Next Article in Journal
Parallel Combination of Inner Capacitance and Ionic Capacitance, Apparently Inconsistent with Stern’s Model
Previous Article in Journal
Applying Different Configurations for the Thermal Management of a Lithium Titanate Oxide Battery Pack
Previous Article in Special Issue
Electrochemical Reduction of CO2: Overcoming Chemical Inertness at Ambient Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Direct Electrochemical Reduction of Bicarbonate to Formate Using Tin Catalyst

1
Eurecat, Centre Tecnològic de Catalunya, C/Marcel·lí Domingo, 43007 Tarragona, Spain
2
Department of Chemical Engineering, Universitat Rovira I Virgili, Av. Països Catalans, 26, 43007 Tarragona, Spain
*
Author to whom correspondence should be addressed.
Submission received: 9 December 2020 / Revised: 5 February 2021 / Accepted: 5 February 2021 / Published: 10 February 2021
(This article belongs to the Special Issue Electroreduction of CO2 to Fuels and Chemicals)

Abstract

:
Nowadays, the self-accelerating increase in global temperatures strengthens the idea that the cutting of CO2 emissions will not be enough to avoid climate change, thus CO2 from the atmosphere must be removed. This gas can be easily trapped by converting it to bicarbonate using hydroxide solutions. However, bicarbonate must be converted into a more valuable product to make this technology profitable. Several studies show great efficiency when reducing bicarbonate solutions saturated with pure CO2 gas to formate. However, those approaches don’t have a real application and our objective was to obtain similar results without pure CO2 saturation. The method consists of electroreduction of the bicarbonate solution using bulk tin (Sn) as catalysts. Tin is a relatively cheap material that, according to previous studies performed in saturated bicarbonate solutions, shows a great selectivity towards formate. The 1H NMR analysis of bicarbonate solutions after electroreduction show that, without pure CO2 gas, the faradic efficiency is around 18% but almost 50% for saturated ones. The formate obtained could be used to power formate/formic acid fuel cells obtaining a battery-like system, with greater energy density than common lithium batteries, but electroreduction efficiency needs to be improved to make them competitive.

1. Introduction

Since the start of the industrial revolution, the atmospheric CO2 levels increased drastically and exceeded the 400 ppm in 2016 [1]. For this reason, some studies indicate that just cutting off or reducing the CO2 emissions will not be enough to avoid or mitigate the effects of climate change, and CO2 removal from the atmosphere is going to be necessary [2]. One way to absorb atmospheric CO2 is by transforming it into bicarbonate using hydroxide solutions, like the system developed by A. Nogalska and R. Garcia [3]. Bicarbonate is a product with low value and not many applications, but its conversion to more useful and valuable hydrocarbons is becoming an economically viable way to reuse it [4].
Many different direct CO2 electroreduction techniques use bicarbonate solutions previously saturated with pure CO2 gas to obtain hydrocarbons or alcohols showing significant selectivity. The main inconvenience of these techniques are its practical applications because CO2 is present in the atmosphere at very low concentrations, and it is almost impossible to saturate a bicarbonate solution with CO2 using atmospheric air, with just a few studies performing it on non-CO2 saturated bicarbonate solutions [5,6,7,8,9,10,11]. Theoretically, direct CO2 electroreduction is performed by CO2 activation on a catalyst surface and its following reaction with two protons (2H+). Generally, the bicarbonate acts as an electrolyte, favoring the CO2 dissolution and increasing the conductivity of the aqueous solution, but, since dissolved bicarbonate is always in equilibrium with dissolved CO2, it is still not clear if the carbon source is bicarbonate or CO2 [12,13,14,15,16].
The electrochemical reduction of CO2 into formate performed with water as a proton source is a very attractive technology, due to the great energy density of formate that can be used to power formate fuel cells [17,18,19], as a future alternative to lithium batteries [20,21]. In addition, this liquid fuel may be easily stored and transported using existing infrastructures. This way, renewable energy, which is intermittent, unpredictable, and with irregular production peaks, could be stored as formate and released during low production periods without any additional emissions of CO2 into the atmosphere.
Our goal is to study the viability of reducing a non-CO2 saturated bicarbonate solution into formate using a bulk tin catalyst, which is a relatively cheap material, simple to use and with great selectivity towards formate in mild conditions [22,23,24,25,26]. For this purpose, electrochemical studies were performed with the use of potassium bicarbonate. Moreover, the influence of CO2 saturation on faradic efficiency of conversion was evaluated.

