Next Article in Journal
A Green Synthesis of 3-Selanyl-Isoflavones via Lipase Mediated Selenylation/Cyclization of Enaminones
Previous Article in Journal
Correction: Aldana et al. Nanocomposite PVDF/TiO2 Photocatalytic Membranes for Micropollutant Removal in Secondary Effluent. Catalysts 2024, 14, 109
Previous Article in Special Issue
The Construction of p/n-Cu2O Heterojunction Catalysts for Efficient CO2 Photoelectric Reduction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CO2 Electroreduction by Engineering the Cu2O/RGO Interphase

by
Matteo Bisetto
1,2,*,
Sourav Rej
3,
Alberto Naldoni
4,
Tiziano Montini
1,5,
Manuela Bevilacqua
5,6,* and
Paolo Fornasiero
1,2,5
1
Department of Chemical and Pharmaceutical Sciences, University of Trieste, Via L. Giorgieri 1, 34127 Trieste, Italy
2
National Interuniversity Consortium of Materials Science and Technology (INSTM), University of Trieste, Via L. Giorgieri 1, 34127 Trieste, Italy
3
Czech Advanced Technology and Research Institute, Regional Centre of Advanced Technologies and Materials (RCPTM), Palacký University Olomouc, Šlechtitelů 27, 77900 Olomouc, Czech Republic
4
Department of Chemistry and NIS Centre, University of Turin, Via P. Giuria 7, 10125 Turin, Italy
5
Institute of Chemistry of Organometallic Compounds (ICCOM), National Research Council (CNR), Trieste Research Unit, University of Trieste, Via L. Giorgieri 1, 34127 Trieste, Italy
6
Institute of Chemistry of Organometallic Compounds (ICCOM), National Research Council (CNR), Via Madonna del Piano 10, 50019 Sesto Fiorentino, Italy
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(7), 412; https://doi.org/10.3390/catal14070412
Submission received: 24 May 2024 / Revised: 20 June 2024 / Accepted: 21 June 2024 / Published: 28 June 2024

Abstract

:
In the present investigation, Cu2O-based composites were successfully prepared through a multistep method where cubic Cu2O nanoparticles (CU Cu2O) have been grown on Reduced Graphene Oxide (RGO) nanosheets. The structural and morphological properties of the materials have been studied through a comprehensive characterization, confirming the coexistence of crystalline Cu2O and RGO. Microscopical imaging revealed the intimate contact between the two materials, affecting the size and the distribution of Cu2O nanoparticles on the support. The features of the improved morphology strongly affected the electrochemical behavior of the composites, increasing the activity and the faradaic efficiencies towards the electrochemical CO2 reduction reaction process. CU Cu2O/RGO 2:1 composite displayed selective CO formation over H2, with higher currents compared to pristine Cu2O (−0.34 mA/cm2 for Cu2O and −0.64 mA/cm2 for CU Cu2O/RGO 2:1 at the voltage of −0.8 vs. RHE and in a CO2 atmosphere) and a faradaic efficiency of 50% at −0.9 V vs. RHE. This composition exhibited significantly higher CO production compared to the pristine materials, indicating a favorable *CO intermediate pathway even at lower voltages. The systematic investigation on the effects of nanostructuration on composition, morphology and catalytic behavior is a valuable solution for the formation of effective interphases for the promotion of catalytic properties providing crucial insights for future catalysts design and applications.

Graphical Abstract

1. Introduction

The significant increase in CO2 greenhouse gas and the care about the security of energy supply have received great interest and are considered as the main challenges of our century. In particular, the conversion of CO2 into useful chemicals, building blocks, energy vectors and/or fuels by photochemistry and electrocatalysis have been considered as one of the most advanced approaches to limit both energy and environmental problems at the same time [1,2,3]. Many promising and low-cost photo/electrocatalysts, non-precious metal-based or metal-free composite materials have been documented in the more recent literature; otherwise the design of adequate candidate composite electrocatalysts in terms of efficiency and selectivity needs to be also related to the critical impact of the adopted synthesis pathway [4,5,6,7,8,9].
In the last three decades, copper oxides, particularly Cu2O, have been considered as important precursors for preparing Cu-based catalysts for electrochemical CO2 reduction reaction (CO2RR), due to the improved electrocatalytic performance compared to untreated metallic Cu [10,11]. In the study presented by D. Ren et al. in 2015, metallic Cu exhibited a Faradaic Efficiency (FE) of 13.8% for ethylene (C2H4) and 0% for ethanol (C2H5OH) at −1.0 V vs. RHE. In contrast, under the same conditions the equivalent oxide-derived Cu film catalyst achieved an optimized FE of ca. 39% and ca. 16% for C2H4 and C2H5OH, respectively, mainly dependent on the specific film thickness [12]. Although copper oxides (CuO and Cu2O) are generally fast reduced to metallic Cu under CO2RR conditions according to the Pourbaix diagram, there are other factors that can be discussed to rationalize the relationship between the activity, the selectivity and stability of oxide-derived Cu catalysts, including morphological modifications, defect creation, local pH changes, etc. [13,14].
Cu-based nanostructures characterized by cubic morphology have showed an increased selectivity toward the desired C2+ reaction products, and are considered a point of reference in the state of the art [15,16,17,18,19]. In fact, it has been proved that specific morphological parameters of cubic NPs, in particular the size and the surface properties, could be responsible for better control of the selectivity and stability [20,21]. On the other hand, it is not completely clear which parameter of the synthesis could be strongly responsible for a particular obtained morphology: the comprehension of this aspect is fundamental to fine-tune the NPs structure and consequently to optimize their electrocatalytic properties [22]. Importantly, the addition of a carbon-based support to well disperse the cubic Cu2O (CU Cu2O) NPs and/or to induce a synergic behavior due to the presence on multiple and different interacting active sites has to be taken into account in order to enhance the interphase role [23,24,25]. Very recently, Ze-lin Wu et al. published a review work focused on the use of graphene-based support material for CO2RR electrocatalysts [26]. In this context, graphene is considered an attractive carbon material due to the particular two-dimensional structure, combined with the high specific surface area, a good chemical and mechanical stability and a facile tunability for anchoring different second phase compounds through physical and chemical coupling interactions [27,28]. For these multiple advantages, graphene has been widely explored as either the catalytically active material or the supporting platform in the construction of advanced CO2RR electrocatalysts. The Reduced Graphene Oxide differs from the precursor Graphene Oxide (GO) in terms of morphology, optical and electrical properties [29,30]. In 2017 E. Jaafar and co-workers highlighted that a folded and wrinkled structure has been observed for RGO with respect to GO [31]. This folding structure can be found on both surface and the edge of RGO due to the losses of oxygen functional groups. Furthermore, they observed that a more folded and wrinkled structure could be produced when the reduction was stressed. By the point of view of electrical behavior, the total current registered for a thin film of RGO deposited on a common glassy carbon support was higher when compared to a similar texture of GO. Once again, this feature was attributed to an oxygen decrease in reduced graphene oxide. In the present work we prepared different composites based on cubic Cu2O/RGO and we have used them as electrocatalysts towards the CO2RR. Different efforts have been spent to characterize the obtained materials, with the aim to examine the intrinsic effects of the coupling between the two counterparts of the heterostructure. We deepened the relationship between the morphological, structural and electric properties of CU Cu2O/RGO composites and the corresponding performance evaluated for CO2RR, to identify the keys for tuning their electrochemical activity and selectivity. Remarkably, catalytic experiments revealed an improved selectivity towards the formation of CO, demonstrating the potential of the proposed material for applications focused on CO2 conversion.

