Next Article in Journal
Anticancer Nanotherapeutics in Clinical Trials: The Work behind Clinical Translation of Nanomedicine
Next Article in Special Issue
Identifying the Active Sites of Heteroatom Graphene as a Conductive Membrane for the Electrochemical Filtration of Organic Contaminants
Previous Article in Journal
Proteomic Profiling Revealed Mitochondrial Dysfunction in Photoreceptor Cells under Hyperglycemia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Insights of Fe2O3 and MoO3 Electrodes for Electrocatalytic CO2 Reduction in Aprotic Media

by
Néstor E. Mendieta-Reyes
,
Alejandra S. Lozano-Pérez
and
Carlos A. Guerrero-Fajardo
*
Departamento de Química, Facultad de Ciencias, Universidad Nacional de Colombia, Bogotá 11321, Colombia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(21), 13367; https://doi.org/10.3390/ijms232113367
Submission received: 6 September 2022 / Revised: 14 October 2022 / Accepted: 16 October 2022 / Published: 2 November 2022

Abstract

:
Transition metal oxides (TMO) have been successfully used as electrocatalytically active materials for CO2 reduction in some studies. Because of the lack of understanding of the catalytic behavior of TMOs, electrochemical methods are used to investigate the CO2 reduction in thin-film nanostructured electrodes. In this context, nanostructured thin films of Fe2O3 and MoO3 in an aprotic medium of acetonitrile have been used to study the CO2 reduction reaction. In addition, a synergistic effect between CO2 and the TMO surface is observed. Faradic cathodic processes not only start at lower potentials than those reported with metal electrodes, but also an increase in capacitive currents is observed, which is directly related to an increase in oxygen vacancies. Finally, the results obtained show CO as a product of the reduction.

1. Introduction

The current atmospheric levels of carbon dioxide and the depletion of fossil fuel reserves raise serious concerns about the resulting effects on the global climate change and on the future of some energy sources [1]. The electrochemical conversion of carbon dioxide is a promising way to address both problems simultaneously; this strategy provides an opportunity for the supply of platform chemicals, which can be used as supplementary raw materials in several industrial processes, or energy sources [2]. A wide variety of methods, such as photocatalysis, electrocatalysis, and heterogeneous catalysis, have all been successful in CO2 reduction [3]. Factors such as solvent, electrolyte, pH, temperature, CO2 concentration, electrode material, surface structure, and electrode potential influence the variety of products obtained [4,5]. Reduction products, including CO, CH4, and CH3OH, among others, have been reported [6,7,8]. This leads to the development of several reaction mechanisms with different numbers of electrons transferred, which determines the final oxidation state of the carbon atom [9]. However, due to the linear geometry with a closed-shell electronic configuration, the high energy level of its LUMO, and the low electron affinity of the molecule (0.6 eV), the reduction of CO2 is a non-spontaneous reaction requiring an energy input. Due to the repulsion between the newly acquired electron and the free electron pairs on the oxygen atoms, the transfer of an electron results in the formation of the anion radical CO2(.−), which causes the molecular structure to bend [10,11,12,13].
In this sense, several reports have shown that the stabilization of the CO2(.−) anion radical is an important step for achieving an efficient reduction of CO2. The CO2 molecule can then be activated on a solid-state catalyst, whose function is to adsorb the CO2 molecules by facilitating electron transference from the catalyst, providing favorable conditions for the bending of the molecule and the desorption of the products [10,11,14,15]. In this regard, the adduct formation of CO2(δ.−) on different metal oxides has been widely reported [16,17].
On the other hand, in an aqueous solution, different factors, such as acid-base equilibria and hydrogen synthesis occurring as a competitive process, act against obtaining high efficiency and selectivity. In this sense, acetonitrile (ACN) is one of the most widely used non-aqueous aprotic solvents in catalysis and electrochemistry, and has been used to study the electrocatalytic CO2 reduction [18,19,20]. Pons et al. studied solutions of tetra-n-butylammonium tetrafluoroborate (TBAF) in dry ACN, finding ACN decomposition at potentials more negative than –3.0 V vs. Ag/Ag+ (0.01 M) [5]. The ACN solution containing tetrabutylammonium salts, such as the electrolyte system, TBAF (0.1 M), has been used for the study of the electrocatalytic reduction of CO2. This non-aqueous system not only offers stability at a potential window more negative than −1.97 V vs. SHE, which is needed for the first CO2 electron transfer, but also that avoids the hydrogen synthesis as a competitive reaction.
Metal oxides are used as electrocatalysts for CO2 reduction [21], and transition metal oxides (TMO) have been theoretically predicted to have efficient active sites for said reduction reaction [22]. In fact, these materials have reactive electrons, empty d orbitals, and crystalline defects, such as oxygen vacancies (OV), grain boundaries, and non-stoichiometry, that are capable of triggering the CO2 molecule activation, inducing the formation of the CO2(δ·−) adduct [11]. Compared with metal catalysts, TMO can further improve the performance of CO2 reduction, which has drawn attention in recent studies. Some researchers have proposed that the active sites for CO2 in metal oxides are due to the reconstruction of their surface, formed by the reduced surface of the metal oxide at highly negative operating potentials [23], whereas others have highlighted metal-oxygen sites as active, leaving an open debate on this topic [24]. In this context and based on the dissociative reduction model previously reported for CO2 conversion at WO3 and TiO2 electrodes, results on Fe2O3 and MoO3 electrodes are shown for the CO2 electrochemical reduction. These findings provide an explanation for the electrocatalytic activity seen in other TMOs that have been previously studied, as well as their relationship with surface processes and the formation of OV, in terms of cathodic current and the initial potential that is related to CO2 reduction [25,26,27].

