Mater. Chem. 2010, 20, 7990–7997. [Google Scholar] [CrossRef]
  21. Haider, Z.; Kang, Y.S. Facile Preparation of Hierarchical TiO2 Nano Structures: Growth Mechanism and Enhanced Photocatalytic H2 Production from Water Splitting Using Methanol as a Sacrificial Reagent. ACS Appl. Mater. Interfaces 2014, 6, 10342–10352. [Google Scholar] [CrossRef] [PubMed]
  22. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Cryst. 1976, A32, 751–767. [Google Scholar] [CrossRef]
  23. Tauc, J.; Grigorovici, R.; Vancu, A. Optical Properties and Electronic Structure of Amorphous Germanium. Phys. Status Solidi 1966, 15, 627–637. [Google Scholar] [CrossRef]
  24. Davis, E.A.; Mott, N.F. Conduction in non-crystalline systems V. Conductivity, optical absorption and photoconductivity in amorphous semiconductors. Philos. Mag. 1970, 22, 903–922. [Google Scholar] [CrossRef]
  25. Reyes-Coronado, D.; Rodríguez-Gattorno, G.; Espinosa-Pesqueira, M.E.; Cab, C.; De Coss, R.; Oskam, G. Phase-pure TiO2 nanoparticles: Anatase, brookite and rutile. Nanotechnology 2008, 19, 145605. [Google Scholar] [CrossRef]
  26. Choudhury, B.; Dey, M.; Choudhury, A. Defect generation, d-d transition, and band gap reduction in Cu-doped TiO2 nanoparticles. Int. Nano Lett. 2013, 3, 25. [Google Scholar] [CrossRef] [Green Version]
  27. Pascher, T.F.; Ončák, M.; van der Linde, C.; Beyer, M.K. UV/Vis Spectroscopy of Copper Formate Clusters: Insight into Metal-Ligand Photochemistry. Chem.-A Eur. J. 2020, 26, 8286–8295. [Google Scholar] [CrossRef]
  28. Desario, P.A.; Pietron, J.J.; Brintlinger, T.H.; McEntee, M.; Parker, J.F.; Baturina, O.; Stroud, R.M.; Rolison, D.R. Oxidation-stable plasmonic copper nanoparticles in photocatalytic TiO2 nanoarchitectures. Nanoscale 2017, 9, 11720–11729. [Google Scholar] [CrossRef] [Green Version]
  29. Tsuji, M.; Hikino, S.; Tanabe, R.; Sano, Y. Synthesis of Bicompartmental Ag/Cu Nanoparticles Using a Two-step Polyol Process. Chem. Lett. 2009, 38, 860–861. [Google Scholar] [CrossRef]
  30. Amendola, V.; Meneghetti, M. Laser ablation synthesis in solution and size manipulation of noble metal nanoparticles. Phys. Chem. Chem. Phys. 2009, 11, 3805–3821. [Google Scholar] [CrossRef]
  31. Cano, E.; Torres, C.L.; Bastidas, J.M. An XPS study of copper corrosion originated by formic acid vapour at 40% and 80% relative humidity. Mater. Corros. 2001, 52, 667–676. [Google Scholar] [CrossRef]
  32. Ethiraj, A.S.; Kang, D.J. Synthesis and characterization of CuO nanowires by a simple wet chemical method. Nanoscale Res. Lett. 2012, 7, 70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Biesinger, M.C.; Lau, L.W.M.; Gerson, A.R.; Smart, R.S.C. Resolving surface chemical states in XPS analysis of first row transition metals, oxides and hydroxides: Sc, Ti, V, Cu and Zn. Appl. Surf. Sci. 2010, 257, 887–898. [Google Scholar] [CrossRef]
  34. Idriss, H. On the wrong assignment of the XPS O1s signal at 531–532 eV attributed to oxygen vacancies in photo- and electro-catalysts for water splitting and other materials applications. Surf. Sci. 2021, 712, 121894. [Google Scholar] [CrossRef]
  35. Kenausis, G.L.; Vörös, J.; Elbert, D.L.; Huang, N.; Hofer, R.; Ruiz-Taylor, L.; Textor, M.; Hubbell, J.A.; Spencer, N.D. Poly(L-lysine)-g-poly(ethylene glycol) layers on metal oxide surfaces: Attachment mechanism and effects of polymer architecture on resistance to protein adsorption. J. Phys. Chem. B 2000, 104, 3298–3309. [Google Scholar] [CrossRef]
  36. Chalastara, K.; Guo, F.; Elouatik, S.; Demopoulos, G.P. Tunable Composition Aqueous-Synthesized. Catalyst 2020, 10, 407. [Google Scholar] [CrossRef] [Green Version]
  37. Itai, T.; Takahashi, Y.; Uruga, T.; Tanida, H.; Iida, A. Selective detection of Fe and Mn species at mineral surfaces in weathered granite by conversion electron yield X-ray absorption fine structure. Appl. Geochem. 2008, 23, 2667–2675. [Google Scholar] [CrossRef]
