Next Article in Journal
Experimental Analysis of Bio-Inspired Vortex Generators on a Blade with S822 Airfoil
Previous Article in Journal
Calcium Borohydride Ca(BH4)2: Fundamentals, Prediction and Probing for High-Capacity Energy Storage Applications, Organic Synthesis and Catalysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancement of the Catalytic Effect on the Electrochemical Conversion of CO2 to Formic Acid Using MXene (Ti3C2Tx)-Modified Boron-Doped Diamond Electrode

by
Prastika Krisma Jiwanti
1,*,
Asmaul Mashad Alfaza
2,
Grandprix T. M. Kadja
3,4,5,*,
Suci A. C. Natalya
3,
Fuja Sagita
3,
Yasuaki Einaga
6,
Aning Purwaningsih
2,
Ilma Amalina
1 and
Intan Nurul Rizki
1
1
Nanotechnology Engineering, Faculty of Advanced Technology and Multidiscipline, Airlangga University, Surabaya 60115, Indonesia
2
Department of Chemistry, Faculty of Science and Technology, Airlangga University, Surabaya 60115, Indonesia
3
Division of Inorganic and Physical Chemistry, Faculty of Mathematics and Natural Sciences, Institut Teknologi Bandung, Jl. Ganesha No. 10, Bandung 40132, Indonesia
4
Center for Catalysis and Reaction Engineering, Institut Teknologi Bandung, Jl. Ganesha No. 10, Bandung 40132, Indonesia
5
Research Center for Nanosciences and Nanotechnology, Institut Teknologi Bandung, Jl. Ganesha No. 10, Bandung 40132, Indonesia
6
Department of Chemistry, Keio University, 3-14-1 Hiyoshi, Yokohama 223-8522, Japan
*
Authors to whom correspondence should be addressed.
Energies 2023, 16(12), 4537; https://doi.org/10.3390/en16124537
Submission received: 17 May 2023 / Revised: 23 May 2023 / Accepted: 31 May 2023 / Published: 6 June 2023
(This article belongs to the Section B: Energy and Environment)

Abstract

:
The rising concentration of carbon dioxide (CO2) as one of the greenhouse gases in the atmosphere is a major source of worry. Electrochemical reduction of CO2 is one of many ways to convert CO2 gas into usable compounds. An electrochemical technique was applied in this study to reduce CO2 using a boron-doped diamond (BDD) working electrode modified with MXene (Ti3C2Tx) material to improve electrode performance. MXene concentrations of 0.5 mg/mL (MXene-BDD 0.5), 1.0 mg/mL (MXene-BDD 1.0), and 2.0 mg/mL (MXene-BDD 2.0) were drop-casted onto the BDD surface. MXene was effectively deposited on top of the BDD surface, with Ti weight loads of 0.12%, 4.06%, and 7.14% on MXene-BDD 0.5, MXene-BDD 1.0, and MXene-BDD 2.0, respectively. The modified working electrode was employed for CO2 electroreduction with optimal CO2 gas aeration. The existence of the MXene substance in BDD reduced the electroreduction overpotential of CO2. For the final result, we found that the MXene-BDD 2.0 electrode effectively generated the most formic acid product with a maximum reduction potential as low as −1.3 V (vs. Ag/AgCl).

