Next Article in Journal
Synthesis of Natural-Like Snow by Ultrasonic Nebulization of Water: Morphology and Raman Characterization
Previous Article in Journal
Biochemical Characterization of a Flavonoid O-methyltransferase from Perilla Leaves and Its Application in 7-Methoxyflavonoid Production
Previous Article in Special Issue
Electrochemical Reduction of CO2 to Formate on Easily Prepared Carbon-Supported Bi Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Continuous Electrochemical Reduction of CO2 to Formate: Comparative Study of the Influence of the Electrode Configuration with Sn and Bi-Based Electrocatalysts

by
Guillermo Díaz-Sainz
*,
Manuel Alvarez-Guerra
and
Angel Irabien
Chemical and Biomolecular Engineering Department, University of Cantabria, ETSIIT, Avda. Los Castros s/n, 39005 Santander, Spain
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(19), 4457; https://doi.org/10.3390/molecules25194457
Submission received: 27 August 2020 / Revised: 23 September 2020 / Accepted: 25 September 2020 / Published: 28 September 2020
(This article belongs to the Special Issue Nanocatalysts for Electrochemical Reduction of CO2)

Abstract

:
Climate change has become one of the most important challenges in the 21st century, and the electroreduction of CO2 to value-added products has gained increasing importance in recent years. In this context, formic acid or formate are interesting products because they could be used as raw materials in several industries as well as promising fuels in fuel cells. Despite the great number of studies published in the field of the electrocatalytic reduction of CO2 to formic acid/formate working with electrocatalysts of different nature and electrode configurations, few of them are focused on the comparison of different electrocatalyst materials and electrode configurations. Therefore, this work aims at presenting a rigorous and comprehensive comparative assessment of different experimental data previously published after many years of research in different working electrode configurations and electrocatalysts in a continuous mode with a single pass of the inputs through the reactor. Thus, the behavior of the CO2 electroreduction to formate is compared operating with Sn and Bi-based materials under Gas Diffusion Electrodes (GDEs) and Catalyst Coated Membrane Electrodes (CCMEs) configurations. Considering the same electrocatalyst, the use of CCMEs improves the performance in terms of formate concentration and energy consumption. Nevertheless, higher formate rates can be achieved with GDEs because they allow operation at higher current densities of up to 300 mA·cm−2. Bi-based-GDEs outperformed Sn-GDEs in all the figures of merit considered. The comparison also highlights that in CCME configuration, the employ of Bi-based-electrodes enhanced the behavior of the process, increasing the formate concentration by 35% and the Faradaic efficiency by 11%.

Graphical Abstract

1. Introduction

Carbon dioxide (CO2) emissions to the atmosphere have gradually increased since the middle of the twentieth century from 326 ppm in 1971 to approximately 417 ppm in 2020 [1]. According to the Sustainable Development Goals (SDGs) established by the United Nations and particularly, to SDG Number 13 defined as Climate Action, CO2 is the most important human-emitted greenhouse gas in the atmosphere. Therefore, shrinking the CO2 emissions should be considered as one of the most important priorities in the current century to mitigate climate change [2].
Different strategies can be considered to reduce CO2 emissions to the atmosphere, such as improving the energy efficiency or developing renewable energy sources and related technologies [3,4]. In this sense, Carbon Capture, Storage and Utilization (CCSU) [5,6,7,8,9] have been suggested as promising approaches for mitigating climate change. In particular, the possibility of converting captured CO2 into fuels and useful industrial chemicals has been especially pointed out [10,11].
Indeed, CO2 can be transformed to value-added chemical products by several routes, such as thermochemical processes or mineralization [12], biological transformation [13], electrochemical [14] or photochemical/photoelectrochemical conversion [15]. Among these approaches, the electrochemical reduction of CO2 toward chemicals with value added has been suggested as an excellent way to store energy from renewable sources in the form of chemical products [16,17].
In this respect, different valued-added products [18,19] can be obtained by CO2 electrocatalytic reduction such as carbon monoxide (CO) [20,21], formic acid (HCOOH) or formate (HCOO) [22,23] or methanol (CH3OH), ethylene (C2H4), and methane (CH4) [24,25,26,27,28].
Among these products, HCOOH or HCOO, depending on pH, are used as raw material in several industries (leather tanning, animal feed, steel pickling, or pharmaceutical) [29]. Furthermore, this product has been recommended as an interesting fuel for low-temperature fuel cells [30,31], as well as a promising hydrogen carrier [32,33]. According to the literature [22,34,35], electrocatalysts of different nature, different electrode configurations, and different electrochemical reactors have been used for studying the electrocatalytic reduction of CO2 to HCOOH/HCOO. On the one hand, copper (Cu) [36], cobalt (Co) [37], molybdenum (Mo) [38], lead (Pb) [39,40,41], indium (In) [42,43,44], palladium (Pd) [45,46], and especially tin (Sn) [47,48,49,50,51,52,53,54] and bismuth (Bi) [55,56,57,58,59,60,61,62] are the most common catalysts investigated for the selective electrochemical reduction of CO2 to HCOOH/HCOO. On the other hand, the electrocatalyst can be used in different electrode configurations, such as in the form of a metal plate, a Gas Diffusion Electrode (GDE), or a Catalyst Coated Membrane Electrode (CCME), operating with different electrochemical reactor configurations and operating conditions.
Moreover, an important problem in CO2 electroreduction is that many different variables may have an influence on the results, which makes it very difficult to compare the results of studies carried out in different experimental conditions. Although there is a great number of studies that focused on the CO2 electroreduction to obtain HCOOH/HCOO using electrocatalysts of diverse nature or cathode configuration, few of them focused on thorough comparisons of the performance of the electrochemical process operating with different catalyst materials and electrode configurations with the same experimental setup and operating conditions [63,64,65]. In this context, this work aims to rigorously compare the behavior of the CO2 electroreduction to obtain HCOO using the same experimental setup but employing electrocatalysts of a different nature, (i) Sn carbon-supported nanoparticles (Sn/C NPs) and (ii) Bi carbon-supported nanoparticles (Bi/C NPs), and under different electrode configurations: (i) GDEs and (ii) CCMEs. The experimental data needed to achieve the aim of manuscript have been previously published after many years of research of our group. A comparative assessment is performed in terms of different relevant figures of merit, including HCOO concentration, HCOO rate, Faradaic efficiency for HCOO (FE), and consumption of energy per kmol of HCOO.

2. Comparative Study of the Electrode Configurations: GDE–CCME

This section will compare the performance of GDE and CCME electrode configuration for the electroreduction of CO2 to HCOO, employing the same experimental setup and the same electrocatalyst. The comparison will be made first considering the same Sn-based catalyst and then using the same Bi-based catalyst.