2. Results and Discussion

Fe2O3 and MoO3 thin-film electrodes were coated on Fluorine-doped Tin Oxide (FTO) substrates following two different synthesis routes, as explained in the Supplementary Information, leading to different morphologies. Synthesized Fe2O3 thin-film surface characterization is shown as to hematite, and a highly coated nanostructured electrode surface displaying 50 nm-thick nanoparticles is observed (Figure S2). An oxide film cross-section image, with approximately 500 nm-thickness, shows partially oriented nanobars. FE-SEM images for MoO3 electrodes, at two different magnification levels of 5000× and 25,000×, are shown in Figure S3. Electrodes prepared by sol-gel process are thoroughly coated, with no uncoated substrate areas (Figure S3a). The morphology of the nanoparticles (Figure S3b) is in accordance with electrodes synthesized by the sol-gel technique. The observed nanoparticles’ diameter is between 20–40 nm, with crystal aggregates well interconnected with each other. The X-ray diffraction pattern of the FTO-coated thin-film electrodes presented in Figure S4 indicates that the as-prepared product presents high-crystallinity.
For electrochemical studies, the prepared electrodes were evaluated in a 0.1 M tetrabutylammonium perchlorate (TBAP)/ACN solution as the supporting electrolyte. The electrolyte solution was purged with CO2 and N2 for 30 min before the electrochemical procedure. Figure 1 shows the voltametric response for both the Fe2O3 and MoO3 working electrodes. The potential window for CO2 reduction was determined by purging with N2. Both Fe2O3 and MO3 electrodes show an increase in the cathodic current, with an onset at around −0.4 V and −0.6 V, respectively, which coincides with the beginning of the accumulation region. From a chemical point of view, such an increase can be seen as corresponding to Fe(ss)(III)/(II) and Mo(ss)(VI)/(V) redox processes in surface sites for Fe2O3 and MoO3 (indicated on the figure). Similar behavior has been observed in semiconductor materials such as TiO2 [28].
It has been reported that the metal reduced species in TMOs could be stabilized by small cations, such as H+ or Li+, through intercalation processes by local bronze formation [29]. However, in our case, one should bear in mind that the tetrabutylammonium cation (TBA+) is bulkier than the cavity available in the crystal lattice of metal oxide. Thus, TBA+ is sterically obstructed for intercalation. This fact not only acts in favor of OV formation, but also, the stabilization of the Fe(ss)(II) and Mo(ss)(V) metal surface centers with high electron density is not efficient. This is according with the previously reported dissociative electroreduction model [30]. Under these conditions, there is an increase in surface states energetically located close to the conduction band, and the OV formation is represented by the following equations:
1 + x MoO 3 + 2 x   e     MoO 3 x + xMoO 4 2
1 + x Fe 2 O 3 + 2 x   e     Fe 2 O 3 x + xFe 2 O 4 2
1 + x Fe 2 O 3 + 2 x   e     Fe 2 O 3 x + 2 xFeO 2
In this dissociative model, the partial electroreduction of MO3 and Fe2O3 is limited to the interfacial region, whereas no significant structural damage occurs in the underlying region of the metal oxide film. According to Equations (1) and (3), the partial electroreduction of Mo and Fe lattice atoms from Mo6+ to Mo5+ and Fe3+ to Fe2+ induces the formation of oxygen-deficient metal oxide, MoO 3 x and Fe 2 O 3 x . Additionally, this mechanism entails a loss of mass due to the electrodissolution of MoO3 and Fe2O3 by the formation of MoO 4 2 and FeO 2 [31].
Under the CO2 purged condition, faradaic currents at −0.6 V and −0.8 V are observed for Fe2O3 and MoO3, respectively. In both cases, the presence of CO2 increased the electrode capacitance, even at potentials where CO2 reduction does not occur. This fact suggests important changes in the electrode state that could be associated with an increase in the surface density of Mo+5- and Fe+2-reduced species as a result of the enhancement of the OV generation. This makes the potential at the onset of the reduction of metal atoms at TMO and the cathodic current associated with CO2 reduction indistinguishable. The increase in the capacitive current has been explained in terms of the increment of OV promoted by increasingly negative applied potential, which favors CO2 electro adsorption according to the equations:
1 + y MoO 3 + yCO 2 + 2 y   e     MoO 3 Y CO 2 y + yMoO 4 2
1 + y Fe 2 O 3 + yCO 2 + 2 y   e     Fe 2 O 3 Y CO 2 y + yFe 2 O 4 2
1 + y Fe 2 O 3 + yCO 2 + 2 y   e     Fe 2 O 3 Y CO 2 y + 2 yFeO 2