2. Materials and Methods

2.1. Materials

Tin foil (99.8% trace metals basis), 0.25 mm thick, provided by Alfa Aesar (Haverhill, MA, USA), was used as a working electrode and highly pure KHCO3 (Bio Ultra 99.5%), supplied by Sigma-Aldrich (St. Louis, MO, USA), with a content of iron lower than 5 mg/Kg was used for preparation of the electrolyte. Milli-Q water was used to prepare solutions. Proton exchange membrane was Nafion® 117 membrane (Sigma-Aldrich). Hydrogen peroxide 30% (v/v) (Sigma-Aldrich) and 95–97% H2SO4 (Serviquimia) were used to prepare cleaning solutions for Nafion membrane. To prepare a standard for 1H NMR analysis, 99.8% D2O from Panreac and 99.7% DMSO (Chromasolv Plus) from Sigma Aldrich were used. CO2 gas used for the electrolyte saturation was pure liquid carbon dioxide (CO2 Premier X10) purchased from Carburos Metálicos.

2.2. Linear Sweep Voltammetry

Linear sweep voltammetry (LSV) was performed prior to electroreduction experiments to evaluate which is the lowest potential that is high enough to allow significant product formation, without experimenting excessive H2 gas formation at high potentials. To study the effects of KHCO3 concentration, three solutions with different concentrations (0.1, 0.5, and 1.5 M) were prepared. Finally, the 1.5 M KHCO3 solution was submitted to saturation with CO2 gas during 30 min at a flow rate of 14.47 sccm to test its effects. Constant stirring was applied to ensure homogeneous solution, improve bicarbonate migration to the electrode, and facilitate the release of residual gas bubbles formed during reaction.
All experiments were carried out in a gas-tight Teflon H-cell (Figure 1 and Figure S1), equipped with the standard three-electrode system. The counter electrode was a bulk Platinum (Pt) foil of 1 cm2, the reference electrode was an Ag/AgCl/3M KCl (0.21 V), and the working electrode consisted of a tin (Sn) foil of 5 cm2. A Nafion membrane was used to separate the working and counter electrode compartments (25 mL each) to prevent oxidation of the reduced products. The membrane was washed by immersion in an 80 °C aqueous solution, for 1 h each, in the following order: 0.5M H2SO4, H2O2 3%, and distilled water to eliminate possible contamination (Figure S2). The working electrode was not pretreated before experiments, as the native SnO2 layer is reported to increase the catalytic activity [27], and the whole system was leaned with distilled water between each experiment. An AutoLab PGSTAT 302N potentiostat (Metrohm, Autolab B.V.) was used in all the electrochemical experiments. All the LSV were conducted at a scan rate of 100 mV/s in a range of potentials between 0 and −2.0 V in the presence of the reference electrode.

2.3. Electroreduction Experiments

The electroreduction experiments were carried out in potentiostatic conditions for 1.5 h in the same Teflon H-cell used for LSV analysis. In addition, constant stirring, CO2 pre-saturation, and membrane-system cleaning were performed as described in the previous section and for the same purposes. The electroreduction potential was chosen based on the LSV analysis and set to −1.6 V. Beyond that, a significant increase of reduction current and gas bubble formation associated with CO (2) and H2 (3) formation can be observed [28]. Some of those bubbles stayed in the working electrode surface, diminishing its active area, and reducing the efficiency of the reaction. The materials and the samples used are the same described in the voltammetry section.
The most probable electroreduction reaction of bicarbonate to formate in these conditions is showed (1), but formate formation from dissolved CO2 is also possible [12,13,14,15,16]:
HCO3−(aq) + 2H+ (aq) + 2e− → HCO2− (aq) + H2O
The most likely side reactions of CO and H2 formation are:
2CO2 (aq) → 2CO (g) + O2 (g)
2H+ + 2e− → H2 (g)