2. Results and Discussion

2.1. Characterization of CU Cu2O/RGO Composites

CU Cu2O/RGO composites were prepared by growing cubic nanoparticles of Cu2O directly on the sheets of RGO (Figure 1). RGO was firstly obtained via a two-step process where graphite was exfoliated and oxidized to form nanosheets of GO through a modified Hummer’s method and later partially reduced by addition of hydrazine on the solution to form RGO nanosheets. Cubic Cu2O nanoparticles were directly grown on the support, using different fractions of RGO. The different amount of RGO affects the stability of the composite and its catalytic behavior, with lower affinity between the two counterparts with a higher fraction of support. For this reason, three different compositions had been selected (CU Cu2O/RGO 2:1, 1:1 and 1:2) in order to understand the effects of the interaction between the two counterparts of the composite towards the overall catalytic behavior.
The crystalline structure of the as-prepared catalyst was revealed by XRD. The synthesis allowed the formation of crystalline Cu2O nanoparticles, with all the typical feature of Cu2O revealed by XRD diffractograms (Figure 2a) [32]. The chemical stability of the nanoparticles was confirmed by the absence of any reflection related to metallic Cu or CuO. Moreover, the Rietveld analysis allowed quantitative information regarding the crystalline structure of Cu2O to be obtained, which showed a lattice parameter of 4.27 Å, in good agreement with previously reported values (Figure S1) [33]. No shifts of the position for Cu2O reflections were detected, suggesting that the introduction of RGO has negligible effects on the crystalline structure of the pristine Cu2O [34]. Raman spectra of the composites (Figure 2b) confirmed the contemporaneous presence of Cu2O and RGO by revealing the characteristic vibrations of both phases [35,36]. From the analysis of the fingerprint, it is possible to observe the typical vibrations of Cu2O and discriminate them from those of CuO, supporting the purity of the chemical phase observed from XRD. The characteristic features of RGO were identified by the typical D-band observed at 1340 cm−1 related to the breathing mode of phonons of A1g symmetry while the G-band at 1592 cm−1 arises from the first-order scattering of E2g phonons by sp2 carbon of GO [37,38]. D-band and G-band are observed also in the pristine RGO (Figure S2) discriminating RGO from GO; indeed, the higher intensity of the D-band peak can be related to the removal of oxygen moieties from GO after reduction and suggested the presence of RGO instead of GO [39,40]. The different intensities between the two families of peaks are indicative of the relative composition for the prepared materials.
TGA measurements performed up to 800 °C in air allowed to understand the thermal stability of the different materials with the variation of the temperature (Figure 2c). The mass of the material decreased in the first part of the thermal treatment due to the desorption of the adsorbed water molecules while the rapid reduction of the mass around 400–500 °C is characteristic of CO and CO2 pyrolysis on graphene-like materials [29,41]. The thermal treatment in air had implications even on the Cu2O material, with a conversion to CuO through the reaction 2Cu2O + O2 → 4CuO [42]. From the analysis of the reported trends, it is possible to calculate the fraction of initial Cu2O on the prepared composites, revealing an experimental fraction of Cu2O similar to the theoretical just in the composite 2:1 and confirming the improved affinity of the two materials in the latter case compared to the other composites (Table S1).
SEM and TEM images revealed that the obtained pristine RGO is characterized by thin and crumpled sheets, with different platelets closely associated with each other (Figure S3). The synthesis of unsupported Cu2O nanoparticles allowed the formation of cubes bounded with six (100) facets and an average dimension of 200 nm (±20 nm) (Figures S3 and S4a). When the synthesis of Cu2O occurred in a dispersion of RGO, the cubic Cu2O nanoparticles were directly grown on the support, thanks to the stabilizing features of RGO sheets (Figure 3). TEM images confirmed the intimate contact between the two materials of the catalyst, with the nanostructures that are fully bounded by the 3D plates of RGO (Figure 3a,b). On the other hand, the formation of the composites altered the dimension of Cu2O nanoparticles, with an average size of 150 nm for Cu2O/RGO 2:1, slightly decrease than the unsupported material (Figure 3c,d). The other compositions showed comparable average sizes, intermediate between the pristine and the supported materials (Figure S4 and Table S2). The formation of the proper composite is demonstrated even through EDX, confirming the presence of characteristics elements such as Cu or O and revealing high percentage of C-species due to the RGO support (Figure 3e).