The adsorption of CO2 would, thus, promote the generation of non-stoichiometric Fe2O(3−y)(CO2)y and MoO(3−y)(CO2)y. In this sense, the adsorption of CO2 induces the stabilization of reduced metal sites. The results show that CO2 binds slightly more strongly to stoichiometrically Ti(III)-deficient sites than to Ti(IV) sites, according to CO2 adsorption studies on TiO2. Therefore, it is reasonable to assume that CO2 adsorption is favored at reduced metal sites (Mo(V) and Fe(II)). Thus, a synergistic effect between CO2 and surface sites is verified. CO2 adsorption not only activates the molecule by the adduct formation, but also induces a promotion of OV and the electron-rich center formation in reduced metal species. By further reducing the applied potential, the resulting high concentration of electrons located in the near-surface region would facilitate charge transfer from the electrode to the adsorbed CO2 molecules, leading to the appearance of faradaic currents. It is worth noting that CO has been detected by gas chromatography analysis (presented in Figure S1 in Supporting Information), and the formation proceeds with two electrons transferred, according to Equations (7) and (8):
1 + y MoO 3 y CO 2 y + 2 y e   MoO 3 y + yCO + yMoO 4   2
1 + y Fe 2 O 3 + yCO 2 + 2 y   e     F e 2 O 3 y + yCO + yFe 2 O 4 2
In the positive-going scan, a pseudocapacitive peak related to the reoxidation of the reduced metal is observed, and an enhancement in the electrode capacitance related to the increase in reduced metal sites Fe(ss)(II)/(III) and Mo(ss)(V)/(VI) is evident. Moreover, in the case of Fe2O3, an additional peak, probably related to the reoxidation of some CO2 reduction product, appears at 0.3 V. This fact is intriguing because it demonstrates that each metal oxide has unique electrocatalytic behavior for CO2 reduction and the formation of reduction products. It is important to highlight that carbon dioxide reduction does not proceed when a bare FTO conducting glass electrode is used.
To determinate the nature of the currents associated with CO2 electrocatalytic reduction, a linear sweep voltammetry (LSV) was performed. At a relatively low scan rate (5 mV·s−1), the capacitive current decreases, and the faradic currents are mainly observed. Figure 2 shows the initial potential at which the cathodic current for CO2 reduction appears, providing further evidence on the faradaic nature of the cathodic currents linked to CO2 reduction. However, the presence of CO2 generates changes in current profiles, and not only produces faradic currents due to CO2 reduction, but also increases capacitive currents due to the formation of electron-rich centers in reduced metal species. Therefore, to determinate the potential at which the faradaic process starts for the metal oxide materials studied, a double potential step chronoamperometry experiment was performed. A potential step was performed from the 1.0 V as initial potential (Pi), at which no faradaic process occurs, to several negative final potentials (Pf). Each step was applied for 59 s.
The chronoamperometric profiles shown in Figure 3 indicate the potential steps associated with each thin-film electrode. For MoO3, when the applied potential returns to Pi, the current in both CO2 and N2 purged is the same. In contrast, after electrode polarization to the initial potential from each final potential with CO2 electrode saturation, Fe2O3 shows an increase in anodic current. This could possibly be related to a reduction product reoxidation. In this sense, it is worth nothing that a wide potential peak at 0.2 V is observed, as can be seen from the cyclic voltammetry results. When a negative potential was applied under N2 and CO2 purging conditions, the currents were not always the same at the end of each step potential. For Fe2O3, an increase in current is observed from −0.3 V after 59 s under CO2 purge conditions, the increase being higher at −0.5 V. For MoO3, said increase in current becomes evident at −0.7 V.
Under CO2 and N2 bubbling conditions, the absolute value of the difference between observed currents as a function of the applied potential step is defined by Equation (9):
j CO 2 max j N 2 max     V = j CO 2 max v j N 2 max v
where |j(CO2)max(V)| and |j(N2)max(V)| are the cathodic currents obtained 59 s after the electrode polarization occurs at the potential Pf.
A current profile calculated using Equation (9) has an increase in the relationship between faradaic and non-faradaic currents, as shown in Figure 4. In each case, the intercept with the potential axis corresponds to the onset potential for faradaic current related with the CO2 reduction. The onset for electrodes used in this study was −0.4 V for Fe2O3, and −0.6 V for MoO3. Furthermore, Figure 4 reveals not only the decrease of the overpotential for CO2 reduction, but also the higher faradaic currents for the Fe2O3 electrode, being the best electrocatalytic electrode under the experimental conditions. It should be noted that differences in electrode interface area may vary and are desirable for proper electrode performance comparison.