2.4. Product Analysis

The analysis of the reaction’s products was performed by 1H NMR (VARIAN Mercury VX400). DMSO 0.010M in D2O was used as an internal standard for quantification of the obtained products. Catholyte samples of 600 μL were mixed with 100 μL of standard. To obtain the peaks coming from products, we suppressed the water peak by irradiation.
Based on the determination of formate concentration though the NMR analysis with Mestrenova, Faradic efficiency FE (%) was calculated considering passed charge and electrons transferred with the following Equation (4):
FE (%) = qexp/qtheo × 100
Experimentally used charge [C] qexp (5) is calculated by:
qexp = Fnz
F stays for Faraday constant 9.6485 × 104 [C/mol], n is the amount of formate moles obtained by proton NMR, and z are electrons needed for bicarbonate conversion to formate based on chemical reaction (2e−).
In addition, theoretical charge qtheo (6) [C] passed charge is obtained chronoamperograms by:
Qexp = −It
where I is the total current [A] and t is the reaction time [s].

3. Results and Discussion

3.1. Linear Sweep Voltammetry (LSV)

The LSV measurements show a small peak between −0.9 V to −1.0 V corresponding to reduction of the small SnO2 layer usually present on bulk Sn [21] (Figure 2). At potentials below −1.2 V, current density increases exponentially in all cases, probably associated with bicarbonate and hydrogen electroreduction. In addition, as current density increases, CO2 bubbling on working electrode (Sn foil) is observed, probably associated with H2 or CO and other residual gasses [29]. Furthermore, oxygen bubbles are observed on the platinum counter electrode. Current density is higher when CO2 presaturation is used (Figure 2b), and the product analysis will determine if this is associated with more bicarbonate electroreduction or not.

3.2. Chronoamperometry

The working potential for electroreduction experiments was set at −1.6 V in order to obtain the greatest energy density without observing excessive formation of residual non-desired gasses (chronoamperograms shown in Figure 3). The current density should increase while augmenting KHCO3 concentration, due to the improved conductivity of the solution and increased reaction rates. As expected, the results confirm an energy density six times higher for 1.5 M KHCO3 solution respectively compared with 0.1 M KHCO3 solution at −1.6 V (Figure 3a). For the CO2 saturated experiments, the same trend is observed with a 33% increase on the current density at −1.6 V for the saturated solution (Figure 3b). The electroreduction current is stable during this interval of time for all experiments, slightly decreasing over time for experiments performed at 1.5 M KHCO3, which could be due to faster exhaustion of reactants, membrane poisoning, or catalyst degradation or even a reduction of solution volume. A small amount of noise most commonly present in experiments at higher currents are due to gas bubbles formation, which attach and detach from the catalyst surface affecting the active area.

3.3. Analysis and Quantification of Products by 1H NMR

The 1H NMR spectra show three singlets at around 8.24, 4.63 and 2.50 ppm corresponding to HCOOK, DMSO, and H2O, respectively (Figure S3). The slightly basic media (~pH 9) of the potassium bicarbonate solution, including the non-observation of an O-H bond of formic acid around 11 ppm, confirm that potassium formate (HCOOK) is the main product, and the acid form is not present for reaction conditions. The observation of only one product in the liquid phase indicates a high selectivity of tin foil towards formate production.
The faradaic efficiency towards formate shows a significative increase from 0% to 18% in 0.1 M and 1.5 M solutions as summarized in Table 1. With the saturation of CO2, the efficiency increases from 18% to 47%, noting the importance of pre-saturating the solution with pure CO2.
As we can see in Table 2, bulk Sn seems to have the highest efficiency compared with other similar bulk catalysts working in similar conditions.