  38. DuBois, J.L.; Mukherjee, P.; Stack, T.D.P.; Hedman, B.; Solomon, E.I.; Hodgson, K.O. A systematic K-edge X-ray absorption spectroscopic study of Cu(III) sites. J. Am. Chem. Soc. 2000, 122, 5775–5787. [Google Scholar] [CrossRef]
  39. Todaro, M.; Sciortino, L.; Gelardi, F.M.; Buscarino, G. Determination of geometry arrangement of copper ions in HKUST-1 by XAFS during a prolonged exposure to air. J. Phys. Chem. C 2017, 121, 24853–24860. [Google Scholar] [CrossRef]
  40. Prestipino, C.; Regli, L.; Vitillo, J.G.; Bonino, F.; Damin, A.; Lamberti, C.; Zecchina, A.; Solari, P.L.; Kongshaug, K.O.; Bordiga, S. Local structure of framework Cu(II) in HKUST-1 metallorganic framework: Spectroscopic characterization upon activation and interaction with adsorbates. Chem. Mater. 2006, 18, 1337–1346. [Google Scholar] [CrossRef]
  41. Shinagawa, T.; Larrazábal, G.O.; Martín, A.J.; Krumeich, F.; Pérez-Ramírez, J. Sulfur-Modified Copper Catalysts for the Electrochemical Reduction of Carbon Dioxide to Formate. ACS Catal. 2018, 8, 837–844. [Google Scholar] [CrossRef] [Green Version]
  42. Kim, S.; Dong, W.J.; Gim, S.; Sohn, W.; Park, J.Y.; Yoo, C.J.; Jang, J.Y.; Lee, J.-L. Shape-controlled bismuth nanoflakes as highly selective catalysts for electrochemical carbon dioxide reduction to formate. Nano Energy 2017, 39, 44–52. [Google Scholar] [CrossRef]
  43. Zheng, T.T.; Jiang, K.; Ta, N.; Hu, Y.F.; Zeng, J.; Liu, J.Y.; Wang, H.T. Large-Scale and Highly Selective CO2 Electrocatalytic Reduction on Nickel Single-Atom Catalyst. Joule 2019, 3, 265–278. [Google Scholar] [CrossRef] [Green Version]
  44. Xue, L.; Zhang, C.; Wu, J.; Fan, Q.Y.; Liu, Y.; Wu, Y.; Li, J.; Zhang, H.; Liu, F.; Zeng, S. Unveiling the reaction pathway on Cu/CeO2 catalyst for electrocatalytic CO2 reduction to CH4. Appl. Catal. B Environ. 2022, 304, 120951. [Google Scholar] [CrossRef]
  45. Varandili, S.B.; Huang, J.; Oveisi, E.; Gregorio, G.L.D.; Mensi, M.; Strach, M.; Vavra, J.; Gadiyar, C.; Bhowmik, A.; Buonsanti, R. Synthesis of Cu/CeO2−x Nanocrystalline Heterodimers with Interfacial Active Sites To Promote CO2 Electroreduction. ACS Catal. 2019, 9, 5035–5046. [Google Scholar] [CrossRef]
  46. Huygh, S.; Bogaerts, A.; Neyts, E.C. How Oxygen Vacancies Activate CO2 Dissociation on TiO2 Anatase (001). J. Phys. Chem. C 2016, 120, 21659–21669. [Google Scholar] [CrossRef]
  47. Kato, K.; Tanaka, H. Visualizing charge densities and electrostatic potentials in materials by synchrotron X-ray powder diffraction. Adv. Phys. X 2016, 1, 55–80. [Google Scholar] [CrossRef] [Green Version]
  48. Ravel, B.; Newville, M. ATHENA, ARTEMIS, HEPHAESTUS: Data analysis for X-ray absorption spectroscopy using IFEFFIT. J. Synchrotron Radiat. 2005, 12, 537–541. [Google Scholar] [CrossRef] [Green Version]
  49. Niu, S.; Li, S.; Du, Y.; Han, X.; Xu, P. How to Reliably Report the Overpotential of an Electrocatalyst. ACS Energy Lett. 2020, 5, 1083–1087. [Google Scholar] [CrossRef] [Green Version]
  50. Zhang, T.; Verma, S.; Kim, S.; Fister, T.T.; Kenis, P.J.A.; Gewirth, A.A. Highly dispersed, single-site copper catalysts for the electroreduction of CO2 to methane. J. Electroanal. Chem. 2020, 875, 113862. [Google Scholar] [CrossRef]
  51. Guan, A.X.; Chen, Z.; Quan, Y.L.; Peng, C.; Wang, Z.Q.; Sham, T.-K.; Yang, C.; Ji, Y.L.; Qian, L.P.; Xu, X.; et al. Boosting CO2 Electroreduction to CH4 via Tuning Neighboring Single-Copper Sites. ACS Energy Lett. 2020, 5, 1044–1053. [Google Scholar] [CrossRef]
  52. Chen, S.H.; Wang, B.Q.; Zhu, J.X.; Wang, L.Q.; Ou, H.H.; Zhang, Z.D.; Liang, X.; Zheng, L.R.; Zhou, L.; Su, Y.-Q.; et al. Lewis Acid Site-Promoted Single-Atomic Cu Catalyzes Electrochemical CO2 Methanation. Nano Lett. 2021, 21, 7325–7331. [Google Scholar] [CrossRef] [PubMed]