On the other hand, TMOs, such as Fe2O3 and MoO3, are n-type semiconductors, and their catalytic activity could be related to the conduction band position [32]. The conduction band position could estimate if the electron transfer is thermodynamically feasible. Although several conduction and valence band position diagrams are found in literature for the thin film electrodes studied, these values correspond to conditions different from those applied in this study, and are generally under vacuum conditions [33]. However, the potential at which the conduction band is found depends on factors such as the type of oxide, charge state of the surface, presence of adsorbed species, pH, and isoelectric point of the material [34]. In this way, the flat band potential was estimated via photoelectrochemical measurements in the presence of low oxidation potential species such as methanol [35]. This potential should be very close to the potential of the conduction band lower edge, and related with the apparition of photocurrents. Figure S5 in Supporting Information shows the LSV profile for Fe2O3 and MoO3 electrodes under transitory illumination, in 0.1 M de TBAF/ACN/+3.0 M CH3OH, purged with N2. The results show that photocurrent onset potentials start at −0.70 V for Fe2O3, and 0.65 V for MoO3. Therefore, the electron transference from the TMO to CO2 molecule and the formation of thermodynamically favored C1 products from CO2 reduction are feasible. Scheme 1 represents the conduction band position for TMO under the conditions of this study, and the potentials (versus SHE at pH 7) for some relevant CO2 redox couples (Scheme 1).
The results show a generalized catalytic behavior for CO2 electrocatalytic reduction over the TMO electrodes used. In all cases, a decrease in the overpotential related to that required for the first CO2 electron transfer is evident. According to the dissociative reduction model, in the presence of CO2, the formation of OV caused by the application of increasingly negative potentials appears to be favored by the stabilization of Fe2+ and Mo5+ species upon CO2 chemisorption. Therefore, it is plausible to think that CO2 could be adsorbed in reduced metal sites, and that the OV generation is favored. Metal reduction is facilitated, causing an increase in capacitive currents in CV profiles, even at potentials where CO2 reduction does not proceed. In this sense, it is reasonable to think that the catalytic surface of the examined TMO materials possesses a different morphology and adsorption energy for CO2, and, therefore, a different activation capability [11,36,37]. Likewise, Tanabe et al. found that basic metal oxides facilitate the transfer of electrons to CO2 molecules adsorbed on the surface by adduct formation [38].
Although the activation of the CO2 molecule by adsorption on the surface of the catalyst favors its reduction, it is not the only factor that determines the electrocatalytic activity of the electrode. The formation of the adsorbed surface does not always lead to CO2 reduction [36]. In this sense, it is important to consider the need of electron-rich active sites giving rise to electron transfer from the TMO surface to the CO2 molecule. Thus, the formation of reduced sites identified as catalytic sites for the reduction of CO2 occurs [21,39]. Density Functional Theory calculations show that OV with n-type semiconductor behavior enhance the surface state near the conduction band [40]. Based on the results obtained, the increment of surface states improves the electrode catalytic activity even at potentials at which CO2 reduction does not occur.
Finally, conduction band position, OV density and their associated energy, as well as surface state density are parameters needed to understand the electrocatalytic behavior of TMO materials. Table 1 presents the initial flat band and faradaic current potentials, Efb and Efi, for the electrocatalytic reduction of CO2.
The initial faradaic onset potential for Fe2O3 electrodes is less negative than those at the the conduction band position. This fact reveals a high density of surface states near the conduction band; from which, electron transfer to the adsorbed CO2 molecules occurs. In the case of MoO3, Efb and Efi are very close to each other. The initial potential for MoO3 is 0.05 V, which is less negative than its flat band potential. This could be related to the fact that the metal Mo has different oxidation numbers; therefore, the excess charge from the formation of OV can be distributed on both surface and bulk atoms, causing the excess charge to delocalize, and the density of electron-rich sites to decrease. The migration capability for OV in TMO such as TiO2 and WO3 is predicted by DFT studies, as well as (0 0 1)-orientation activation energies of 0.76 V and 0.14 V, respectively [41,42]. This means that excess electrons resulting from OV formation in TMO such as MoO3 may migrate to the bulk structure more seamlessly than Fe2O3 [43], which suggests a greater availability of Lewis basic active sites on the TMO surface for CO2 reduction when the OV migration energy is higher.