4. Conclusions

During this brief study, we confirmed the possibility to reduce bicarbonate solutions to formate without using pure CO2, with significant amounts for non-CO2 saturated solutions (18%) and almost 50% for saturated ones.
Bicarbonate can be easily obtained from atmospheric air by CO2 capture, and there is an unlimited stock of it. In addition, CO2 pollutes the atmosphere and increases the greenhouse effect, with serious consequences for our planet and society, so pulling it out of the atmosphere will contribute to mitigate its negative consequences. The direct conversion of atmospheric CO2 stored as bicarbonate will make this technology more competitive in the market, and to have an alternative for formic acid/formate production which is nowadays obtained by a non-renewable and very contaminating process. Another great application of formic acid is fuel cells because they have higher energy density than lithium batteries and they are safer than hydrogen fuels cells. Then, the possibility to have more competitive fuel cells with much higher energy density than current lithium batteries will also increase the competitiveness of renewable energies allowing the possibility to store more energy during its normally irregular energy production peaks.
To conclude, we can affirm that increasing the electroreduction efficiency and selectivity of bicarbonate to formate will become a great advance in energy storage systems, thus making renewable energies even more competitive, then investing in this technologies is a great way to fight climate change and take economical profit at the same time.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2673-3293/2/1/6/s1, Table S1: List of abbreviations, Figure S1: 3D rendered design of Teflon H-cell used for the experiments, Figure S2: Nafion membrane cleaning procedure, Figure S3: 1H NMR spectra of products after electroreduction.

Author Contributions

Conceptualization, R.G.-V.; methodology, A.B.N. and A.N.; investigation, A.B.N. and A.N.; writing—original draft preparation, A.B.N. and A.N.; writing—review and editing, A.N. and R.G.-V.; visualization, A.B.N.; supervision, R.G.-V. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Regional Agency for the Business Competitiveness of the Generalitat de Catalunya (EURECAT-ACCIÓ). Moreover, the Vincente Lopez scholarship from EURECAT was provided for Andreu Bonet Navarro pre-doctoral studies and Ministerio de Economía y Competitividad (ENE2017-86711-C3-3-R).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

“Not applicable” for studies not involving humans.

Data Availability Statement

Data is contained within the article or supplementary material.