2.1. Sn-Based Electrodes

The results of different experiments carried out with the same catalyst (Sn NPs) using different working electrode configurations—GDEs (Sn/C-GDEs) and CCMEs (Sn/C-CCMEs)—are reported and discussed in references [66,67], respectively. A summary of the main results of the performance with both configurations is summarized in Table 1. Although the operating conditions were the same, depositing directly the catalytic material over the membrane (Sn/C-CCME) instead of depositing the catalyst over a carbonaceous support (Sn/C-GDE) improved the performance of the electrochemical process, as shown in Table 1. Noteworthy results in terms of HCOO concentration and energy consumption per kmol of HCOO using the Sn-based catalyst can be obtained using Sn/C-CCME as a cathode in the electrochemical reactor, with values of 19.2 g·L−1 and 244 kWh·kmol−1, respectively, which represent noticeably better results than those with the Sn/C-GDE configuration. This improvement can be in part attributed to the better behavior of the electrochemical reactor using a solid polymer electrolyte (with the Sn/C-CCME configuration) instead of a liquid electrolyte (with the Sn/C-GDE configuration). This may be attributed to the fact that the humidified CO2 gas input stream enhances the delivery of gaseous CO2 to the working electrode surface, and therefore, the solubility limitation of CO2 in the electrolyte.
Despite the improvement in the HCOO concentration and the energy consumption per kmol of HCOO, the HCOO rate and the FE for HCOO did not show a substantial enhancement. As can be seen in Table 1, the FE for HCOO was very similar in both configurations (42.3 and 49.4%), and the HCOO rate obtained using the Sn/C-CCME configuration decreased from 4.38 to 1.15 mmol·m−2·s−1 because of the low values of current density supplied to the electrochemical reactor using the Sn/C-CCME cathode (45 mA·cm−2) configuration with respect to Sn/C-GDE configuration (90, 150, and 200 mA·cm−2). In this sense, the current density employed in Sn/C-CCME was fixed to that value, since an increase in this variable results in higher cell potentials with respect to the use of Sn/C-GDE at the same operating conditions.
Thus, it can be concluded that an Sn/C-GDE configuration is more suitable for operating at higher current densities, and therefore achieving higher rates of CO2 reduction and formation of product, but the process results are more energetically favorable under an Sn/C-CCME configuration.

2.2. Bi-Based Electrodes

In Table 2, the comparison between the GDE configuration (Bi/C-GDE) [68] and the CCME configuration (Bi/C-CCME) [69], employing the same Bi/C NPs as electrocatalyst, is carried out, in terms of HCOO concentration, HCOO rate, FE for HCOO, and the energy consumption per kmol of HCOO.
It is noteworthy that Bi/C-GDEs were able to operate with a high current density up to 300 mA·cm−2, obtaining an HCOO concentration of 5.2 g·L−1 with an FE for HCOO of 70.6%, an HCOO rate of 10.97 mmol·m−2·s−1, and an energy consumption per kmol of HCOO of 410 kWh·kmol−1 of HCOO. Nevertheless, the possibility of obtaining higher HCOO concentration (18.0 g·L−1) with Bi/C-GDEs was by means of lowering the catholyte flow and at the expense of a decrease in both FE for HCOO (45.1%) and HCOO rate (4.67 mmol·m−2·s−1). As illustrated in Table 2, working with a Bi/C-CCME configuration, promising results can be achieved in terms of HCOO concentration, FE for HCOO, and energy consumption per kmol of HCOO. The best result using a Bi/C-CCME configuration allowed obtaining an HCOO concentration approximately 44% higher with respect to the highest HCOO concentration obtained using Bi/C-GDE, and simultaneously, the FE for to HCOO is approximately 21.5% higher with an important saving in the energy consumption per kmol of HCOO of around 50%. Similarly, as was found for Sn-based electrodes, the HCOO rate decreased using a Bi/C-CCMEs configuration with respect to the Bi/C-GDE configuration because of the lower current densities supplied to the electrochemical filter press by potentiostat–galvanostat.
Finally, as can be concluded from the analysis of the results in Table 2, the Bi/C-CCME configuration improves the performance of the electrochemical conversion of CO2 to HCOO in terms of HCOO concentration, FE for HCOO, and the energy consumption per kmol of HCOO. However, the HCOO rate did not show an important improvement for the electrocatalytic reduction of CO2 to HCOO because the CCME configuration did not allow operating with high current densities values.

3. Comparative Study of the Catalyst Nature

This section is focused on the comparison of the performance of the CO2 electroreduction to obtain HCOO employing two electrocatalysts of different nature: Sn carbon-supported nanoparticles (Sn/C NPs) and Bi carbon-supported nanoparticles (Bi/C NPs), considering the same experimental setup and the same electrode configuration. The comparison of the two catalysts will be made first using a GDE configuration and then employing a CCME configuration.

3.1. Sn vs. Bi in GDE Configuration

A comparative study of the behavior of Sn/C NPs and Bi/C NPs in the form of a GDE is performed. For a clearer visualization, the results obtained using the GDE configurations with Sn/C NPs and Bi/C NPs are summarized in Figure 1 and Figure 2.
As explained in Section 2, the current density and the catholyte flow per geometric area are considered the most influential variables in the performance of the process for the electrochemical conversion of CO2 to HCOO with GDEs The influence of current density in FE for HCOO (Figure 1a) and the HCOO rate (Figure 1b) with both electrocatalytic materials can be clearly studied in Figure 1. The current density ranged between 90 and 300 mA·cm−2 operating with the same catholyte flow (5.7 mL·min−1) to allow a fair comparison. Nevertheless, it was not feasible to supply current densities higher than 200 mA·cm−2 using Sn/C-GDEs because of the high cell potential obtained operating with this current density (>4 V). Despite the decrease in the FE for HCOO, from 69 to 55% (with Sn/C-GDEs as working electrode) and from 92 to 80% (with Bi/C-GDEs as working electrode), when the current increased from 90 to 200 mA·cm−2, an increase in the HCOO rate was observed. Moreover, it is important to point out that a noteworthy result in terms of FE for HCOO, of approximately 70%, was achieved using Bi/C-GDEs supplied with a current density of 300 mA·cm−2.
As illustrated in Figure 1b, on the one hand, the HCOO rate using Sn/C-GDEs increased approximately 70% (from 3.23 to 5.61 mmol·m−2·s−1) when the current density increased from 90 to 200 mA·cm−2. In contrast, HCOO rates up to 8.33 mmol·m−2·s−1 were obtained using Bi/C-GDEs, which was approximately 50% higher than those with Sn/C-GDE under the same operating conditions and with a current density of 200 mA·cm−2. Nevertheless, the highest HCOO rate was obtained with a current density of 300 mA·cm−2 (with Bi/C-GDEs as working electrode), achieving an excellent value of 10.97 mmol·m−2·s−1.
A summary of the results in terms of HCOO concentration obtained at different current densities and catholyte flows is represented in Figure 2. On the one hand, operating at a certain same current density (90 or 200 mA·cm−2), a higher HCOO concentration collected in the output stream of the electrochemical filter press cell is obtained if the catholyte flow is lowered from 5.7 to 0.7 mL·min−1·cm−2. As happened with the FE and the HCOO rate, the performance of the electrochemical cell in terms of HCOO concentration is better working with Bi/C-GDEs than with Sn/C-GDEs. The highest HCOO concentration obtained was 18.0 g·L−1, which was 7% higher than operating with the same value of current density and catholyte flow in Sn/C-GDEs (Figure 2). However, working with a current density = 200 mA·cm−2, the highest difference between the Bi/C-GDEs and the Sn/C-GDEs was operating with a catholyte flow = 5.7 mL·min−1, obtaining an HCOO concentration of 3.95 g·L−1 (using Bi/C-GDE as cathode), which was 50% higher with respect to the value of HCOO concentration using Sn/C-GDE.
In addition, the comparison of Bi and Sn-based GDEs in terms of energy consumption per kmol of HCOO is summarized in Table 3. Bi-based electrocatalysts in the form of GDEs resulted in needing less energy consumption per kmol of HCOO at different current densities and catholyte flows compared with Sn/C GDEs.
After the analysis of these results, it can be concluded that the use of GDEs with Bi/C NPs gives better results for the electrocatalytic reduction of CO2 to HCOO in terms of all figures of merit studied (FE for HCOO, HCOO rate, HCOO concentration and energy consumption per kmol of HCOO). Moreover, it should be highlighted that Bi/C-GDEs allow operating with remarkable values of current density (300 mA·cm−2). In the next section, a similar comparison was carried out between Bi/C NPs and Sn/C NPs in the form of CCMEs.