3. Materials and Methods

The preparation of Fe2O3 and MoO3 electrodes was carried out by chemical bath deposition and the sol-gel method, respectively [28]. Experimental details for oxide thin-film synthesis and the characterization are presented in the Supporting Information.
A computer-controlled micro-Autolab potentiostat was used for all electrochemical experiments. A conventional three-electrode glass cell was used with an Ag/AgCl(s)/KCl (aq,sat) reference electrode and a Pt wire as a counter electrode in all the experiments. The supporting electrolyte was 0.1 M tetrabutylammonium perchlorate (TBAP) (≥99.0% from Sigma-Aldrich, Madrid, Spain) in ACN (99.88% anhydrous from Scharlau, Multisolvent, HPLC grade, Barcelona, Spain). In all the experiments, the electrolyte was purged for 30 min with either N2 (Alphagaz 99.999%, Madrid, Spain) or CO2 (Air Liquid, 99.98%, Madrid, Spain). A conversion of potentials measured to SHE scale is the same as previously reported, and according to Pavlishchuk and Addison. Detailed information about materials, methods, and the potential conversion scale are presented in the Supporting Information.

4. Conclusions

Based on the results obtained, TMOs present an electrocatalytic behavior for the reduction of CO2 in ACN medium. In this sense, it could be affirmed that the location of the conduction band, the ability to generate OV, and charge accumulation in surface states are factors that can modify the charge transfer from TMO to CO2. The results obtained show that, in general terms, the less negative (in potential) the conduction band position is, the more positive the potential at which the processes of faradic reduction of CO2 are initiated. This behavior is expected, since there is a correlation between the location of the edge of the conduction band and the electronic surface states responsible for the CO2 activation. In this sense, future studies that allow obtaining information on the generation energy and the migration of OVs may be relevant for the evaluation of a TMO as a possible (photo)cathode for CO2 reduction. Finally, it should be noted that carbon monoxide was obtained as a reduction product, which is in accordance with what is expected for aprotic systems.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms232113367/s1. References [44,45] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, N.E.M.-R. and C.A.G.-F.; methodology, N.E.M.-R. and A.S.L.-P.; formal analysis, N.E.M.-R. and A.S.L.-P.; investigation, N.E.M.-R.; data curation, A.S.L.-P.; writing—original draft preparation, A.S.L.-P.; writing—review and editing, N.E.M.-R.; supervision, C.A.G.-F.; project administration, C.A.G.-F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by MINCIENCIAS through National Doctoral Scholarship (567-2019), and (848-2019) postdoctoral fellowship.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Nor applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to acknowledge the support of the Universidad Nacional de Colombia, its facilities, and the use of its equipment.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sinsel, S.R.; Riemke, R.L.; Hoffmann, V.H. Challenges and solution technologies for the integration of variable renewable energy sources—A review. Renew. Energy 2020, 145, 2271–2285. [Google Scholar] [CrossRef]
  2. Kondratenko, E.V.; Mul, G.; Baltrusaitis, J.; Larrazábal, G.O.; Pérez-Ramírez, J. Status and perspectives of CO2 conversion into fuels and chemicals by catalytic, photocatalytic and electrocatalytic processes. Energy Environ. Sci. 2013, 6, 3112–3135. [Google Scholar] [CrossRef] [Green Version]
  3. Lv, J.; Yin, R.; Zhou, L.; Li, J.; Kikas, R.; Xu, T.; Wang, Z.; Jin, H.; Wang, X.; Wang, S. Microenvironment Engineering for the Electrocatalytic CO2 Reduction Reaction. Angew. Chem. Int. Ed. 2022, 61, e202207252. [Google Scholar] [CrossRef] [PubMed]
  4. Moura de Salles Pupo, M.; Kortlever, R. Electrolyte Effects on the Electrochemical Reduction of CO2. ChemPhysChem 2019, 20, 2926–2935. [Google Scholar] [CrossRef] [Green Version]