2.2. Electrochemical Behavior of CU Cu2O/RGO Composites and CO2RR Tests

The electrocatalytic behavior of the different materials was firstly characterized using the typical three-electrode setup, depositing an ink of the specific material on the surface of a Rotating Disc Electrode (RDE, Figures S5 and S6). The presented Linear Sweep Voltammetries (LSVs) and ChronoAmperometries (CAs) showed a different behavior under the Ar and CO2 atmosphere, indicating a dependency of the activity from the atmosphere and an overall influence on the catalytic pathway of the electrocatalysts. The formation of the composites has implications on the overall activity of Cu2O, with higher current of the Cu2O/RGO system due to the improved dispersion of Cu2O cubes and the intimate contact within the sheets of RGO (i.e., −0.34 mA/cm2 for Cu2O and −0.64 mA/cm2 for CU Cu2O/RGO 2:1 at the voltage of −0.8 V vs. RHE in the CO2 atmosphere; this feature can be observed for all the other composites). The reduction peak of Cu2O to metal Cu is observed at −0.2 V vs. RHE, showing a modification of the material due to the application of the negative voltage. Concerning the stability of the catalyst, no decays or rapid changes of the currents were observed during the chronoamperometric tests; otherwise, a general decrease in the current was observed after a few minutes from the application of the voltage followed by a stable response for the 90 min of electrochemical characterization. The decreasing of the current is related to the rapid reduction of the Cu(I) phase to metallic copper under the application of a constant voltage and the subsequent stabilization of the phase during the first minutes of the analysis [43]. Moreover, according to the Electrochemical Impedance Spectroscopy (EIS) data summarized in the Nyquist plot of Figure S7, the total impedance strongly decreases for the composite materials with respect to both the pristine CU Cu2O and the carbon RGO support, accounting for enhanced charge transfer and diffusive contributes; furthermore, in the range of ca. 1–5 kHz frequencies, Z values referred to the coupled electrocatalysts are still lower than the separate phases, suggesting an improved charge transfer that could be mediated by the intimate contact of Cu and RGO (Figure S7b).
An in-deep electrochemical CO2RR characterization had been performed to unveil the role of CO2 on different catalysts behavior, analyzing the performances of the materials in terms of current and productivity. Chronoamperometries at different voltages (−0.7 V, −0.9 V and −1.1 V vs. RHE) were used to evaluate the dependency of the activity from the applied potential. The detailed characterization at the voltage of −0.9 V is presented in Figure 4 while the other voltages are reported in the supporting information (Figures S8 and S9). Compared to the pristine materials, all the composites showed an improved current density especially at lower potentials (i.e., −1.79 mA/cm2 for Cu2O and −2.72 mA/cm2 for CU Cu2O/RGO 2:1 at the voltage of −0.9 V vs. RHE in the CO2 atmosphere), evidencing that the higher exposed area of the materials and the intrinsic affinity between the two counterparts affected the overall electrochemical response in a beneficial way (Figure 4a). The different products from the catalytic process were tracked using gas and liquid chromatography, analyzing the overall quantities and calculating the relative total FEs at the proper potential for the 120 min of characterization.
The improved electrochemical phase boundary of the CU Cu2O/RGO texture increased the current of the materials, and it had implications on the selectivity and productivity, limiting the formation of H2 and shifting the activity towards the CO2RR. Moreover, for the quasi-total reactivity of bare RGO toward the HER process [44], the different composites showed higher faradaic efficiencies for the formation of CO2RR products, in particular CO and HCOO (Figure 4b). Among the different catalytic pathways, CO2 can be reduced forming CO or formic acid, depending on which part of the CO2 molecule interacts with the surface of the catalyst [45]. As it was reported, 100-like Cu domains are able to trigger the formation of C-C bond and selectively induce the production of *CO species [46]. Therefore, the addition of cubic Cu2O has implications on the overall behavior of RGO, leading to the adsorption of CO2 molecules via a *CO2 intermediate and explaining a favorable CO formation compared to the other products [47]. As a result of the different catalytic responses of the composites, the F E H 2 dropped from 73% for the pristine RGO to 18% for the unsupported CU Cu2O, and a contemporaneous improving of the selectivity for the CO formation is observed, with the composite CU Cu2O/RGO 2:1 characterized by a F E C O of 50%, the highest F E C O between all the materials. The material showed also a F E H C O O of 18%, higher than the other prepared composites. Formic acid is considered an interesting hydrogen storage thanks to the high energy density and the fact that, being a liquid product, it can be separated easily from the rest of the other products of catalysis [48]. In this optic, the comparison of F E H C O O for the different materials allows confirmation of the beneficial interaction between the two counterparts of the composite. The obtained selectivity, coupled with the increased current of the composite, provided a higher amount of CO during the experiments, more than double compared with unsupported Cu2O (29 µmol vs. 14 µmol, respectively) and suggested a favored pathway toward the formation of CO (Figure 4c). The high fraction of RGO in the composites 1:1 and 1:2 still has implication on the overall H2 selectivity, with an approximately F E H 2 of 60%, and productivity, affected also by the higher current. The analysis of the ratio between the moles of CO and the moles of all the products quantified during the CA test for these specified compositions showed a decrease in the ratio due to the increased selectivity toward HER and a correlation with the increasing of RGO content (Figure 4d). On the other hand, CU Cu2O/RGO 2:1 showed the highest CO/products ratio and confirmed the improved selectivity toward the formation of CO. The overall behaviors follow similar trends even at the other considered voltages. At −0.7 V vs. RHE, the activity of HER becomes predominant compared to CO2RR, favored by the high concentration of protons in the aqueous-based electrolytes (Figure S8) [49,50]. Cu2O/RGO 2:1 still shows the highest F E C O (38%) compared to all the other materials: the selectivity towards HER is the lowest meaning that the synergistic effects of coupling between the two counterparts of the composite suppress the formation of H2 and improved the activity of CO2RR. This composition shows interesting properties in terms of selectivity and mostly productivity, with a ratio of CO compared to all the material that is almost double (0.39) respect to those obtained for other prepared composites, during the 120 min of experiment. The high faradaic efficiency at a relatively low applied voltage suggested an efficient formation of CO by the Cu2O/RGO 2:1 composite, attesting improved selectivity compared to RGO-based systems reported in the literature (Table S3) [51,52,53,54,55]. At the most reductive voltage (−1.1 V vs. RHE), the high overvoltage relies on an improved selectivity of the different catalysts towards the CO2RR process and higher productivities for ethylene, formic acid and targeted CO. The effects of the voltages on the selectivity for the different fractions between Cu2O and RGO in the composite are highlighted in Figure 5. Cu2O/RGO 1:1 and 1:2 showed a similar catalytic response, with an important fraction of H2 produced and low selectivity towards CO2RR products, even at more reductive voltages. This might be attributed to the intrinsic activity of RGO towards HER and the similar sizes of Cu2O nanoparticles in the two composites, since this parameter is considered crucial for the different behavior of an electrocatalyst. The activity changes when the fraction of Cu2O is higher than RGO, where an improved dispersion of Cu2O nanoparticles is observed. In this case, the lowest dimensions of Cu2O nanoparticles and the synergistic effects due to the coupling with RGO sheets allowed an improved activity of CO2RR reaction, with the highest F E C O at all the considered voltages.

3. Materials and Methods

All reagents were pure and of analytical grade. Graphite, NaNO3, H2SO4, KMnO4 and H2O2 were utilized for the modified Hummer’s method and the preparation of GO while, for the reduction of GO to RGO, NaOH and H2N2·2H2O were used. SDS, CuCl2, NH2OH*HCl and NaOH were used as precursors for the synthesis of Cu2O nanoparticles.