3.2. Sn vs. Bi in CCME Configuration

As previously discussed in Section 2, the key variables studied in the CCMEs configuration are the temperature and the water flow in the input CO2 stream [67,69]. In order to compare the performance of Sn/C-CCMEs and Bi/C-CCMEs, all the experiments considered for the study in this section were carried out operating with a catalyst load = 0.75 mg·cm−2 and a current density = 45 mA·cm−2.
First, a comparative of the performance of Sn/C-CCMEs and Bi/C-CCMEs at different temperatures is carried out in terms of Faradaic efficiency for HCOO (Figure 3a), HCOO rate (Figure 3b), energy consumption per kmol of HCOO (Figure 3c), and HCOO concentration (Figure 3d). The range of temperature studied in both configurations was between 20 and 50 °C. Using either Bi or Sn as electrocatalyst, when the temperature was increased from 20 to 50 °C, all the figures of merit analyzed in this work got worse due to the promotion of the hydrogen evolution reaction. The best results were obtained using Bi-based electrodes operating at ambient condition of temperature, whereas when the temperature was increased to 50 °C, the best performance of the electrochemical process was with the use of Sn/C NPs instead of Bi/C NPs.
Supplying the CO2 stream to the electrochemical reactor with Bi-based CCMEs at a temperature = 20 °C, the FE for HCOO, the HCOO rate, the energy consumption per kmol of HCOO, and the HCOO concentration obtained were 47.2%, 1.10 mmol·m−2·s−1, 312.1 kWh·kmol−1, and 22.3 g·L−1, respectively. Nevertheless, at the same conditions and using Sn/C-CCMEs, the figures of merit obtained were 47.3%, 1.10 mmol·m−2·s−1, 226.6 kWh·kmol−1, and 18.4 g·L−1, respectively. Therefore, although the FE and the HCOO rate were very similar, the employ of Bi as electrocatalyst gives the best result in terms of HCOO concentration (21% higher with Bi), and the use of Sn gives better results in term of energy consumption per kmol of HCOO (27% lower consumption than with Bi). It is interesting to reiterate that raising the temperature from 20 to 50 °C resulted in worse results in terms of all the figures of merit, as illustrated in Figure 3.
The influence of the water flow in the CO2 stream was compared operating at ambient condition of temperature (20 °C) in the CO2 input stream to the electrochemical reactor, as summarized graphically in Figure 4. The range of the water flow in the CO2 stream studied using both catalyst materials ranged from 0.15 to 1 g·h−1. It can be seen in Figure 4a,b that the results of FE and rate obtained using Sn-based electrodes and Bi-based electrodes were very similar in the range of water flow studied. However, the performance of the electrochemical filter press cell was optimized with a water flow in the CO2 stream = 0.5 g·h−1. On the one hand, using Bi in the form of CCMEs, a FE, a HCOO rate, an energy consumption, and a HCOO concentration of 54.8%, 1.28 mmol·m−2·s−1, 265.8 kWh·kmol−1, and 25.9 g·L−1, were obtained, respectively. Nevertheless, using Sn/C-CCMEs with the same operating conditions, the HCOO concentration lowered 35 points in percentage to 19.2 g·L−1, keeping similar values in terms of FE, HCOO rate, and energy consumption (49.4%, 1.15 mmol·m−2·s−1, and 244 kWh·kmol−1, respectively).
In both kinds of electrodes (Sn-based CCMEs and Bi-based CCMEs), raising the water flow in the CO2 input stream from 0.5 to 1 g·h−1, the behavior of the electrochemical process to give HCOO in terms of all figures of merit analyzed got worse. On the one hand, the FE and the HCOO rate (Figure 4a,b) lowered from 54.8% to 47.8% and from 1.28 to 1.11 mmol·m−2·s−1, respectively, with the use of Bi-based electrodes, and at the same conditions under the Sn-CCME configuration, the FE decreased from 49.4% to 46.3% and the HCOO rate decreased from 1.15 to 1.08 mmol·m−2·s−1, respectively. On the other hand, as illustrated in Figure 4c,d, the HCOO concentration decreased from 25.9 to 22.6 g·L−1 and from 19.2 to 17.5 g·L−1, whereas the energy consumption per kmol of HCOO increased to 319.9 and 249 kWh·kmol−1 using Bi-based and Sn-based electrodes, respectively.
Furthermore, the results show that the use of Bi-based materials in the form of CCMEs improves the performance of the electrochemical process for the electrocatalytic reduction of CO2 to HCOO in terms of FE for HCOO, HCOO rate, and HCOO concentration. However, the employ of Sn/C NPs in the form of CCMEs working at the same operating conditions allows obtaining HCOO with less energy consumption per kmol of HCOO.
Finally, further research is still required to overcome current limitations and develop processes with performances that simultaneously optimize all the figures of merit analyzed in this study. In this regard, research efforts should focus on the development of new filter press configurations [70,71] such as the use of a three-compartment electrochemical reactor, bipolar membranes [72], or working electrode configurations [73] to enhance the mass transfer of the reagents and products in the counter and working electrode, and simultaneously address the synthesis of innovative electrocatalysts for both cathode [74] and anode [75] to reduce the energy consumption in both compartments of the electrochemical reactor.

4. Experimental Conditions

All the experiments considered in this comparative analysis were carried out working in a continuous mode with a single pass of the input streams through the electrochemical filter press reactor. The experimental setup included an electrochemical filter press reactor as the core element, a potentiostat–galvanostat, tanks, peristaltic pumps, and a Vapor Delivery Module for operating in a gaseous phase at the cathode side under the CCME configuration. In order to carry out a rigorous comparative study of the electrode configurations, the same electrocatalysts for both configurations were employed: (i) Sn/C-NPs or (ii) Bi/C-NPs. More detail about the synthesis and the characterization of Sn/C NPs and Bi/C-NPs are described in detail in Castillo et al. 2017 [66] and Ávila-Bolívar et al. 2019 [76], respectively.
In contrast, the behavior of the two different electrocatalysts (Sn/C NPs and Bi/C NPs) was compared under the same configuration and experimental setup in both studies: (i) GDEs (Figure 5a), which are described as detailed in Castillo et al. 2017 [66] and Díaz-Sainz et al. 2019 [68], and (ii) CCMEs (Figure 5b), whose fabrication and characterization are described in Díaz-Sainz et al. 2018 [67] and Díaz-Sainz et al. 2020 [69]. In this context, the size of both quasi-spherical Sn/C and Bi/C NPs are 10–15 nm and 9.3 ± 1.6 nm, respectively. The thickness values of the different layers shown in Figure 5 in each of the electrodes compared in this work are summarized in Table 4. In addition, the same operating conditions were employed in the references previously mentioned, as reviewed in Table 5.
In the GDE configuration, the current density and the liquid electrolyte flow per geometric area are considered key variables in the electrochemical conversion of CO2 to HCOO, while in the CCMEs configuration, the key variables studied will be the temperature and the water flow in the input CO2 stream. It is important to note that in the CCME configuration, due to the characteristics of CCME and unlike in GDEs, it is not feasible to work at high values of current density because this implies huge increases in cell potentials. This is why the comparison in Section 3.2. will be carried out using data at a fixed current density of 45 mA·cm−2. In contrast, the temperature is only considered as a key variable for CCMEs because its influence in a GDE configuration is much more limited, but operating in a CCME configuration for the gas-phase electrocatalytic reduction of CO2 to HCOO, this variable has an important influence in the amount of water vapor condensed over the CCME surface as well as the water flow in the CO2 input stream, which is a crucial aspect in the performance of the process.
Finally, a Dimensionally Stable Anode (DSA) (number 1, Figure 5), a leak-free Ag/AgCl, and a Nafion 117 membrane (number 2, Figure 5) were used as a counter electrode, as a reference electrode, and as a cationic exchange membrane. In addition, the concentration of the HCOO produced by the electrocatalytic reduction of CO2 was analyzed by ion-chromatography technique.