  5. Ait Ahsaine, H.; Zbair, M.; BaQais, A.; Arab, M. CO2 Electroreduction over Metallic Oxide, Carbon-Based, and Molecular Catalysts: A Mini-Review of the Current Advances. Catalysts 2022, 12, 450. [Google Scholar] [CrossRef]
  6. Yang, J.H.; Park, S.J.; Rhee, C.K.; Sohn, Y. Photocatalytic CO2 Reduction and Electrocatalytic H2 Evolution over Pt(0,II,IV)-Loaded Oxidized Ti Sheets. Nanomaterials 2020, 10, 1909. [Google Scholar] [CrossRef]
  7. Zhang, Q.; Li, Y.; Zhu, H.; Suo, B. High-Performance of Electrocatalytic CO2 Reduction on Defective Graphene-Supported Cu4S2 Cluster. Catalysts 2022, 12, 454. [Google Scholar] [CrossRef]
  8. Tackett, B.M.; Gomez, E.; Chen, J.G. Net reduction of CO2 via its thermocatalytic and electrocatalytic transformation reactions in standard and hybrid processes. Nat. Catal. 2019, 2, 381–386. [Google Scholar] [CrossRef]
  9. Xu, S.; Carter, E.A. Theoretical Insights into Heterogeneous (Photo)electrochemical CO2 Reduction. Chem. Rev. 2018, 119, 6631–6669. [Google Scholar] [CrossRef]
  10. Freund, H.J.; Roberts, M. Surface chemistry of carbon dioxide. Surf. Sci. Rep. 1996, 25, 225–273. [Google Scholar] [CrossRef]
  11. Indrakanti, V.P.; Kubicki, J.D.; Schobert, H.H. Photoinduced activation of CO2 on Ti-based heterogeneous catalysts: Current state, chemical physics-based insights and outlook. Energy Environ. Sci. 2009, 2, 745–758. [Google Scholar] [CrossRef]
  12. Lamy, E.; Nadjo, L.; Savéant, J.M. Standard potential and kinetic parameters of the electrochemical reduction of carbon dioxide in dimethylformamide. J. Electroanal. Chem. Interfacial Electrochem. 1977, 78, 403–407. [Google Scholar] [CrossRef]
  13. Koppenol, W. Oxyradical reactions: From bond-dissociation energies to reduction potentials. FEBS Lett. 1990, 264, 165–167. [Google Scholar] [CrossRef] [Green Version]
  14. Shah, S.S.A.; Sufyan Javed, M.; Najam, T.; Molochas, C.; Khan, N.A.; Nazir, M.A.; Xu, M.; Tsiakaras, P.; Bao, S.J. Metal oxides for the electrocatalytic reduction of carbon dioxide: Mechanism of active sites, composites, interface and defect engineering strategies. Coord. Chem. Rev. 2022, 471, 214716. [Google Scholar] [CrossRef]
  15. Rasko, J.; Solymosi, F. Infrared Spectroscopic Study of the Photoinduced Activation of CO2 on TiO2 and Rh/TiO2 Catalysts. J. Phys. Chem. 1994, 98, 7147–7152. [Google Scholar] [CrossRef]
  16. Perissinotti, L.L.; Brusa, M.A.; Grela, M.A. Yield of Carboxyl Anion Radicals in the Photocatalytic Degradation of Formate over TiO2 Particles. Langmuir 2001, 17, 8422–8427. [Google Scholar] [CrossRef]
  17. Habisreutinger, S.N.; Schmidt-Mende, L.; Stolarczyk, J.K. Photocatalytic Reduction of CO2 on TiO2 and Other Semiconductors. Angew. Chem. Int. Ed. 2013, 52, 7372–7408. [Google Scholar] [CrossRef] [PubMed]
  18. Ludwig, T.; Singh, A.R.; Nørskov, J.K. Acetonitrile Transition Metal Interfaces from First Prin-ciples. J. Phys. Chem. Lett. 2020, 11, 9802–9811. [Google Scholar] [CrossRef]
  19. Gomes, R.J.; Birch, C.; Cencer, M.M.; Li, C.; Son, S.B.; Bloom, I.D.; Assary, R.S.; Amanchukwu, C.V. Probing Electrolyte Influence on CO2 Reduction in Aprotic Solvents. J. Phys. Chem. C 2022, 126, 13595–13606. [Google Scholar] [CrossRef]
  20. Kuntyi, R.; Zozulya, G.; Shepida, M. CO2 Electroreduction in Organic Aprotic Solvents: A Mini Review. J. Chem. 2022, 2022, 1306688. [Google Scholar] [CrossRef]
  21. Bandi, A. Electrochemical Reduction of Carbon Dioxide on Conductive Metallic Oxides. J. Electrochem. Soc. 1990, 137, 2157–2160. [Google Scholar] [CrossRef]
  22. Chu, S.; Yan, X.; Choi, C.; Hong, S.; Robertson, A.W.; Masa, J.; Han, B.; Jung, Y.; Sun, Z. Stabilization of Cu+ by tuning a CuO–CeO2 interface for selective electrochemical CO2 reduction to ethylene. Green Chem. 2020, 22, 6540–6546. [Google Scholar] [CrossRef]
  23. Lum, Y.; Ager, J.W. Stability of Residual Oxides in Oxide-Derived Copper Catalysts for Electrochemical CO2 Reduction Investigated with 18O Labeling. Angew. Chem. Int. Ed. 2017, 57, 551–554. [Google Scholar] [CrossRef]
  24. Zou, Y.; Wang, S. An Investigation of Active Sites for electrochemical CO2 Reduction Reactions: From In Situ Characterization to Rational Design. Adv. Sci. 2021, 8, 2003579. [Google Scholar] [CrossRef]