3.1. Preparation of the CU Cu2O/RGO Composites

RGO was prepared via an adapted two-step method where graphite was initially converted to Graphene Oxide (GO) through a modified Hummer’s method and reduced to obtain RGO [56,57]. In a typical procedure, 0.2 g of Graphite and 0.2 g of NaNO3 were mixed in a round bottom flask with 10 mL of H2SO4 (95%). The solution was stirred at 400 rpm and kept in an ice bath for 30 min. Potassium Permanganate (KMnO4, 1.2 g) was gradually added to the solution. After five minutes, 16 mL of Milli-Q H2O was added, and the entire solution was stirred for one hour. GO is formed through addition of 40 mL of Milli-Q H2O and 1.2 mL of H2O2 and stirring for one hour. The obtained powder of GO was washed with a 1:3 HCl solution to remove all the metal traces from the obtained material and with absolute ethanol; GO was then separated by filtration and dried. In the second step, GO was reduced to RGO by addition of a strong reducing agent such as H2N2·2H2O. In a typical procedure, 50 mg of GO was mixed with 50 mL of H2O in a bottom flask. The pH of the mixture was corrected to 10–11 by the addition of 0.1 M NaOH solution and 4 mL of H2N2·2H2O to start the reduction. The entire solution was stirred under reflux for one hour at 90 °C. RGO was filtrated under vacuum using a Millipore filter paper (pore size: 0.2 μm), washed with Milli-Q H2O and absolute ethanol and dried at 50 °C for 12 h. All the composites were prepared via impregnation, growing Cu2O nanoparticles on the support through an adapted method [58]. In particular, a proper amount of RGO (5, 10 and 20 mg, respectively, for the 2:1, 1:1 and 1:2 composites) was mixed with 0.348 g of SDS and 35.65 mL of Milli-Q H2O and sonicated for 20 min at 30 °C. In total, 2 mL of a CuCl2 0.1 M solution were introduced dropwise under vigorous stirring and 0.72 mL of NaOH 1M and 1.6 mL of NH2OH·HCl were added very quickly to the solution of CuCl2, forming the Cu2O nucleation seeds. The mixture was aged for 60 min at 30 °C, while the color of the solution changed from light blue to orange, confirming the formation of Cu2O on the support. The materials were washed three times with absolute ethanol, separated from the solution by filtration and dried for 12 h at 45 °C in order to remove the surfactants from the surface of Cu2O.

3.2. Characterization

The crystalline structure of the different materials was investigated by X-ray diffraction (XRD) using a high-resolution X-ray powder diffractometer (PANalytical X’Pert Pro MPD) with Cu Kα radiation (λ = 0.1541 nm). Raman spectra have been acquired using a Renishaw InVia™ confocal Raman microscope coupled with a 632 nm laser. The morphology of fabricated nanostructures was investigated using a TEM JEOL 2010 with a LaB6 emission gun operating at 160 kV and a Hitachi FE-SEM 4800 SEM. EDX measurements had been performed using a Jeol-7900F SEM microscope with accelerating voltage of 5 kV. An Autolab PGSTAT 302N electrochemical workstation (Metrohm, The Netherlands) coupled with Nova softwarehad been used for all the electrochemical measurements. The analysis of the gaseous products (CO, CH4, CH2CH2, CH3CH3) was performed using a J&W Select Permanent Gases/CO2 column connected to a methanizer and a FID detector. The analysis of H2 amount was performed using a Molsieve 5A plot column (Restek, 30m, 0.53 mm ID, 30 µm film) connected to a TCD detector. HCOOH accumulated in the liquid phase has been determined by ion chromatography using a Metrohm 833 instrument, mounting a column Metrosep A Supp 19-250/4.0 with eluent NaHCO3 1.0 mM/Na2CO3 3.2 mM.

3.3. Electrochemical Characterization

The Working Electrodes (WE) were prepared by drop casting 10 µL of an ink of the catalyst on a Rotating Disk Electrode (RDE), coupled with a graphite electrode as the Counter Electrode (CE) and a Saturated Calomel Electrode (SCE) as the Reference Electrode (RE). The ink was prepared dissolving 3.0 mg of catalyst in 1 mL of a 1:9 2-propanol:H2O solution and 50 µL of Nafion. All the electrochemical characterization had been performed by rotating the RDE at 1600 RPM in a 0.1 M KHCO3 solution. Linear Sweep Voltammetries (LSV) had been performed by choosing a lower voltage limit and using a scan rate of 5 mV/s. The voltages were converted to RHE using the Equation (1)
V R H E = V S C E + 0.244   V + 0.059 × p H
Each LSV was repeated for three times to desorb all the gaseous species adsorbed on the surface of the electrode and to obtain a clear indication of the electrochemical behavior of the bare materials. All the experiments were carried out both in the Ar and CO2 atmosphere, correcting the voltage with iR compensation. Even if the determination of the ECSA was performed to better characterize the electrochemical properties [59], it was decided to normalize the current measurements to the geometric area of the electrode due to the variability of the electrocatalysts deposition. Electrochemical Impedance Spectroscopy (EIS) analysis was performed at Open Circuit Voltage (OCV) in the frequency range of 100 kHz to 1 Hz. According to the literature [60], a Thin-Film (TF)–Rotating Disc Working Electrode (three–electrodes) cell setup was used to minimize the capacitance related to ionomer and carbon components.

3.4. CO2RR Experiments

The WE was prepared by drop casting an ink of different catalyst powder on top of Toray carbon paper support. In a typical experiment, 3.0 mg of catalyst, 0.9 mL of milliQ H2O, 0.1 mL of i-propanol and 50 μL of Nafion were mixed and sonicated for one hour. In total, 100 μL of solution was deposited on both sides of the carbon support, covering an area of 10 mm × 5 mm. The properly modified Toray paper support is used as WE in a gas-tight electrochemical cell, equipped with a platinum wire (1.00 mm as diameter, 99.9% of purity, Sigma Aldrich, St. Louis, MI, USA) as the CE and Ag/AgCl sat. electrode as the RE. The electrochemical characterization was performed purging 20 mL/min of CO2 in 15 mL of KHCO3 0.1 M (purity > 99.95%, Sigma Aldrich). The different electrocatalyst materials were studied performing chronoamperometries for 120 min, analyzing the gaseous products through gas-chromatography and the liquid products by ionic chromatography by taking 1 mL of aliquot and replacing it with fresh electrolyte. A 5 cycle-CV with a scan rate of 100 mV/s was performed before each experiment in order to clean the surface of the catalyst from the adsorbed gaseous species. CAs had been studied at −1.3 V, −1.5 V and −1.7 V vs. Ag/AgCl, using the Equation (2) to convert the voltages to RHE.
V R H E = V A g / A g C l + 0.197   V + 0.059 × p H
All the voltages were corrected for iR compensation and normalized for the geometrical area and each WE was substituted after each electrochemical characterization. The total Faradaic Efficiency (FE) was calculated from the final concentration of the different products during the time by Equation (3):
F E = n e × n × F Q
where n e the number of electrons involved in the particular reductive reaction, n is the total moles of the product, F is the Faraday constant and Q is the total charge.