5. Conclusions

This work is a comprehensive comparative assessment of different experimental data previously published after many years of research for the electrocatalytic reduction of CO2 to HCOO in different working electrode configurations and electrocatalysts in a continuous mode with a single pass of the inputs through the reactor.
First, the comparison focused on operation with the same electrocatalyst: Sn/C NPs and Bi/C NPs, which were compared in different kinds of working electrode configurations. Considering the same electrocatalyst, the use of CCMEs improves the performance in terms of HCOO concentration, Faradaic efficiency, and energy consumption when compared with GDEs of that same electrocatalyst. However, the HCOO rate worsened because of the low values of current densities that had to be supplied to the electrochemical reactor in the CCME configuration.
Moreover, considering the same operating conditions, a rigorous comparison of both electrocatalysts in the form of GDEs and CCMEs has been carried out. Firstly, using Bi/C-GDEs, the performance of the electrochemical reactor was improved in all the figures of merit analyzed (FE for HCOO, HCOO rate, energy consumption per kmol of HCOO, and HCOO concentration) with respect to the Sn/C-GDEs operating at the same conditions. The same comparative study was performed using CCMEs as a cathode configuration. In this new scenario, the use of Bi/C NPs improved the HCOO concentration in 35%, the FE for HCOO in 11%, and the HCOO rate in 11% with respect to the employ of Sn carbon-supported nanoparticles. Nevertheless, the energy consumption per kmol of HCOO worsened with the use of Bi-based electrodes in 9%.
Finally, despite notable advances achieved, before the electrochemical conversion of CO2 to HCOO could be an industrial reality, further research is still required to optimize all the figures of merit analyzed in this study.

Author Contributions

Conceptualization, M.A.-G. and A.I.; Methodology, G.D.-S. and M.A.-G.; Investigation, G.D.-S.; Data Curation, G.D.-S. and M.A.-G.; Writing—Original Draft Preparation, G.D.-S. and M.A.-G.; Writing—Review and Editing, G.D.-S., M.A.-G. and A.I.; Visualization, G.D.-S. and M.A.-G.; Supervision, M.A.-G. and A.I.; Project Administration, A.I.; Funding Acquisition, A.I. All authors have read and agreed to the published version of the manuscript.

Funding

The authors of this work want to acknowledge financial support from the Spanish Ministry of Economy and Competitiveness, through the project CTQ2016-76231-C2-1-R (AEI/FEDER, UE).