  25. Figueiredo, M.C.; Ledezma-Yanez, I.; Koper, M.T.M. In Situ Spectroscopic Study of CO2 Electroreduction at Copper Electrodes in Acetonitrile. ACS Catal. 2016, 6, 2382–2392. [Google Scholar] [CrossRef]
  26. Rudnev, A.V.; Zhumaev, U.E.; Kuzume, A.; Vesztergom, S.; Furrer, J.; Broekmann, P.; Wandlowski, T. (The promoting effect of water on the electroreduction of CO2 in acetonitrile. Electrochim. Acta 2016, 189, 38–44. [Google Scholar] [CrossRef]
  27. de Castro, I.A.; Datta, R.S.; Ou, J.Z.; Castellanos-Gomez, A.; Sriram, S.; Daeneke, T.; Kalantar-Zadeh, K. Molybdenum Oxides—From Fundamentals to Functionality. Adv. Mater. 2017, 29, 1701619. [Google Scholar] [CrossRef]
  28. Mendieta-Reyes, N.E.; Díaz-García, A.K.; Gómez, R. Simultaneous Electrocatalytic CO2 Reduction and Enhanced Electrochromic Effect at WO3 Nanostructured Electrodes in Acetonitrile. ACS Catal. 2018, 8, 1903–1912. [Google Scholar] [CrossRef] [Green Version]
  29. Masetti, E.; Dini, D.; Decker, F. The electrochromic response of tungsten bronzes MxWO3 with different ions and insertion rates. Sol. Energy Mater. Sol. Cells 1995, 39, 301–307. [Google Scholar] [CrossRef]
  30. Foley, J.K.; Korzeniewski, C.; Pons, S. Anodic and cathodic reactions in acetonitrile/tetra-n-butylammonium tetrafluoroborate: An electrochemical and infrared spectroelectrochemical study. Can. J. Chem. 1988, 66, 201–206. [Google Scholar] [CrossRef]
  31. Hepel, M.; Wickham, D. Large Cation Model of Dissociative Reduction of WO3−x Lattice Studied by EQCN and AFM. ECS Trans. 2009, 19, 11–23. [Google Scholar] [CrossRef]
  32. Chava, R.K.; Son, N.; Kang, M. Band structure alignment transitioning strategy for the fabrication of efficient photocatalysts for solar fuel generation and environmental remediation applications. J. Colloid Interface Sci. 2022, 627, 247–260. [Google Scholar] [CrossRef]
  33. Jian, J.; Sun, J. A Review of Recent Progress on Silicon Carbide for Photoelectrochemical Water Splitting. Sol. RRL 2020, 4, 2000111. [Google Scholar] [CrossRef] [Green Version]
  34. Wang, D.; Liu, Z.P.; Yang, W.M. Proton-Promoted Electron Transfer in Photocatalysis: Key Step for Photocatalytic Hydrogen Evolution on Metal/Titania Composites. ACS Catal. 2017, 7, 2744–2752. [Google Scholar] [CrossRef]
  35. Ismael, M. Highly effective ruthenium-doped TiO2 nanoparticles photocatalyst for visible-light-driven photocatalytic hydrogen production. New J. Chem. 2019, 43, 9596–9605. [Google Scholar] [CrossRef]
  36. Breyer, C.; Reichert, D.; Seidel, J.; Hüttl, R.; Mertens, F.; Kureti, S. Kinetic modeling of the adsorption and desorption of CO2 on α-Fe2O3. Phys. Chem. Chem. Phys. 2015, 17, 27011–27018. [Google Scholar] [CrossRef]
  37. Liu, L.; Lin, M.; Liu, Z.; Sun, H.; Zhao, X. Density functional theory study of CO2 and H2O adsorption on a monoclinic WO3(001) surface. Chem. Res. Chin. Univ. 2017, 33, 255–260. [Google Scholar] [CrossRef]
  38. Wicke, E.J.R., Anderson, M., Boudart. (Eds.): Catalysis, Science and Technology. Vol. 4, Springer-Verlag, Berlin, Heidelberg, New York 1983. 289 Seiten, Preis: DM 132,-. Ber. Bunsenges. Phys. Chem. 1984, 88, 586. [CrossRef]
  39. Rao, C.N.R. Transition Metal Oxides. Annu. Rev. Phys. Chem. 1989, 40, 291–326. [Google Scholar] [CrossRef]
  40. Di Valentin, C.; Pacchioni, G.; Selloni, A. Reduced and n-Type Doped TiO2: Nature of Ti3+ Species. J. Phys. Chem. C 2009, 113, 20543–20552. [Google Scholar] [CrossRef]
  41. Cheng, H.; Selloni, A. Energetics and diffusion of intrinsic surface and subsurface defects on anatase TiO2 (101). J. Chem. Phys. 2009, 131, 054703. [Google Scholar] [CrossRef] [PubMed]
  42. Le, H.M.; Vu, N.H.; Phan, B.T. Migrations of oxygen vacancy in tungsten oxide (WO3): A density functional theory study. Comput. Mater. Sci. 2014, 90, 171–176. [Google Scholar] [CrossRef]
  43. Yin, W.J.; Wen, B.; Zhou, C.; Selloni, A.; Liu, L.M. Excess electrons in reduced rutile and anatase TiO2. Surf. Sci. Rep. 2018, 73, 58–82. [Google Scholar] [CrossRef] [Green Version]
  44. Hasani-Sadrabadi, M.M.; Dashtimoghadam, E.; Nasseri, R.; Karkhaneh, A.; Majedi, F.S.; Mokarram, N.; Renaud, P.; Jacob, K.I. Cellulose nanowhiskers to regulate the microstructure of perfluorosulfonate ionomers for high-performance fuel cells. J. Mater. Chem. A 2014, 2, 11334. [Google Scholar] [CrossRef]