4. Conclusions

In summary, in this work we successfully prepared different composites of Cu2O/RGO through a multistep approach where nanocubes of Cu2O were grown on crumpled RGO nanosheets. The detailed characterization of the prepared materials provided valuable insights into the structural and morphological properties, observing the purity of the crystalline phase and the contemporaneous coexistence of Cu2O and RGO in the composites. SEM and TEM images unveiled the morphology of the materials, revealing an intimate contact between the two counterparts of the material and the beneficial effects on size and distribution of Cu2O nanoparticles. The electrocatalytic performance of the synthetized materials was evaluated through a three-electrode setup highlighting the improved overall activity of the composites thanks to the enhanced dispersion of Cu2O and the intimate contact with RGO sheets. CO2RR experiments revealed an enhanced current density for the composites, especially at lower voltages, and improved faradaic efficiencies for the products of the reduction, in particular CO. The sample Cu2O/RGO 2:1 showed a selective formation of CO over H2, with a faradaic efficiency of 50% at −0.9 V vs. RHE and revealed the advantageous pathway towards the formation of *CO intermediate even at lower voltages. The increased current (−2.72 mA/cm2 for CU Cu2O/RGO 2:1, compared to −1.79 mA/cm2 for Cu2O at the voltage of −0.9 V vs. RHE and in the CO2 atmosphere) and selectivity in this composition resulted in significantly higher amounts of produced CO, providing a valuable solution for the formation of effective interphases and the promotion of catalytic properties in nanostructured materials.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal14070412/s1, Figure S1: Rietveld analysis of the CU Cu2O/RGO 2:1 diffractogram. Figure S2: Raman spectrum of pure RGO with the characteristics vibrations. Table S1: Weight fraction of the different CU Cu2O/RGO materials from TGA analysis. Figure S3: SEM and TEM images of pristine RGO and CU Cu2O. Figure S4: Size distribution for the different materials. Table S2: Size of Cu2O nanoparticles on the different composites. Figure S5: LSVs at −0.6 V, −0.8 V and −1.0 V vs. RHE for the different materials. Figure S6: CAs at −0.6 V, −0.8 V and −1.0 V vs. RHE for the different materials. Figure S7: Nyquist plot at OCV of the different materials performed in CO2 saturated atmosphere. Figure S8: CO2RR tests for the different composites of CU Cu2O/RGO at the voltage of −0.7 V vs. RHE. Figure S9: CO2RR tests for the different composites of CU Cu2O/RGO at the voltage of −1.1 V vs. RHE. Table S3: Comparison with literature. Table S4: ECSA obtained for the different materials. Refences.

Author Contributions

Conceptualization, M.B. (Matteo Bisetto), M.B. (Manuela Bevilacqua). and P.F.; methodology, M.B. (Matteo Bisetto), S.R., M.B. (Manuela Bevilacqua) and T.M.; validation, T.M. and A.N.; formal analysis, M.B. (Matteo Bisetto); investigation, M.B. (Matteo Bisetto) and S.R.; data curation, M.B. (Matteo Bisetto) and M.B. (Manuela Bevilacqua); writing—original draft preparation, M.B. (Matteo Bisetto) and M.B. (Manuela Bevilacqua); writing—review and editing, A.N., T.M. and P.F.; resources, A.N. and P.F.; supervision, P.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the European Community, (projects H2020–LC–SC3-2019-NZE-RES-CC), grant agreement number 884444. A.N. acknowledges the support from the Project CH4.0 under the MIUR program “Dipartimenti di Eccellenza 2023-2027” (CUP: D13C2200352001).