  53. Guan, A.X.; Yang, C.; Wang, Q.H.; Qian, L.P.; Cao, J.Y.; Zhang, L.J.; Wu, M.L.; Zheng, G.F.; Zhou, L.; Su, Y.-Q.; et al. DAtomic-Level Copper Sites for Selective CO2 Electroreduction to Hydrocarbon. ACS Sustain. Chem. Eng. 2021, 9, 13536–13544. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) TiO2, (b) Cu-TiO2, and (c) Cu-TiO2-H. The simulated XRD patterns of anatase, brookite, and Cu are also presented.
Figure 1. XRD patterns of (a) TiO2, (b) Cu-TiO2, and (c) Cu-TiO2-H. The simulated XRD patterns of anatase, brookite, and Cu are also presented.
Catalysts 12 00478 g001
Figure 2. Diffuse reflectance spectra of (a) TiO2, (b) Cu-TiO2, and (c) Cu-TiO2-H samples. The inset is the Tauc plots for the determination of the bandgap of the samples.
Figure 2. Diffuse reflectance spectra of (a) TiO2, (b) Cu-TiO2, and (c) Cu-TiO2-H samples. The inset is the Tauc plots for the determination of the bandgap of the samples.
Catalysts 12 00478 g002
Figure 3. High-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) image and EDS mapping images of Cu-TiO2-H.
Figure 3. High-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) image and EDS mapping images of Cu-TiO2-H.
Catalysts 12 00478 g003
Figure 4. Deconvoluted Cu 2p3/2 XPS spectra of Cu-TiO2 (a) and Cu-TiO2-H (b). The observed spectra, fitting curves, calculated patterns, and deconvoluted curves are denoted by circle, solid line, and dashed line, respectively.
Figure 4. Deconvoluted Cu 2p3/2 XPS spectra of Cu-TiO2 (a) and Cu-TiO2-H (b). The observed spectra, fitting curves, calculated patterns, and deconvoluted curves are denoted by circle, solid line, and dashed line, respectively.
Catalysts 12 00478 g004
Figure 5. Cu K-edge spectra (A) and first derivative of Cu K-edge XANES spectra (B) of reference compounds. (a) Cu-TiO2, (b) Cu-TiO2-H. The inset figure shows overlay of the spectra of CuO and (a) Cu-TiO2.
Figure 5. Cu K-edge spectra (A) and first derivative of Cu K-edge XANES spectra (B) of reference compounds. (a) Cu-TiO2, (b) Cu-TiO2-H. The inset figure shows overlay of the spectra of CuO and (a) Cu-TiO2.
Catalysts 12 00478 g005
Figure 6. LSV curves under CO2 flow (A) on an electrode employing (a) TiO2, (b) Cu-TiO2, and (c) Cu-TiO2-H catalyst. FEs for different products over TiO2 (B), Cu-TiO2 (C), and Cu-TiO2-H (D) catalysts at various potentials. Comparison of FECH4/FEC1+C2 (E) on (a) TiO2, (b) Cu-TiO2, and (c) Cu-TiO2-H catalyst. Comparison of CH4 partial current density (F) of (a) TiO2, (b) Cu-TiO2, and (c) Cu-TiO2-H catalyst.
Figure 6. LSV curves under CO2 flow (A) on an electrode employing (a) TiO2, (b) Cu-TiO2, and (c) Cu-TiO2-H catalyst. FEs for different products over TiO2 (B), Cu-TiO2 (C), and Cu-TiO2-H (D) catalysts at various potentials. Comparison of FECH4/FEC1+C2 (E) on (a) TiO2, (b) Cu-TiO2, and (c) Cu-TiO2-H catalyst. Comparison of CH4 partial current density (F) of (a) TiO2, (b) Cu-TiO2, and (c) Cu-TiO2-H catalyst.
Catalysts 12 00478 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Anzai, A.; Liu, M.-H.; Ura, K.; Noguchi, T.G.; Yoshizawa, A.; Kato, K.; Sugiyama, T.; Yamauchi, M. Cu Modified TiO2 Catalyst for Electrochemical Reduction of Carbon Dioxide to Methane. Catalysts 2022, 12, 478. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12050478

AMA Style

Anzai A, Liu M-H, Ura K, Noguchi TG, Yoshizawa A, Kato K, Sugiyama T, Yamauchi M. Cu Modified TiO2 Catalyst for Electrochemical Reduction of Carbon Dioxide to Methane. Catalysts. 2022; 12(5):478. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12050478

Chicago/Turabian Style

Anzai, Akihiko, Ming-Han Liu, Kenjiro Ura, Tomohiro G. Noguchi, Akina Yoshizawa, Kenichi Kato, Takeharu Sugiyama, and Miho Yamauchi. 2022. "Cu Modified TiO2 Catalyst for Electrochemical Reduction of Carbon Dioxide to Methane" Catalysts 12, no. 5: 478. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12050478

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