Acknowledgments

The authors would like to acknowledge Ramon Guerrero Grueso for technical support with 1H-NMR analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Smith, M.R.; Myers, S.S. Impact of Anthropogenic CO2 Emissions on Global Human Nutrition. Nat. Clim. Chang. 2018, 8, 834–839. [Google Scholar] [CrossRef]
  2. Workman, M.; McGlashan, N.; Chalmers, H.; Shah, N. An Assessment of Options for CO2 Removal from the Atmosphere. Energy Procedia 2011, 4, 2877–2884. [Google Scholar] [CrossRef] [Green Version]
  3. Nogalska, A.; Zukowska, A.; Garcia-Valls, R. Atmospheric CO2 Capture for the Artificial Photosynthetic System. Sci. Total Environ. 2018, 621, 186–192. [Google Scholar] [CrossRef]
  4. 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]
  5. Hosseini, S.; Kheawhom, S.; Soltani, S.M.; Aroua, M.K. Electrochemical Reduction of Bicarbonate on Carbon Nanotube-Supported Silver Oxide: An Electrochemical Impedance Spectroscopy Study. J. Environ. Chem. Eng. 2018, 6, 1033–1043. [Google Scholar] [CrossRef] [Green Version]
  6. Spichiger-Ulmann, M.; Augustynski, J. Electrochemical Reduction of Bicarbonate Ions at a Bright Palladium Cathode. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1985, 81, 713–716. [Google Scholar] [CrossRef]
  7. Li, T.; Lees, E.W.; Goldman, M.; Salvatore, D.A.; Weekes, D.M.; Berlinguette, C.P. Electrolytic Conversion of Bicarbonate into CO in a Flow Cell. Joule 2019, 3, 1487–1497. [Google Scholar] [CrossRef] [Green Version]
  8. Min, X.; Kanan, M.W. Pd-Catalyzed Electrohydrogenation of Carbon Dioxide to Formate: High Mass Activity at Low Overpotential and Identification of the Deactivation Pathway. J. Am. Chem. Soc. 2015, 137, 4701–4708. [Google Scholar] [CrossRef] [PubMed]
  9. Zhong, H.; Fujii, K.; Nakano, Y.; Jin, F. Effect of CO2 Bubbling into Aqueous Solutions Used for Electrochemical Reduction of CO2 for Energy Conversion and Storage. J. Phys. Chem. C 2014, 119, 55–61. [Google Scholar] [CrossRef]
  10. Kortlever, R.; Tan, K.H.; Kwon, Y.; Koper, M.T.M. Electrochemical Carbon Dioxide and Bicarbonate Reduction on Copper in Weakly Alkaline Media. J. Solid State Electrochem. 2013, 17, 1843–1849. [Google Scholar] [CrossRef]
  11. Hori, Y. Electrolytic Reduction of Bicarbonate Ion at a Mercury Electrode. J. Electrochem. Soc. 1983, 130, 2387–2390. [Google Scholar] [CrossRef]
  12. Sreekanth, N.; Phani, K.L. Selective Reduction of CO2 to Formate through Bicarbonate Reduction on Metal Electrodes: New Insights Gained from SG/TC Mode of SECM. Chem. Commun. 2014, 50, 11143–11146. [Google Scholar] [CrossRef] [PubMed]
  13. 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]
  14. Isa, P.; Louis, H.; Adesina, K.; Akpan, E. Review article Understanding the Mechanism of Electrochemical Reduction of CO2 Using Cu/Cu-Based Electrodes. Asian J. Nano Mater. 2018, 1, 183–224. [Google Scholar]
  15. Garza, A.J.; Bell, A.T.; Gordon, M.H. Mechanism of CO2 Reduction at Copper Surfaces: Pathways to C2 Products. ACS Catal. 2018, 8, 1490–1499. [Google Scholar] [CrossRef] [Green Version]
  16. Hussain, J.; Skúlason, E.; Jónsson, H. Computational Study of Electrochemical CO2 Reduction at Transition Metal Electrodes. Procedia Comput. Sci. 2015, 51, 1865–1871. [Google Scholar] [CrossRef]
  17. Aslam, N.; Masdar, M.; Kamarudin, S.; Daud, W.R.W. Overview on Direct Formic Acid Fuel Cells (DFAFCs) as an Energy Sources. APCBEE Procedia 2012, 3, 33–39. [Google Scholar] [CrossRef] [Green Version]
  18. Rees, N.V.; Compton, R.G. Sustainable Energy: A Review of Formic Acid Electrochemical Fuel Cells. J. Solid State Electrochem. 2011, 15, 2095–2100. [Google Scholar] [CrossRef]