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

We would like to thank Luca Mascaretti for the measurements at RCPTM, Giuliano Giambastiani for the TGA measurements and Elvio Merlac from the University of Trieste for the development of the electrochemical device.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Khdary, N.H.; Alayyar, A.S.; Alsarhan, L.M.; Alshihri, S.; Mokhtar, M. Metal Oxides as Catalyst/Supporter for CO2 Capture and Conversion, Review. Catalysts 2022, 12, 300. [Google Scholar] [CrossRef]
  2. Xia, Q.; Zhang, K.; Zheng, T.; An, L.; Xia, C.; Zhang, X. Integration of CO2 Capture and Electrochemical Conversion. ACS Energy Lett. 2023, 8, 2840–2857. [Google Scholar] [CrossRef]
  3. He, Y.; Müller, F.H.; Palkovits, R.; Zeng, F.; Mebrahtu, C. Tandem Catalysis for CO2 Conversion to Higher Alcohols: A Review. Appl. Catal. B Environ. 2024, 345, 123663. [Google Scholar] [CrossRef]
  4. Ma, Y.; Wang, Z.; Xu, X.; Wang, J. Review on Porous Nanomaterials for Adsorption and Photocatalytic Conversion of CO2. Cuihua Xuebao/Chinese J. Catal. 2017, 38, 1956–1969. [Google Scholar] [CrossRef]
  5. Lee, Y.Y.; Jung, H.S.; Kim, J.M.; Kang, Y.T. Photocatalytic CO2 Conversion on Highly Ordered Mesoporous Materials: Comparisons of Metal Oxides and Compound Semiconductors. Appl. Catal. B Environ. 2018, 224, 594–601. [Google Scholar] [CrossRef]
  6. Christoforidis, K.C.; Fornasiero, P. Photocatalysis for Hydrogen Production and CO2 Reduction: The Case of Copper-Catalysts. ChemCatChem 2019, 11, 368–382. [Google Scholar] [CrossRef]
  7. Verma, R.; Belgamwar, R.; Polshettiwar, V. Plasmonic Photocatalysis for CO2 Conversion to Chemicals and Fuels. ACS Mater. Lett. 2021, 3, 574–598. [Google Scholar] [CrossRef]
  8. Rej, S.; Bisetto, M.; Naldoni, A.; Fornasiero, P. Well-Defined Cu2O Photocatalysts for Solar Fuels and Chemicals. J. Mater. Chem. A 2021, 9, 5915–5951. [Google Scholar] [CrossRef]
  9. Zhang, Y.; Huang, Y.; Zhu, S.S.; Liu, Y.Y.; Zhang, X.; Wang, J.J.; Braun, A. Covalent S-O Bonding Enables Enhanced Photoelectrochemical Performance of Cu2S/Fe2O3 Heterojunction for Water Splitting. Small 2021, 17, 2100320. [Google Scholar] [CrossRef]
  10. Ma, X.; Zhang, Y.; Fan, T.; Wei, D.; Huang, Z.; Zhang, Z.; Zhang, Z.; Dong, Y.; Hong, Q.; Chen, Z.; et al. Facet Dopant Regulation of Cu2O Boosts Electrocatalytic CO2 Reduction to Formate. Adv. Funct. Mater. 2023, 33, 2213145. [Google Scholar] [CrossRef]
  11. Jun, M.; Kwak, C.; Lee, S.Y.; Joo, J.; Kim, J.M.; Im, D.J.; Cho, M.K.; Baik, H.; Hwang, Y.J.; Kim, H.; et al. Microfluidics-Assisted Synthesis of Hierarchical Cu2O Nanocrystal as C2-Selective CO2 Reduction Electrocatalyst. Small Methods 2022, 6, 2200074. [Google Scholar] [CrossRef] [PubMed]
  12. Ren, D.; Deng, Y.; Handoko, A.D.; Chen, C.S.; Malkhandi, S.; Yeo, B.S. Selective Electrochemical Reduction of Carbon Dioxide to Ethylene and Ethanol on Copper(I) Oxide Catalysts. ACS Catal. 2015, 5, 2814–2821. [Google Scholar] [CrossRef]
  13. Mistry, H.; Varela, A.S.; Bonifacio, C.S.; Zegkinoglou, I.; Sinev, I.; Choi, Y.W.; Kisslinger, K.; Stach, E.A.; Yang, J.C.; Strasser, P.; et al. Highly Selective Plasma-Activated Copper Catalysts for Carbon Dioxide Reduction to Ethylene. Nat. Commun. 2016, 7, 12123. [Google Scholar] [CrossRef]
  14. Wang, S.; Kou, T.; Baker, S.E.; Duoss, E.B.; Li, Y. Recent Progress in Electrochemical Reduction of CO2 by Oxide-Derived Copper Catalysts. Mater. Today Nano 2020, 12, 100096. [Google Scholar] [CrossRef]
  15. Liu, J.; Cheng, L.; Wang, Y.; Chen, R.; Xiao, C.; Zhou, X.; Zhu, Y.; Li, Y.; Li, C. Dynamic Determination of Cu+ Roles for CO2 Reduction on Electrochemically Stable Cu2O-Based Nanocubes. J. Mater. Chem. A 2022, 10, 8459–8465. [Google Scholar] [CrossRef]
  16. Jiang, Y.; Wang, X.; Duan, D.; He, C.; Ma, J.; Zhang, W.; Liu, H.; Long, R.; Li, Z.; Kong, T.; et al. Structural Reconstruction of Cu2O Superparticles toward Electrocatalytic CO2 Reduction with High C2+ Products Selectivity. Adv. Sci. 2022, 9, 2105292. [Google Scholar] [CrossRef]
  17. Larrazábal, G.O.; Okatenko, V.; Chorkendorff, I.; Buonsanti, R.; Seger, B. Investigation of Ethylene and Propylene Production from CO2 Reduction over Copper Nanocubes in an MEA-Type Electrolyzer. ACS Appl. Mater. Interfaces 2022, 14, 7779–7787. [Google Scholar] [CrossRef] [PubMed]
  18. Guo, S.; Liu, Y.; Huang, Y.; Wang, H.; Murphy, E.; Delafontaine, L.; Chen, J.; Zenyuk, I.V.; Atanassov, P. Promoting Electrolysis of Carbon Monoxide toward Acetate and 1-Propanol in Flow Electrolyzer. ACS Energy Lett. 2023, 8, 935–942. [Google Scholar] [CrossRef]
  19. Wu, Q.; Du, R.; Wang, P.; Waterhouse, G.I.N.; Li, J.; Qiu, Y.; Yan, K.; Zhao, Y.; Zhao, W.W.; Tsai, H.J.; et al. Nanograin-Boundary-Abundant Cu2O-Cu Nanocubes with High C2+ Selectivity and Good Stability during Electrochemical CO2 Reduction at a Current Density of 500 mA/cm2. ACS Nano 2023, 17, 12884–12894. [Google Scholar] [CrossRef]
  20. Zaza, L.; Rossi, K.; Buonsanti, R. Well-Defined Copper-Based Nanocatalysts for Selective Electrochemical Reduction of CO2 to C2 Products. ACS Energy Lett. 2022, 7, 1284–1291. [Google Scholar] [CrossRef]
  21. Rossi, K.; Buonsanti, R. Shaping Copper Nanocatalysts to Steer Selectivity in the Electrochemical CO2 Reduction Reaction. Acc. Chem. Res. 2022, 55, 629–637. [Google Scholar] [CrossRef] [PubMed]
  22. Liu, L.; Corma, A. Metal Catalysts for Heterogeneous Catalysis: From Single Atoms to Nanoclusters and Nanoparticles. Chem. Rev. 2018, 118, 4981–5079. [Google Scholar] [CrossRef] [PubMed]
  23. Zhai, Z.; Guo, X.; Jiao, Z.; Jin, G.; Guo, X.Y. Graphene-Supported Cu2O Nanoparticles: An Efficient Heterogeneous Catalyst for C-O Cross-Coupling of Aryl Iodides with Phenols. Catal. Sci. Technol. 2014, 4, 4196–4199. [Google Scholar] [CrossRef]
  24. Abu-Zied, B.M.; Hussein, M.A.; Khan, A.; Asiri, A.M. Cu-Cu2O@graphene Nanoplatelets Nanocomposites: Facile Synthesis, Characterization, and Electrical Conductivity Properties. Mater. Chem. Phys. 2018, 213, 168–176. [Google Scholar] [CrossRef]