Acknowledgments

The authors gratefully acknowledge the Spanish Ministry of Economy and Competitiveness, through the project CTQ2016-76231-C2-1-R (AEI/FEDER, UE) for financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Overview of Greenhouse Gases: United States Environmental Protection Agency. Available online: https://www.epa.gov/ghgemissions/overview-greenhouse-gases (accessed on 16 June 2020).
  2. Sustainable Development Goals: United Nations. Available online: https://sustainabledevelopment.un.org/ (accessed on 16 June 2020).
  3. Irabien, A.; Alvarez-Guerra, M.; Albo, J.; Domínguez-Ramos, A. Electrochemical conversion of CO2 to value-added products. In Electrochememical Water Wastewater Treatment; Martínez-Huitle, C.A., Rodrigo, M.A., Scialdone, O., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 29–59. [Google Scholar]
  4. Qiao, J.; Liu, Y.; Zhan, J. Electrochemical Reduction of Carbon Dioxide. Fundamentals and Technologies; CRC Press: Boca Raton, FL, USA, 2016; ISBN 9781482258257. [Google Scholar]
  5. Bui, M.; Adjiman, C.S.; Bardow, A.; Anthony, E.J.; Boston, A.; Brown, S.; Fennell, P.S.; Fuss, S.; Galindo, A.; Hackett, L.A.; et al. Carbon capture and storage (CCS): The way forward. Energy Environ. Sci. 2018, 11, 1062–1176. [Google Scholar] [CrossRef] [Green Version]
  6. Sohaib, Q.; Vadillo, J.M.; Gómez-Coma, L.; Albo, J.; Druon-Bocquet, S.; Irabien, A.; Sanchez-Marcano, J. CO2 capture with room temperature ionic liquids; coupled absorption/desorption and single module absorption in membrane contactor. Chem. Eng. Sci. 2020, 223, 115719. [Google Scholar] [CrossRef]
  7. Nocito, F.; Dibenedetto, A. Atmospheric CO2 mitigation technologies: Carbon capture utilization and storage. Curr. Opin. Green Sustain. Chem. 2020, 21, 34–43. [Google Scholar] [CrossRef]
  8. Mustafa, A.; Lougou, B.G.; Shuai, Y.; Wang, Z.; Tan, H. Current technology development for CO2 utilization into solar fuels and chemicals: A review. J. Energy Chem. 2020, 49, 96–123. [Google Scholar] [CrossRef]
  9. Vadillo, J.M.; Gómez-Coma, L.; Garea, A.; Irabien, A. CO2 desorption performance from imidazolium ionic liquids by membrane vacuum regeneration technology. Membranes 2020, 10, 234. [Google Scholar] [CrossRef]
  10. 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]
  11. Ting, L.R.L.; Yeo, B.S. Recent advances in understanding mechanisms for the electrochemical reduction of carbon dioxide. Curr. Opin. Electrochem. 2018, 8, 126–134. [Google Scholar] [CrossRef]
  12. Zheng, Y.; Zhang, W.; Li, Y.; Chen, J.; Yu, B.; Wang, J.; Zhang, L.; Zhang, J. Energy related CO2 conversion and utilization: Advanced materials/nanomaterials, reaction mechanisms and technologies. Nano Energy 2017, 40, 512–539. [Google Scholar] [CrossRef]
  13. Yaashikaa, P.R.; Senthil Kumar, P.; Varjani, S.J.; Saravanan, A. A review on photochemical, biochemical and electrochemical transformation of CO2 into value-added products. J. CO2 Util. 2019, 33, 131–147. [Google Scholar] [CrossRef]
  14. Merino-Garcia, I.; Alvarez-Guerra, E.; Albo, J.; Irabien, A. Electrochemical membrane reactors for the utilisation of carbon dioxide. Chem. Eng. J. 2016, 305, 104–120. [Google Scholar] [CrossRef] [Green Version]
  15. Castro, S.; Albo, J.; Irabien, A. Photoelectrochemical Reactors for CO2 Utilization. ACS Sustain. Chem. Eng. 2018, 6, 15877–15894. [Google Scholar] [CrossRef] [Green Version]
  16. Kumar, B.; Brian, J.P.; Atla, V.; Kumari, S.; Bertram, K.A.; White, R.T.; Spurgeon, J.M. New trends in the development of heterogeneous catalysts for electrochemical CO2 reduction. Catal. Today 2016, 270, 19–30. [Google Scholar] [CrossRef]
  17. Lu, Q.; Jiao, F. Electrochemical CO2 reduction: Electrocatalyst, reaction mechanism, and process engineering. Nano Energy 2016, 29, 439–456. [Google Scholar] [CrossRef] [Green Version]
  18. Endrődi, B.; Bencsik, G.; Darvas, F.; Jones, R.; Rajeshwar, K.; Janáky, C. Continuous-flow electroreduction of carbon dioxide. Prog. Energy Combust. Sci. 2017, 62, 133–154. [Google Scholar] [CrossRef]
  19. Gao, D.; Arán-Ais, R.M.; Jeon, H.S.; Roldan Cuenya, B. Rational catalyst and electrolyte design for CO2 electroreduction towards multicarbon products. Nat. Catal. 2019, 2, 198–210. [Google Scholar] [CrossRef]
  20. Ren, S.; Joulié, D.; Salvatore, D.; Torbensen, K.; Wang, M.; Robert, M.; Berlinguette, C.P. Molecular electrocatalysts can mediate fast, selective CO2 reduction in a flow cell. Science 2019, 365, 367–369. [Google Scholar] [CrossRef]
  21. Jouny, M.; Hutchings, G.S.; Jiao, F. Carbon monoxide electroreduction as an emerging platform for carbon utilization. Nat. Catal. 2019, 2, 1062–1070. [Google Scholar] [CrossRef]
  22. Du, D.; Lan, R.; Humphreys, J.; Tao, S. Progress in inorganic cathode catalysts for electrochemical conversion of carbon dioxide into formate or formic acid. J. Appl. Electrochem. 2017, 47, 661–678. [Google Scholar] [CrossRef] [Green Version]
  23. Taheri, A.; Berben, L.A. Making C-H bonds with CO2: Production of formate by molecular electrocatalysts. Chem. Commun. 2016, 52, 1768–1777. [Google Scholar] [CrossRef]
  24. Wu, Y.; Jiang, Z.; Lu, X.; Liang, Y.; Wang, H. Domino electroreduction of CO2 to methanol on a molecular catalyst. Nature 2019, 575, 639–642. [Google Scholar] [CrossRef]
  25. 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]
  26. Zhang, W.; Hu, Y.; Ma, L.; Zhu, G.; Wang, Y.; Xue, X.; Chen, R.; Yang, S.; Jin, Z. Progress and Perspective of Electrocatalytic CO2 Reduction for Renewable Carbonaceous Fuels and Chemicals. Adv. Sci. 2018, 5, 1700275. [Google Scholar] [CrossRef] [PubMed]
  27. Zhang, L.; Merino-Garcia, I.; Albo, J.; Sánchez-Sánchez, C.M. Electrochemical CO2 reduction reaction on cost-effective oxide-derived copper and transition metal–nitrogen–carbon catalysts. Curr. Opin. Electrochem. 2020, 23, 65–73. [Google Scholar] [CrossRef]
  28. Zhao, J.; Xue, S.; Barber, J.; Zhou, Y.; Meng, J.; Ke, X. An overview of Cu-based heterogeneous electrocatalysts for CO2 reduction. J. Mater. Chem. A 2020, 8, 4700–4734. [Google Scholar] [CrossRef]
  29. Agarwal, A.S.; Zhai, Y.; Hill, D.; Sridhar, N. The electrochemical reduction of carbon dioxide to formate/formic acid: Engineering and economic feasibility. ChemSusChem 2011, 4, 1301–1310. [Google Scholar] [CrossRef]
  30. Bienen, F.; Kopljar, D.; Löwe, A.; Aßmann, P.; Stoll, M.; Rößner, P.; Wagner, N.; Friedrich, A.; Klemm, E. Utilizing Formate as an Energy Carrier by Coupling CO2 Electrolysis with Fuel Cell Devices. Chem. Ing. Tech. 2019, 91, 872–882. [Google Scholar] [CrossRef]
  31. An, L.; Chen, R. Direct formate fuel cells: A review. J. Power Sources 2016, 320, 127–139. [Google Scholar] [CrossRef]