  19. Nogalska, A.; Navarro, A.B.; Garcia-Valls, R. MEA Preparation for Direct Formate/Formic Acid Fuel Cell—Comparison of Palladium Black and Palladium Supported on Activated Carbon Performance on Power Generation in Passive Fuel Cell. Membranes 2020, 10, 355. [Google Scholar] [CrossRef] [PubMed]
  20. Sharaf, O.Z.; Orhan, M.F. An Overview of Fuel Cell Technology: Fundamentals and Applications. Renew. Sustain. Energy Rev. 2014, 32, 810–853. [Google Scholar] [CrossRef]
  21. Mohapatra, A.; Tripathy, S. A Critical Review of the use of Fuel Cells Towards Sustainable Management of Resources. In IOP Conference Series: Materials Science and Engineering, Proceedings of the International Conference on Mechanical, Materials and Renewable Energy, Sikkim, India, 8–10 December 2017; IOP Publishing: Bristol, UK, 2018; Volume 377, p. 012135. [Google Scholar] [CrossRef]
  22. Prakash, G.K.S.; Viva, F.A.; Olah, G.A. Electrochemical Reduction of CO2 over Sn-Nafion® Coated Electrode for a Fuel-Cell-like Device. J. Power Sources 2013, 223, 68–73. [Google Scholar] [CrossRef]
  23. Cui, C.; Wang, H.; Zhu, X.; Han, J.; Ge, Q. A DFT Study of CO2 Electrochemical Reduction on Pb(211) and Sn(112). Sci. China Chem. 2015, 58, 607–613. [Google Scholar] [CrossRef]
  24. Bumroongsakulsawat, P.; Kelsall, G. Effect of Solution pH on CO: Formate Formation Rates during Electrochemical Reduction of Aqueous CO2 at Sn Cathodes. Electrochim. Acta 2014, 141, 216–225. [Google Scholar] [CrossRef]
  25. Saravanan, K.; Basdogan, Y.; Dean, J.; Keith, J.A. Computational Investigation of CO2 Electroreduction on Tin Oxide and Predictions of Ti, V, Nb and Zr Dopants for Improved Catalysis. J. Mater. Chem. A 2017, 5, 11756–11763. [Google Scholar] [CrossRef]
  26. Medina-Ramos, J.; Pupillo, R.C.; Keane, T.P.; DiMeglio, J.L.; Rosenthal, J. Efficient Conversion of CO2 to CO Using Tin and Other Inexpensive and Easily Prepared Post-Transition Metal Catalysts. J. Am. Chem. Soc. 2015, 137, 5021–5027. [Google Scholar] [CrossRef]
  27. Zhang, R.; Lv, W.; Lei, L. Role of the Oxide Layer on SN Electrode in Electrochemical Reduction of CO2 to Formate. Appl. Surf. Sci. 2015, 356, 24–29. [Google Scholar] [CrossRef]
  28. Sheng, W.; Kattel, S.; Yao, S.; Yan, B.; Liang, Z.; Hawxhurst, C.J.; Wu, Q.; Chen, J.G. Electrochemical Reduction of CO2 to Synthesis Gas with Controlled CO/H2 ratios. Energy Environ. Sci. 2017, 10, 1180–1185. [Google Scholar] [CrossRef]
  29. Rumayor, M.; Dominguez-Ramos, A.; Irabien, A. Formic Acid Manufacture: Carbon Dioxide Utilization Alternatives. Appl. Sci. 2018, 8, 914. [Google Scholar] [CrossRef] [Green Version]
  30. Noda, H.; Ikeda, S.; Oda, Y.; Imai, K.; Maeda, M.; Ito, K. Electrochemical Reduction of Carbon Dioxide at Various Metal Electrodes in Aqueous Potassium Hydrogen Carbonate Solution. Bull. Chem. Soc. Jpn. 1990, 63, 2459–2462. [Google Scholar] [CrossRef] [Green Version]
  31. Nakagawa, S.; Kudo, A.; Azuma, M.; Sakata, T. Effect of Pressure on the Electrochemical Reduction of CO2 on Group VIII Metal Electrodes. J. Electroanal. Chem. Interfacial Electrochem. 1991, 308, 339–343. [Google Scholar] [CrossRef]
  32. Yang, D.; Zhu, Q.; Chen, C.; Liu, H.; Liu, Z.; Zhao, Z.; Zhang, X.; Liu, S.; Han, B. Selective Electroreduction of Carbon Dioxide to Methanol on Copper Selenide Nanocatalysts. Nat. Commun. 2019, 10, 1–9. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Cross section of Teflon H-cell used for the experiments. The cell consists of two compartments (Anodic and Cathodic) filled with the bicarbonate solution and separated by a proton exchange membrane (4) to prevent re-oxidation of reduced products on the anode. The cathodic compartment contains the working electrode (2), where the CO2 is reduced and the reference electrode (1) (Ag/AgCl (0.21 V) is used to maintain a stable potential reading. In the anodic compartment, there is the counter electrode (3) (Pt foil) where water splitting takes place. For the saturation of solution with CO2, two inlets (5) and outlets (6) for gas were used.