  25. Alhaddad, M.; Navarro, R.M.; Hussein, M.A.; Mohamed, R.M. Visible Light Production of Hydrogen from Glycerol over Cu2O-gC3N4 Nanocomposites with Enhanced Photocatalytic Efficiency. J. Mater. Res. Technol. 2020, 9, 15335–15345. [Google Scholar] [CrossRef]
  26. Wu, Z.; Wang, C.; Zhang, X.; Guo, Q.; Wang, J. Graphene-Based CO2 Reduction Electrocatalysts: A Review. New Carbon Mater. 2024, 39, 100–130. [Google Scholar] [CrossRef]
  27. Zhang, D.; Wei, D.; Cui, Z.; Wang, S.; Yang, S.; Cao, M.; Hu, C. Improving Water Splitting Performance of Cu2O through a Synergistic “Two-Way Transfer” Process of Cu and Graphene. Phys. Chem. Chem. Phys. 2014, 16, 25531–25536. [Google Scholar] [CrossRef] [PubMed]
  28. Nemiwal, M.; Zhang, T.C.; Kumar, D. Graphene-Based Electrocatalysts: Hydrogen Evolution Reactions and Overall Water Splitting. Int. J. Hydrogen Energy 2021, 46, 21401–21418. [Google Scholar] [CrossRef]
  29. Sharma, V.; Jain, Y.; Kumari, M.; Gupta, R.; Sharma, S.K.; Sachdev, K. Synthesis and Characterization of Graphene Oxide (GO) and Reduced Graphene Oxide (RGO) for Gas Sensing Application. Macromol. Symp. 2017, 376, 1700006. [Google Scholar] [CrossRef]
  30. Yoon, Y.; Kye, H.; Yang, W.S.; Kang, J.W. Comparing Graphene Oxide and Reduced Graphene Oxide as Blending Materials for Polysulfone and Polyvinylidene Difluoride Membranes. Appl. Sci. 2020, 10, 2015. [Google Scholar] [CrossRef]
  31. Jaafar, E.; Kashif, M.; Sahari, S.K.; Ngaini, Z. Study on Morphological, Optical and Electrical Properties of Graphene Oxide (GO) and Reduced Graphene Oxide (RGO). Mater. Sci. Forum 2018, 917, 112–116. [Google Scholar] [CrossRef]
  32. Wang, W.; Ning, H.; Yang, Z.; Feng, Z.; Wang, J.; Wang, X.; Mao, Q.; Wu, W.; Zhao, Q.; Hu, H.; et al. Interface-Induced Controllable Synthesis of Cu2O Nanocubes for Electroreduction CO2 to C2H4. Electrochim. Acta 2019, 306, 360–365. [Google Scholar] [CrossRef]
  33. Susman, M.D.; Feldman, Y.; Vaskevich, A.; Rubinstein, I. Chemical Deposition of Cu2O Nanocrystals with Precise Morphology Control. ACS Nano 2014, 8, 162–174. [Google Scholar] [CrossRef] [PubMed]
  34. An, X.; Li, K.; Tang, J. Cu2O/Reduced Graphene Oxide Composites for the Photocatalytic Conversion of CO2. ChemSusChem 2014, 7, 1086–1093. [Google Scholar] [CrossRef] [PubMed]
  35. Fu, W.; Liu, Z.; Wang, T.; Liang, J.; Duan, S.; Xie, L.; Han, J.; Li, Q. Promoting C2+ Production from Electrochemical CO2 Reduction on Shape-Controlled Cuprous Oxide Nanocrystals with High-Index Facets. ACS Sustain. Chem. Eng. 2020, 8, 15223–15229. [Google Scholar] [CrossRef]
  36. Xu, J.; Wang, R.; Chen, X.; Zhou, R.; Zhang, J. Cu2SnS3 Nanocrystals Decorated RGO Nanosheets towards Efficient and Stable Hydrogen Evolution Reaction in Both Acid and Alkaline Solutions. Mater. Today Energy 2020, 17, 100435. [Google Scholar] [CrossRef]
  37. Khanra, P.; Kuila, T.; Kim, N.H.; Bae, S.H.; Yu, D.-S.; Lee, J.H. Simultaneous Bio-Functionalization and Reduction of Graphene Oxide by Baker’s Yeast. Chem. Eng. J. 2012, 183, 526–533. [Google Scholar] [CrossRef]
  38. Sadhukhan, S.; Ghosh, T.K.; Rana, D.; Roy, I.; Bhattacharyya, A.; Sarkar, G.; Chakraborty, M.; Chattopadhyay, D. Studies on Synthesis of Reduced Graphene Oxide (RGO) via Green Route and Its Electrical Property. Mater. Res. Bull. 2016, 79, 41–51. [Google Scholar] [CrossRef]
  39. Wang, Y.; Shi, Z.X.; Yin, J. Facile Synthesis of Soluble Graphene via a Green Reduction of Graphene Oxide in Tea Solution and Its Biocomposites. ACS Appl. Mater. Interfaces 2011, 3, 1127–1133. [Google Scholar] [CrossRef]
  40. Cui, P.; Lee, J.; Hwang, E.; Lee, H. One-Pot Reduction of Graphene Oxide at Subzero Temperatures. Chem. Commun. 2011, 47, 12370–12372. [Google Scholar] [CrossRef]
  41. Kuila, T.; Mishra, A.K.; Khanra, P.; Kim, N.H.; Lee, J.H. Recent Advances in the Efficient Reduction of Graphene Oxide and Its Application as Energy Storage Electrode Materials. Nanoscale 2013, 5, 52–71. [Google Scholar] [CrossRef] [PubMed]
  42. Gao, L.; Pang, C.; He, D.; Shen, L.; Gupta, A.; Bao, N. Synthesis of Hierarchical Nanoporous Microstructures via the Kirkendall Effect in Chemical Reduction Process. Sci. Rep. 2015, 5, 16061. [Google Scholar] [CrossRef] [PubMed]
  43. Möller, T.; Scholten, F.; Thanh, T.N.; Sinev, I.; Timoshenko, J.; Wang, X.; Jovanov, Z.; Gliech, M.; Roldan Cuenya, B.; Varela, A.S.; et al. Electrocatalytic CO2 Reduction on CuOx Nanocubes: Tracking the Evolution of Chemical State, Geometric Structure, and Catalytic Selectivity Using Operando Spectroscopy. Angew. Chemie Int. Ed. 2020, 59, 17974–17983. [Google Scholar] [CrossRef] [PubMed]
  44. Nallal, M.; Park, K.H.; Park, S.; Kim, J.; Shenoy, S.; Chuaicham, C.; Sasaki, K.; Sekar, K. Cu2O/Reduced Graphene Oxide Nanocomposites for Electrocatalytic Overall Water Splitting. ACS Appl. Nano Mater. 2022, 5, 17271–17280. [Google Scholar] [CrossRef]
  45. Zhong, Y.; Wang, S.; Li, M.; Ma, J.; Song, S.; Kumar, A.; Duan, H.; Kuang, Y.; Sun, X. Rational Design of Copper-Based Electrocatalysts and Electrochemical Systems for CO2 Reduction: From Active Sites Engineering to Mass Transfer Dynamics. Mater. Today Phys. 2021, 18, 100354. [Google Scholar] [CrossRef]
  46. Dattila, F.; Garclá-Muelas, R.; López, N. Active and Selective Ensembles in Oxide-Derived Copper Catalysts for CO2 Reduction. ACS Energy Lett. 2020, 5, 3176–3184. [Google Scholar] [CrossRef]
  47. Cao, X.; Cao, G.; Li, M.; Zhu, X.; Han, J.; Ge, Q.; Wang, H. Enhanced Ethylene Formation from Carbon Dioxide Reduction through Sequential Catalysis on Au Decorated Cubic Cu2O Electrocatalyst. Eur. J. Inorg. Chem. 2021, 2021, 2353–2364. [Google Scholar] [CrossRef]
  48. Zheng, T.; Liu, C.; Guo, C.; Zhang, M.; Li, X.; Jiang, Q.; Xue, W.; Li, H.; Li, A.; Pao, C.W.; et al. Copper-Catalysed Exclusive CO2 to Pure Formic Acid Conversion via Single-Atom Alloying. Nat. Nanotechnol. 2021, 16, 1386–1393. [Google Scholar] [CrossRef] [PubMed]
  49. Lee, C.W.; Kim, C.; Min, B.K. Theoretical Insights into Selective Electrochemical Conversion of Carbon Dioxide. Nano Converg. 2019, 6, 8. [Google Scholar] [CrossRef]
  50. Goyal, A.; Marcandalli, G.; Mints, V.A.; Koper, M.T.M. Competition between CO2 Reduction and Hydrogen Evolution on a Gold Electrode under Well-Defined Mass Transport Conditions. J. Am. Chem. Soc. 2020, 142, 4154–4161. [Google Scholar] [CrossRef]