  32. Preuster, P.; Albert, J. Biogenic Formic Acid as a Green Hydrogen Carrier. Energy Technol. 2018, 6, 501–509. [Google Scholar] [CrossRef] [Green Version]
  33. Eppinger, J.; Huang, K.W. Formic Acid as a Hydrogen Energy Carrier. ACS Energy Lett. 2017, 2, 188–195. [Google Scholar] [CrossRef] [Green Version]
  34. Han, N.; Ding, P.; He, L.; Li, Y.; Li, Y. Promises of Main Group Metal–Based Nanostructured Materials for Electrochemical CO2 Reduction to Formate. Adv. Energy Mater. 2019, 10, 1902338. [Google Scholar] [CrossRef]
  35. Zhao, S.; Li, S.; Guo, T.; Zhang, S.; Wang, J.; Wu, Y.; Chen, Y. Advances in Sn-Based Catalysts for Electrochemical CO2 Reduction. Nano-Micro Lett. 2019, 11, 62. [Google Scholar] [CrossRef] [Green Version]
  36. Sen, S.; Liu, D.; Palmore, G.T.R. Electrochemical reduction of CO2 at copper nanofoams. ACS Catal. 2014, 4, 3091–3095. [Google Scholar] [CrossRef]
  37. Gao, S.; Lin, Y.; Jiao, X.; Sun, Y.; Luo, Q.; Zhang, W.; Li, D.; Yang, J.; Xie, Y. Partially oxidized atomic cobalt layers for carbon dioxide electroreduction to liquid fuel. Nature 2016, 529, 68–71. [Google Scholar] [CrossRef] [PubMed]
  38. Yuan, M.; Sahin, S.; Cai, R.; Abdellaoui, S.; Hickey, D.P.; Minteer, S.D.; Milton, R.D. Creating a Low-Potential Redox Polymer for Efficient Electroenzymatic CO2 Reduction. Angew. Chem. Int. Ed. 2018, 57, 6582–6586. [Google Scholar] [CrossRef] [PubMed]
  39. Pander, J.E.; Lum, J.W.J.; Yeo, B.S. The importance of morphology on the activity of lead cathodes for the reduction of carbon dioxide to formate. J. Mater. Chem. A 2019, 7, 4093–4101. [Google Scholar] [CrossRef]
  40. Alvarez-Guerra, M.; Quintanilla, S.; Irabien, A. Conversion of carbon dioxide into formate using a continuous electrochemical reduction process in a lead cathode. Chem. Eng. J. 2012, 207–208, 278–284. [Google Scholar] [CrossRef]
  41. Gálvez-Vázquez, M.D.J.; Moreno-García, P.; Guo, H.; Hou, Y.; Dutta, A.; Waldvogel, S.R.; Broekmann, P. Leaded Bronze Alloy as a Catalyst for the Electroreduction of CO2. ChemElectroChem 2019, 6, 2324–2330. [Google Scholar] [CrossRef]
  42. Luo, W.; Xie, W.; Li, M.; Zhang, J.; Züttel, A. 3D hierarchical porous indium catalyst for highly efficient electroreduction of CO2. J. Mater. Chem. A 2019, 7, 4505–4515. [Google Scholar] [CrossRef]
  43. Zha, B.; Li, C.; Li, J. Efficient electrochemical reduction of CO2 into formate and acetate in polyoxometalate catholyte with indium catalyst. J. Catal. 2020, 382, 69–76. [Google Scholar] [CrossRef]
  44. Yuan, X.; Luo, Y.; Zhang, B.; Dong, C.; Lei, J.; Yi, F.; Duan, T.; Zhu, W.; He, R. Decoration of in nanoparticles on In2S3 nanosheets enables efficient electrochemical reduction of CO2. Chem. Commun. 2020, 56, 4212–4215. [Google Scholar] [CrossRef]
  45. Klinkova, A.; De Luna, P.; Dinh, C.T.; Voznyy, O.; Larin, E.M.; Kumacheva, E.; Sargent, E.H. Rational Design of Efficient Palladium Catalysts for Electroreduction of Carbon Dioxide to Formate. ACS Catal. 2016, 6, 8115–8120. [Google Scholar] [CrossRef]
  46. Gao, D.; Zhou, H.; Cai, F.; Wang, D.; Hu, Y.; Jiang, B.; Cai, W.-B.; Chen, X.; Si, R.; Yang, F.; et al. Switchable CO2 electroreduction via engineering active phases of Pd nanoparticles. Nano Res. 2017, 10, 2181–2191. [Google Scholar] [CrossRef]
  47. Lee, W.; Kim, Y.E.; Youn, M.H.; Jeong, S.K.; Park, K.T. Catholyte-Free Electrocatalytic CO2 Reduction to Formate. Angew. Chem. Int. Ed. 2018, 57, 6883–6887. [Google Scholar] [CrossRef]
  48. Li, D.; Wu, J.; Liu, T.; Liu, J.; Yan, Z.; Zhen, L.; Feng, Y. Tuning the pore structure of porous tin foam electrodes for enhanced electrochemical reduction of carbon dioxide to formate. Chem. Eng. J. 2019, 375, 122024. [Google Scholar] [CrossRef]
  49. He, G.; Tang, H.; Wang, H.; Bian, Z. Highly Selective and Active Pd-In/three-dimensional Graphene with Special Structure for Electroreduction CO2 to Formate. Electroanalysis 2018, 30, 84–93. [Google Scholar] [CrossRef]
  50. Proietto, F.; Schiavo, B.; Galia, A.; Scialdone, O. Electrochemical conversion of CO2 to HCOOH at tin cathode in a pressurized undivided filter-press cell. Electrochim. Acta 2018, 277, 30–40. [Google Scholar] [CrossRef]
  51. Alvarez-Guerra, M.; Del Castillo, A.; Irabien, A. Continuous electrochemical reduction of carbon dioxide into formate using a tin cathode: Comparison with lead cathode. Chem. Eng. Res. Des. 2014, 92, 692–701. [Google Scholar] [CrossRef] [Green Version]
  52. Del Castillo, A.; Alvarez-Guerra, M.; Irabien, A. Continuous electroreduction of CO2 to formate using Sn gas diffusion electrodes. AIChE J. 2014, 60, 3557–3564. [Google Scholar] [CrossRef]
  53. Del Castillo, A.; Alvarez-Guerra, M.; Solla-Gullón, J.; Sáez, A.; Montiel, V.; Irabien, A. Electrocatalytic reduction of CO2 to formate using particulate Sn electrodes: Effect of metal loading and particle size. Appl. Energy 2015, 157, 165–173. [Google Scholar] [CrossRef]
  54. Fu, Y.; Wang, T.; Zheng, W.; Lei, C.; Yang, B.; Chen, J.; Li, Z.; Lei, L.; Yuan, C.; Hou, Y. Nanoconfined Tin Oxide within N-Doped Nanocarbon Supported on Electrochemically Exfoliated Graphene for Efficient Electroreduction of CO2 to Formate and C1 Products. ACS Appl. Mater. Interf. 2020, 12, 16178–16185. [Google Scholar] [CrossRef]
  55. Zhang, X.; Sun, X.; Guo, S.-X.; Bond, A.M.; Zhang, J. Formation of lattice-dislocated bismuth nanowires on copper foam for enhanced electrocatalytic CO2 reduction at low overpotential. Energy Environ. Sci. 2019, 12, 1334–1340. [Google Scholar] [CrossRef]
  56. Wu, D.; Huo, G.; Chen, W.Y.; Fu, X.Z.; Luo, J.L. Boosting formate production at high current density from CO2 electroreduction on defect-rich hierarchical mesoporous Bi/Bi2O3 junction nanosheets. Appl. Catal. B Environ. 2020, 271, 118957. [Google Scholar] [CrossRef]
  57. Wang, Q.; Zhu, C.; Wu, C.; Yu, H. Direct synthesis of bismuth nanosheets on a gas diffusion layer as a high-performance cathode for a coupled electrochemical system capable of electroreduction of CO2 to formate with simultaneous degradation of organic pollutants. Electrochim. Acta 2019, 319, 138–147. [Google Scholar] [CrossRef]
  58. Yang, F.; Elnabawy, A.O.; Schimmenti, R.; Song, P.; Wang, J.; Peng, Z.; Yao, S.; Deng, R.; Song, S.; Lin, Y.; et al. Bismuthene for highly efficient carbon dioxide electroreduction reaction. Nat. Commun. 2020, 11, 1088. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Zhu, C.; Wang, Q.; Wu, C. Rapid and scalable synthesis of bismuth dendrites on copper mesh as a high-performance cathode for electroreduction of CO2 to formate. J. CO2 Util. 2020, 36, 96–104. [Google Scholar] [CrossRef]