  45. Pavlishchuk, V.V.; Addison, A.W. Conversion constants for redox potentials measured versus different reference electrodes in acetonitrile solutions at 25 °C. Inorg. Chim. Acta 2000, 298, 97–102. [Google Scholar] [CrossRef]
Figure 1. Cyclic voltammograms for CO2 reduction over thin-film electrodes: (a) Fe2O3 and (b) MoO3 in 0.1 M TBAF/ACN electrolyte solutions; black and red plots correspond to CO2 and N2 purged electrolytes, respectively. (ss) corresponds to metal atoms in surface states.
Figure 1. Cyclic voltammograms for CO2 reduction over thin-film electrodes: (a) Fe2O3 and (b) MoO3 in 0.1 M TBAF/ACN electrolyte solutions; black and red plots correspond to CO2 and N2 purged electrolytes, respectively. (ss) corresponds to metal atoms in surface states.
Ijms 23 13367 g001
Figure 2. Linear sweep voltammograms taken in 0.1 M TBAF/ACN with TMO thin-film electrodes at 5 mV·s−1 scanning rate: (a) Fe2O3, (b) MoO3. The black and red curves correspond to scans under an atmosphere of N2 and CO2, respectively.
Figure 2. Linear sweep voltammograms taken in 0.1 M TBAF/ACN with TMO thin-film electrodes at 5 mV·s−1 scanning rate: (a) Fe2O3, (b) MoO3. The black and red curves correspond to scans under an atmosphere of N2 and CO2, respectively.
Ijms 23 13367 g002
Figure 3. Double potential step chronoamperometry results for the electrocatalytic reduction of CO2 in 0.1 M TBAF/ACN electrolyte solution with thin-film electrodes: (a) Fe2O3 and (b) MoO3. The red and black graphs correspond to the electrolyte purged with N2 and CO2, individually. Blue arrow corresponds to initial potential (Pi). Arrows indicate potential steps.
Figure 3. Double potential step chronoamperometry results for the electrocatalytic reduction of CO2 in 0.1 M TBAF/ACN electrolyte solution with thin-film electrodes: (a) Fe2O3 and (b) MoO3. The red and black graphs correspond to the electrolyte purged with N2 and CO2, individually. Blue arrow corresponds to initial potential (Pi). Arrows indicate potential steps.
Ijms 23 13367 g003
Figure 4. Profile of current associated with CO2 reduction faradaic processes for examined TMO materials as a function of applied potential in square-wave chronoamperometric experiments.
Figure 4. Profile of current associated with CO2 reduction faradaic processes for examined TMO materials as a function of applied potential in square-wave chronoamperometric experiments.
Ijms 23 13367 g004
Scheme 1. Representation of the conduction band position measured by flat band potential determination and some CO2 redox couples related with the onset of the faradic currents.
Scheme 1. Representation of the conduction band position measured by flat band potential determination and some CO2 redox couples related with the onset of the faradic currents.
Ijms 23 13367 sch001
Table 1. Initial and flat band potential for TMO electrodes. Potentials were measured versus an Ag/AgCl/KClsat reference electrode.
Table 1. Initial and flat band potential for TMO electrodes. Potentials were measured versus an Ag/AgCl/KClsat reference electrode.
ElectrodeE/V (Efb)E/V (Efi)
Fe2O3−0.70−0.4
WO3−0.50−0.6
MoO3−0.60−0.6
TiO2−1.55−1.3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mendieta-Reyes, N.E.; Lozano-Pérez, A.S.; Guerrero-Fajardo, C.A. Insights of Fe2O3 and MoO3 Electrodes for Electrocatalytic CO2 Reduction in Aprotic Media. Int. J. Mol. Sci. 2022, 23, 13367. https://doi.org/10.3390/ijms232113367

AMA Style

Mendieta-Reyes NE, Lozano-Pérez AS, Guerrero-Fajardo CA. Insights of Fe2O3 and MoO3 Electrodes for Electrocatalytic CO2 Reduction in Aprotic Media. International Journal of Molecular Sciences. 2022; 23(21):13367. https://doi.org/10.3390/ijms232113367

Chicago/Turabian Style

Mendieta-Reyes, Néstor E., Alejandra S. Lozano-Pérez, and Carlos A. Guerrero-Fajardo. 2022. "Insights of Fe2O3 and MoO3 Electrodes for Electrocatalytic CO2 Reduction in Aprotic Media" International Journal of Molecular Sciences 23, no. 21: 13367. https://doi.org/10.3390/ijms232113367

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