  51. Cao, C.; Wen, Z. Cu Nanoparticles Decorating RGO Nanohybrids as Electrocatalyst toward CO2 Reduction. J. CO2 Util. 2017, 22, 231–237. [Google Scholar] [CrossRef]
  52. Ning, H.; Mao, Q.; Wang, W.; Yang, Z.; Wang, X.; Zhao, Q.; Song, Y.; Wu, M. N-Doped Reduced Graphene Oxide Supported Cu2O Nanocubes as High Active Catalyst for CO2 Electroreduction to C2H4. J. Alloys Compd. 2019, 785, 7–12. [Google Scholar] [CrossRef]
  53. Bochlin, Y.; Korin, E.; Bettelheim, A. Different Pathways for CO2 Electrocatalytic Reduction by Confined CoTMPyP in Electrodeposited Reduced Graphene Oxide. ACS Appl. Energy Mater. 2019, 2, 8434–8440. [Google Scholar] [CrossRef]
  54. Nguyen, D.L.T.; Lee, C.W.; Na, J.; Kim, M.C.; Tu, N.D.K.; Lee, S.Y.; Sa, Y.J.; Won, D.H.; Oh, H.S.; Kim, H.; et al. Mass Transport Control by Surface Graphene Oxide for Selective CO Production from Electrochemical CO2 Reduction. ACS Catal. 2020, 10, 3222–3231. [Google Scholar] [CrossRef]
  55. Jiang, X.; Wang, Q.; Xiao, X.; Chen, J.; Shen, Y.; Wang, M. Interfacial Engineering of Bismuth with Reduced Graphene Oxide Hybrid for Improving CO2 Electroreduction Performance. Electrochim. Acta 2020, 357, 136840. [Google Scholar] [CrossRef]
  56. Ramakrishnan, M.C.; Thangavelu, R.R. Synthesis and Characterization of Reduced Graphene Oxide. Adv. Mater. Res. 2013, 678, 56–60. [Google Scholar] [CrossRef]
  57. Bansal, K.; Singh, J.; Dhaliwal, A.S. Synthesis and Characterization of Graphene Oxide and Its Reduction with Different Reducing Agents. IOP Conf. Ser. Mater. Sci. Eng. 2022, 1225, 012050. [Google Scholar] [CrossRef]
  58. Chanda, K.; Rej, S.; Huang, M.H. Investigation of Facet Effects on the Catalytic Activity of Cu2O Nanocrystals for Efficient Regioselective Synthesis of 3,5-Disubstituted Isoxazoles. Nanoscale 2013, 5, 12494–12501. [Google Scholar] [CrossRef]
  59. Zhuang, T.T.; Liang, Z.Q.; Seifitokaldani, A.; Li, Y.; De Luna, P.; Burdyny, T.; Che, F.; Meng, F.; Min, Y.; Quintero-Bermudez, R.; et al. Steering Post-C-C Coupling Selectivity Enables High Efficiency Electroreduction of Carbon Dioxide to Multi-Carbon Alcohols. Nat. Catal. 2018, 1, 421–428. [Google Scholar] [CrossRef]
  60. Singh, R.K.; Devivaraprasad, R.; Kar, T.; Chakraborty, A.; Neergat, M. Electrochemical Impedance Spectroscopy of Oxygen Reduction Reaction (ORR) in a Rotating Disk Electrode Configuration: Effect of Ionomer Content and Carbon-Support. J. Electrochem. Soc. 2015, 162, F489–F498. [Google Scholar] [CrossRef]
Figure 1. Experimental procedure for the synthesis of cubic Cu2O/RGO electrocatalysts.
Figure 1. Experimental procedure for the synthesis of cubic Cu2O/RGO electrocatalysts.
Catalysts 14 00412 g001
Figure 2. (a) XRD patterns for pristine nanoparticles of Cu2O and for the composite with RGO. (b) Raman spectra for the different synthetized composites where it is possible to observe the characteristic vibrations of Cu2O (#) and RGO (*). (c) Thermal Gravimetric Analysis for the three different composites in air with the temperature range 50–800 °C.
Figure 2. (a) XRD patterns for pristine nanoparticles of Cu2O and for the composite with RGO. (b) Raman spectra for the different synthetized composites where it is possible to observe the characteristic vibrations of Cu2O (#) and RGO (*). (c) Thermal Gravimetric Analysis for the three different composites in air with the temperature range 50–800 °C.
Catalysts 14 00412 g002
Figure 3. Microscope analysis for the composite Cu2O/RGO 2:1. (a,b) TEM images at different magnification levels. (c) SEM image of the cubic nanoparticles of Cu2O growth on RGO surface. (d) Size distribution of Cu2O nanoparticles (count: 400 nanoparticles). (e) EDX image and elemental distribution for the considered composite.
Figure 3. Microscope analysis for the composite Cu2O/RGO 2:1. (a,b) TEM images at different magnification levels. (c) SEM image of the cubic nanoparticles of Cu2O growth on RGO surface. (d) Size distribution of Cu2O nanoparticles (count: 400 nanoparticles). (e) EDX image and elemental distribution for the considered composite.
Catalysts 14 00412 g003
Figure 4. CO2RR tests for the different composites of CU Cu2O/RGO at the voltage of −0.9 V vs. RHE. (a) Two hour chronoamperometries under saturated CO2 atmosphere. (b) FEs of gaseous and liquid products. (c) Products distribution (μmol) of the different compounds after 2 h chronoamperometries. (d) Production profile of CO during the experiment for the different composite materials.
Figure 4. CO2RR tests for the different composites of CU Cu2O/RGO at the voltage of −0.9 V vs. RHE. (a) Two hour chronoamperometries under saturated CO2 atmosphere. (b) FEs of gaseous and liquid products. (c) Products distribution (μmol) of the different compounds after 2 h chronoamperometries. (d) Production profile of CO during the experiment for the different composite materials.
Catalysts 14 00412 g004
Figure 5. Faradaic efficiencies at the different voltages for the three composite CU Cu2O/RGO. (a) CU Cu2O/RGO 1:2, (b) CU Cu2O/RGO 1:1 and (c) CU Cu2O/RGO 2:1.
Figure 5. Faradaic efficiencies at the different voltages for the three composite CU Cu2O/RGO. (a) CU Cu2O/RGO 1:2, (b) CU Cu2O/RGO 1:1 and (c) CU Cu2O/RGO 2:1.
Catalysts 14 00412 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bisetto, M.; Rej, S.; Naldoni, A.; Montini, T.; Bevilacqua, M.; Fornasiero, P. CO2 Electroreduction by Engineering the Cu2O/RGO Interphase. Catalysts 2024, 14, 412. https://doi.org/10.3390/catal14070412

AMA Style

Bisetto M, Rej S, Naldoni A, Montini T, Bevilacqua M, Fornasiero P. CO2 Electroreduction by Engineering the Cu2O/RGO Interphase. Catalysts. 2024; 14(7):412. https://doi.org/10.3390/catal14070412

Chicago/Turabian Style

Bisetto, Matteo, Sourav Rej, Alberto Naldoni, Tiziano Montini, Manuela Bevilacqua, and Paolo Fornasiero. 2024. "CO2 Electroreduction by Engineering the Cu2O/RGO Interphase" Catalysts 14, no. 7: 412. https://doi.org/10.3390/catal14070412

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