  60. Tran-Phu, T.; Daiyan, R.; Fusco, Z.; Ma, Z.; Amal, R.; Tricoli, A. Nanostructured β-Bi2O3 Fractals on Carbon Fibers for Highly Selective CO2 Electroreduction to Formate. Adv. Funct. Mater. 2020, 30, 1–8. [Google Scholar] [CrossRef]
  61. Lu, P.; Gao, D.; He, H.; Wang, Q.; Liu, Z.; Dipazir, S.; Yuan, M.; Zu, W.; Zhang, G. Facile synthesis of a bismuth nanostructure with enhanced selectivity for electrochemical conversion of CO2 to formate. Nanoscale 2019, 11, 7805–7812. [Google Scholar] [CrossRef]
  62. García de Arquer, F.P.; Bushuyev, O.S.; De Luna, P.; Dinh, C.T.; Seifitokaldani, A.; Saidaminov, M.I.; Tan, C.S.; Quan, L.N.; Proppe, A.; Kibria, M.G.; et al. 2D Metal Oxyhalide-Derived Catalysts for Efficient CO2 Electroreduction. Adv. Mater. 2018, 30, 6–11. [Google Scholar] [CrossRef]
  63. Kumawat, A.S.; Sarkar, A. Comparative Study of Carbon Supported Pb, Bi and Sn Catalysts for Electroreduction of Carbon Dioxide in Alkaline Medium. J. Electrochem. Soc. 2017, 164, H1112–H1120. [Google Scholar] [CrossRef]
  64. Vennekoetter, J.B.; Sengpiel, R.; Wessling, M. Beyond the catalyst: How electrode and reactor design determine the product spectrum during electrochemical CO2 reduction. Chem. Eng. J. 2019, 364, 89–101. [Google Scholar] [CrossRef]
  65. Vennekötter, J.B.; Scheuermann, T.; Sengpiel, R.; Wessling, M. The electrolyte matters: Stable systems for high rate electrochemical CO2 reduction. J. CO2 Util. 2019, 32, 202–213. [Google Scholar]
  66. Del Castillo, A.; Alvarez-Guerra, M.; Solla-Gullón, J.; Sáez, A.; Montiel, V.; Irabien, A. Sn nanoparticles on gas diffusion electrodes: Synthesis, characterization and use for continuous CO2 electroreduction to formate. J. CO2 Util. 2017, 18, 222–228. [Google Scholar] [CrossRef] [Green Version]
  67. Díaz-Sainz, G.; Alvarez-Guerra, M.; Solla-Gullón, J.; García-Cruz, L.; Montiel, V.; Irabien, A. Catalyst coated membrane electrodes for the gas phase CO2 electroreduction to formate. Catal. Today 2020, 346, 58–64. [Google Scholar] [CrossRef]
  68. Díaz-Sainz, G.; Alvarez-Guerra, M.; Solla-Gullón, J.; García-Cruz, L.; Montiel, V.; Irabien, A. CO2 electroreduction to formate: Continuous single-pass operation in a filter-press reactor at high current densities using Bi gas diffusion electrodes. J. CO2 Util. 2019, 34, 12–19. [Google Scholar]
  69. Díaz-Sainz, G.; Alvarez-Guerra, M.; Solla-Gullón, J.; García-Cruz, L.; Montiel, V.; Irabien, A. Gas–liquid–solid reaction system for CO2 electroreduction to formate without using supporting electrolyte. AIChE J. 2020, 66, e16299. [Google Scholar]
  70. Xia, C.; Zhu, P.; Jiang, Q.; Pan, Y.; Liang, W.; Stavitsk, E. Continuous production of pure liquid fuel solutions via electrocatalytic CO2 reduction using solid-electrolyte devices. Nat. Energy 2019, 5, 776–785. [Google Scholar] [CrossRef]
  71. Díaz-Sainz, G.; Alvarez-Guerra, M.; Ávila-Bolívar, B.; Solla-Gullón, J.; Montiel, V.; Irabien, A. Improving trade-offs in the figures of merit of gas-phase single-pass continuous CO2 electrocatalytic reduction to formate. Chem. Eng. J. 2020, 405, 126965. [Google Scholar] [CrossRef]
  72. Chen, Y.; Vise, A.; Klein, W.E.; Cetinbas, F.C.; Myers, D.J.; Smith, W.A.; Deutsch, T.G.; Neyerlin, K.C. A Robust, Scalable Platform for the Electrochemical Conversion of CO2 to Formate: Identifying Pathways to Higher Energy Efficiencies. ACS Energy Lett. 2020, 5, 1825–1833. [Google Scholar] [CrossRef]
  73. Marcos-Madrazo, A.; Casado-Coterillo, C.; Irabien, Á. Sustainable Membrane-Coated Electrodes for CO2 Electroreduction to Methanol in Alkaline Media. ChemElectroChem 2019, 6, 5273–5282. [Google Scholar] [CrossRef]
  74. Merino-Garcia, I.; Albo, J.; Krzywda, P.; Mul, G.; Irabien, A. Bimetallic Cu-based hollow fibre electrodes for CO2 electroreduction. Catal. Today 2020, 346, 34–39. [Google Scholar] [CrossRef]
  75. Hosseini, M.G.; Mahmoodi, R.; Daneshvari-Esfahlan, V. Ni@Pd core-shell nanostructure supported on multi-walled carbon nanotubes as efficient anode nanocatalysts for direct methanol fuel cells with membrane electrode assembly prepared by catalyst coated membrane method. Energy 2018, 161, 1074–1084. [Google Scholar] [CrossRef]
  76. Ávila-Bolívar, B.; García-Cruz, L.; Montiel, V.; Solla-Gullón, J. Electrochemical Reduction of CO2 to Formate on Easily Prepared Carbon-Supported Bi Nanoparticles. Molecules 2019, 24, 2032. [Google Scholar]
Figure 1. Influence of current density on (a) Faradaic Efficiency for HCOO (%) and (b) HCOO rate (mmol·m−2·s−1) in the current density range of 90–300 mA·cm−2 operating with an electrolyte flow = 5.7 mL·min−1. Sn/C-GDEs and Bi/C-GDEs are indicated in green and red, respectively. Data taken from references [66,68]. Sn-C: Sn carbon-supported nanoparticles; Bi/C: Sn carbon-supported nanoparticles.
Figure 1. Influence of current density on (a) Faradaic Efficiency for HCOO (%) and (b) HCOO rate (mmol·m−2·s−1) in the current density range of 90–300 mA·cm−2 operating with an electrolyte flow = 5.7 mL·min−1. Sn/C-GDEs and Bi/C-GDEs are indicated in green and red, respectively. Data taken from references [66,68]. Sn-C: Sn carbon-supported nanoparticles; Bi/C: Sn carbon-supported nanoparticles.
Molecules 25 04457 g001aMolecules 25 04457 g001b
Figure 2. Influence of electrolyte flow on the HCOO concentration (g·L−1) in the electrolyte flow range of 0.7–5.7 mL·min−1 applied at different current densities: 90 and 200 mA·cm−2 and with different electrocatalyst material (Sn/C NPs and Bi/C NPs). Data taken from references [66,68].
Figure 2. Influence of electrolyte flow on the HCOO concentration (g·L−1) in the electrolyte flow range of 0.7–5.7 mL·min−1 applied at different current densities: 90 and 200 mA·cm−2 and with different electrocatalyst material (Sn/C NPs and Bi/C NPs). Data taken from references [66,68].
Molecules 25 04457 g002
Figure 3. Influence of temperature on (a) Faradaic Efficiency for HCOO (%), (b) HCOO rate (mmol·m−2·s−1), (c) energy consumption per kmol of HCOO (kWh·kmol−1), and (d) HCOO concentration (g·L−1) in the temperature range of 20–50 °C applied at a constant current density = 45 mA·cm−2, a catalyst loading = 0.75 mg·cm−2, and a relative humidity = 100%. Sn/C-CCMEs and Bi/C-CCMEs are indicated in green and red, respectively. Data taken from references [67,69].
Figure 3. Influence of temperature on (a) Faradaic Efficiency for HCOO (%), (b) HCOO rate (mmol·m−2·s−1), (c) energy consumption per kmol of HCOO (kWh·kmol−1), and (d) HCOO concentration (g·L−1) in the temperature range of 20–50 °C applied at a constant current density = 45 mA·cm−2, a catalyst loading = 0.75 mg·cm−2, and a relative humidity = 100%. Sn/C-CCMEs and Bi/C-CCMEs are indicated in green and red, respectively. Data taken from references [67,69].