Figure 1. Cross section of Teflon H-cell used for the experiments. The cell consists of two compartments (Anodic and Cathodic) filled with the bicarbonate solution and separated by a proton exchange membrane (4) to prevent re-oxidation of reduced products on the anode. The cathodic compartment contains the working electrode (2), where the CO2 is reduced and the reference electrode (1) (Ag/AgCl (0.21 V) is used to maintain a stable potential reading. In the anodic compartment, there is the counter electrode (3) (Pt foil) where water splitting takes place. For the saturation of solution with CO2, two inlets (5) and outlets (6) for gas were used.
Electrochem 02 00006 g001
Figure 2. (a) LSV) at different potassium bicarbonate (KHCO3) concentrations of non-CO2 saturated solutions; (b) comparison of a linear sweep voltammetry (LSV) between saturated with CO2 and non-saturated potassium bicarbonate (KHCO3) aqueous solutions.
Figure 2. (a) LSV) at different potassium bicarbonate (KHCO3) concentrations of non-CO2 saturated solutions; (b) comparison of a linear sweep voltammetry (LSV) between saturated with CO2 and non-saturated potassium bicarbonate (KHCO3) aqueous solutions.
Electrochem 02 00006 g002
Figure 3. (a) Chronoamperogram at −1.6 V of different potassium bicarbonate (KHCO3) concentrations of non-CO2 saturated solutions; (b) chronoamperogram at −1.6 V of CO2 saturated and non-saturated bicarbonate solution.
Figure 3. (a) Chronoamperogram at −1.6 V of different potassium bicarbonate (KHCO3) concentrations of non-CO2 saturated solutions; (b) chronoamperogram at −1.6 V of CO2 saturated and non-saturated bicarbonate solution.
Electrochem 02 00006 g003
Table 1. Summary of the effects of the potassium bicarbonate (KHCO3) concentration and the CO2 pre-saturation on the faradaic efficiency of the electroreduction to formate.
Table 1. Summary of the effects of the potassium bicarbonate (KHCO3) concentration and the CO2 pre-saturation on the faradaic efficiency of the electroreduction to formate.
KHCO3 (aq) Concentration (M)0.1 M0.5 M1.5 M1.5 M
CO2 pre-saturation NoNoNoYes
Faradaic efficiency (%)n.d.81847
n.d.—not detected.
Table 2. Summary of CO2 electroreduction efficiencies towards HCOOH/HCOOK of similar bulk catalysts.
Table 2. Summary of CO2 electroreduction efficiencies towards HCOOH/HCOOK of similar bulk catalysts.
ElectrodeReference ElectrodeElectrolyteCO2 SaturationFaradaic Efficiency (%) towards HCOOK/HCOOHRef.
Ag (99.98%) electrode−1.6 V vs. Ag/AgCl saturated with KCl0.1M KHCO3 aqueous solutionNon[30]
Au (99.95%) electrode6
Pd Metal1.8 V vs. Ag/AgCl0.1 M KHCO3 aqueous solutionNo4.4[31]
Cu-Based catalystsAg/Ag+ with 0.01 M0.5 M KHCO3 aqueous solutionYes3–15[32]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bonet Navarro, A.; Nogalska, A.; Garcia-Valls, R. Direct Electrochemical Reduction of Bicarbonate to Formate Using Tin Catalyst. Electrochem 2021, 2, 64-70. https://0-doi-org.brum.beds.ac.uk/10.3390/electrochem2010006

AMA Style

Bonet Navarro A, Nogalska A, Garcia-Valls R. Direct Electrochemical Reduction of Bicarbonate to Formate Using Tin Catalyst. Electrochem. 2021; 2(1):64-70. https://0-doi-org.brum.beds.ac.uk/10.3390/electrochem2010006

Chicago/Turabian Style

Bonet Navarro, Andreu, Adrianna Nogalska, and Ricard Garcia-Valls. 2021. "Direct Electrochemical Reduction of Bicarbonate to Formate Using Tin Catalyst" Electrochem 2, no. 1: 64-70. https://0-doi-org.brum.beds.ac.uk/10.3390/electrochem2010006

Article Metrics

Back to TopTop