Molecules 25 04457 g003aMolecules 25 04457 g003bMolecules 25 04457 g003c
Figure 4. Influence of water flow in the CO2 stream on (a) Faradaic efficiency for HCOO (%), (b) HCOO rate (mmol·m−2·s−1), (c) energy consumption per kmol of HCOO (kWh·kmol−1), and (d) HCOO concentration (g·L−1) in the water flow range of 0.15–1 g·h−1 applied at a constant current density = 45 mA·cm−2, a catalyst loading = 0.75 mg·cm−2, and a temperature = 20 °C Sn/C-CCMEs and Bi/C-CCMEs are indicated in green and red, respectively. Data taken from references [67,69].
Figure 4. Influence of water flow in the CO2 stream on (a) Faradaic efficiency for HCOO (%), (b) HCOO rate (mmol·m−2·s−1), (c) energy consumption per kmol of HCOO (kWh·kmol−1), and (d) HCOO concentration (g·L−1) in the water flow range of 0.15–1 g·h−1 applied at a constant current density = 45 mA·cm−2, a catalyst loading = 0.75 mg·cm−2, and a temperature = 20 °C Sn/C-CCMEs and Bi/C-CCMEs are indicated in green and red, respectively. Data taken from references [67,69].
Molecules 25 04457 g004aMolecules 25 04457 g004b
Figure 5. Scheme of (a) Gas Diffusion Electrode (GDE) configuration and (b) Catalyst Coated Membrane Electrode (CCME) configuration (1: the counter electrode; 2: the cationic exchange membrane; 3: the catalytic layer; 4: the microporous layer, and 5: the carbon support). Please note that there are no elements 4 and 5 in (b) because in the CCME configuration, the catalyst is deposited directly on the membrane, avoiding the use of a carbon support and microporous layer.
Figure 5. Scheme of (a) Gas Diffusion Electrode (GDE) configuration and (b) Catalyst Coated Membrane Electrode (CCME) configuration (1: the counter electrode; 2: the cationic exchange membrane; 3: the catalytic layer; 4: the microporous layer, and 5: the carbon support). Please note that there are no elements 4 and 5 in (b) because in the CCME configuration, the catalyst is deposited directly on the membrane, avoiding the use of a carbon support and microporous layer.
Molecules 25 04457 g005
Table 1. Figures of merit (current density, flow in the cathode side, HCOO concentration, HCOO rate, Faradaic efficiency (FE) for HCOO and energy consumption per kmol of HCOO) for the electrocatalytic reduction of CO2 to HCOO employing different Sn-based electrodes (Sn/C-GDEs and Sn/C-CCMEs). Data taken from references [66,67].
Table 1. Figures of merit (current density, flow in the cathode side, HCOO concentration, HCOO rate, Faradaic efficiency (FE) for HCOO and energy consumption per kmol of HCOO) for the electrocatalytic reduction of CO2 to HCOO employing different Sn-based electrodes (Sn/C-GDEs and Sn/C-CCMEs). Data taken from references [66,67].
Figures of MeritConfiguration
Sn/C-GDEs [66]Sn/C-CCMEs [67]
Flow in the cathode side5.75.75.70.7~0.008 (0.5 g·h1)
(mL·min−1)
Current density9015020020045
(mA·cm−2)
Absolute cell potential3.13.74.04.32.2
(V)
HCOO concentration1.52.52.716.919.2
(g·L−1)
HCOO rate3.235.455.614.381.15
(mmol·m−2·s−1)
FE for HCOO69.47054.142.349.4
(%)
Energy consumption239282396513244
(kWh·kmol−1)
Table 2. Figures of merit (current density, HCOO concentration, HCOO rate, FE for HCOO, and energy consumption per kmol of HCOO) for the electrocatalytic reduction of CO2 to HCOO employing different Bi-based electrodes (GDEs and CCMEs). Data taken from references [68,69].
Table 2. Figures of merit (current density, HCOO concentration, HCOO rate, FE for HCOO, and energy consumption per kmol of HCOO) for the electrocatalytic reduction of CO2 to HCOO employing different Bi-based electrodes (GDEs and CCMEs). Data taken from references [68,69].
Figures of MeritConfiguration
Bi/C-GDEs [68]Bi/C-CCMEs [69]
Flow in the cathode side5.75.75.75.70.7~0.008 (0.5 g·h−1)
(mL·min−1)
Current density9015020030020045
(mA·cm−2)
Absolute cell potential3.13.74.25.44.52.7
(V)
HCOO concentration2.03.063.95.218.025.9
(g·L−1)
HCOO rate4.316.468.3310.974.671.28
(mmol·m−2·s−1)
FE for HCOO92.483.180.470.645.154.8
(%)
Energy consumption177240277410535266
(kWh·kmol−1)
Table 3. Energy consumption per kmol of HCOO for the electrocatalytic reduction of CO2 to HCOO working with Sn/C-GDEs and Bi/C-GDEs at different operating conditions (current density and catholyte flow). Data taken from references [66,68].
Table 3. Energy consumption per kmol of HCOO for the electrocatalytic reduction of CO2 to HCOO working with Sn/C-GDEs and Bi/C-GDEs at different operating conditions (current density and catholyte flow). Data taken from references [66,68].
Operating ConditionEnergy Consumption (kWh·kmol−1 of HCOO)
Current Density (mA·cm−2)Catholyte Flow (mL·min−1)Sn/C-GDEs [66]Bi/C-GDEs [68]
905.7239177
150282240
200396277
901.5267186
200395364
0.7544535
Table 4. Value of thickness of the different layers shown in Figure 5 in each of the electrodes employed in the references [66,67,68,69].
Table 4. Value of thickness of the different layers shown in Figure 5 in each of the electrodes employed in the references [66,67,68,69].
Sn/C GDEsBi/C GDEsSn/C CCMEsBi/C CCMEs
Thickness12 mm
2183 μm
350–60 μm15–20 μm15 μm15–20 μm
4100 μm100–125 μmNo microporous layer
5190 μm No carbon support
Table 5. Value of the different operating conditions taken in the references [66,67,68,69].
Table 5. Value of the different operating conditions taken in the references [66,67,68,69].
Operating ConditionValue
Anolyte flow (mL·min−1)5.7
KOH concentration in anolyte (mol·L−1)1
CO2 flow (mL·min−1)200
Catalyst loading (mg·cm−2)0.75
Reaction time (min)90
Electrode area (cm2)10

Share and Cite

MDPI and ACS Style

Díaz-Sainz, G.; Alvarez-Guerra, M.; Irabien, A. Continuous Electrochemical Reduction of CO2 to Formate: Comparative Study of the Influence of the Electrode Configuration with Sn and Bi-Based Electrocatalysts. Molecules 2020, 25, 4457. https://doi.org/10.3390/molecules25194457

AMA Style

Díaz-Sainz G, Alvarez-Guerra M, Irabien A. Continuous Electrochemical Reduction of CO2 to Formate: Comparative Study of the Influence of the Electrode Configuration with Sn and Bi-Based Electrocatalysts. Molecules. 2020; 25(19):4457. https://doi.org/10.3390/molecules25194457

Chicago/Turabian Style

Díaz-Sainz, Guillermo, Manuel Alvarez-Guerra, and Angel Irabien. 2020. "Continuous Electrochemical Reduction of CO2 to Formate: Comparative Study of the Influence of the Electrode Configuration with Sn and Bi-Based Electrocatalysts" Molecules 25, no. 19: 4457. https://doi.org/10.3390/molecules25194457

Article Metrics

Back to TopTop