1. Introduction

The causes of extreme usage of fossil fuel in transportation and industry has been well reported, due to high CO2 gas emissions and the shortage of fossil fuel [1,2]. According to recent research, the use of fossil fuel energy contributes to 96.5% of CO2 emissions. Heavy CO2 emissions led to global warming [3]. Since the industrial revolution, human activity has increased by 30%, and it is expected to continue growing year by year [4]. Indonesia generated 617.51 million tons of CO2 gas emissions in 2019 [5]. As a result, an alternative energy source that possibly keeps CO2 levels in the atmosphere stable is necessary. There have been attempts to overcome this problem through carbon storage; the other method is CO2 conversion, a process that creates chemical stocks or further converts it back to fuels. As a result, these substances may be put to good use and may improve its value [6]. As for CO2 conversion, researchers have been attracted to electrochemical techniques because of their mild working conditions, adjustable reaction conditions and rates, reusable catalysts and electrolytes, simple equipment, diverse product selectivity, and higher conversion efficiency [7,8]. The CO2 molecule may be reduced through electrochemical reduction into CO2•−, which can then be used to generate a number of different compounds [9,10,11].
Meanwhile, the boron-doped diamond (BDD) electrode is a carbon-based electrode [12], which has excellent properties, such as high physical and chemical stability, wide potential window, low background current, and biocompatibility [13,14,15,16]. This wide potential window suppresses the hydrogen evolution reaction that is widely known to be the competitor for CO2 electrochemical reactions, especially in aqueous solutions. The Einaga group has been working on CO2 reduction for years using the polycrystalline BDD electrode, and formic acid (HCOOH) is known to be the main product on the bare BDD condition, with its potential around −2.2 V [17,18]. This high reduction potential can be suppressed by modifying the BDD surface with metal nanoparticles or higher catalytic materials. As an example, the modification of the BDD surface with nanoparticles such as Cu or Pd can suppress the reduction potential, as well as the production of higher carbon-containing compound products.
On the other hand, Ti3C2Tx serves as one of the members of MXene, a class of 2D morphological carbide metal transition materials with nanoscale thicknesses (nanosheets) [19]. Tx in the Ti3C2Tx structure stands for the surface terminal functional groups, i.e., –OH, –O–, –F, and/or –Cl. This material can be modified by adding specific functional groups. This material’s further important attributes include excellent conductivity, mechanical strength, and flexibility [20]. MXene is studied computationally and prospected for CO2 reduction reaction [21]. However, experimental studies are required to realize the potential of MXene. In our previous study, we found that modifying Ti3C2Tx on a bare BDD electrode might increase the catalytic effect and potentially reduce the CO2 overpotential reduction required on a bare BDD, which can then be utilized for an actual CO2 reduction application [22]. In this report, the electrochemical reduction product of CO2 on the surface of a Ti3C2Tx (MXene)-modified BDD electrode with mainly HCOOH will be investigated. The CO2 reduction to HCOOH on the surface of the MXene-modified BDD electrode is predicted to be greater than that on the bare BDD electrode.

2. Materials and Methods

2.1. Materials

The materials used in this research were isopropanol, KOH (≥85%), KCl (99.0–100.5%), HCl (37%), HNO3 (≥90%), H2SO4 (95–98%), HCOOH (≥96%), and a Nafion membrane from Sigma Aldrich. A microwave plasma-assisted chemical vapor deposition (MP-CVD, Model AX-5400, CORNES Technology Corp., Tokyo, Japan) was used to produce 1% (B/C) boron-doped diamond on the surface of a silicon wafer (111) for 6 h. All chemicals were used without additional purification.

2.2. Preparation and Characterization of the MXene-BDD Electrode

MXene was prepared, based on our previous work, using LiF–HCl treatment [23]. The as-prepared MXene was then characterized with scanning electron microscopy (SEM, Hitachi SU3500, Japan), X-ray diffraction (XRD, Bruker D8 Advance, Germany), and Raman spectroscopy (HORIBA HR Evolution Raman, Japan). To prepare the modified MXene-BDD electrode, first, the MXene solution was prepared at concentrations of 0.5 mg/mL (MXene-BDD 0.5), 1.0 mg/mL (MXene-BDD 1.0), and 2.0 mg/mL (MXene-BDD 2.0) in ultrapure water. Further, 20 µL of each solution was drop-casted onto the surface of the BDD electrode and dried at room temperature. After that, the modified electrodes were characterized with SEM and were employed for CO2 electrochemical reduction.

2.3. CO2 Electrochemical Reduction Process

The CO2 electrochemical reduction processes were performed in a two-compartment, closed electrochemical cell, with the cathode compartment carrying 0.5 M KCl electrolyte and the anode compartment carrying 0.5 M KOH electrolyte. The electrolyte used in this work was based on our previous work [24]. The use of the KOH aqueous solution might suppress the overpotential in the anolyte side, as reported in a previous work [25]. The Nafion membrane, an ion exchange membrane, separated the two compartments. This cell used a three-electrode system, with Pt as the counter electrode, Ag/AgCl as the reference electrode, and the modified and unmodified BDD electrodes with a geometric area of 0.754 cm2 as the working electrodes. Prior to each reduction process, N2 and CO2 gas purging was carried out for 15 min, and the CO2 electrochemical reduction was performed for 1 h. All electrochemical data were recorded using Emstat3+ blue PalmSens. High performance liquid chromatography (HPLC, Agilent 1260, Japan) was used to evaluate the product of the CO2 reduction.

3. Results

3.1. MXene-BDD Electrode Preparation and Characterization

SEM was used to study the surface morphology of the as-prepared MXene. It can be seen that the MXene sheets had a layered structure, showing the nature of two-dimensional nanomaterials (Figure 1a,b). The XRD result, depicted in Figure 1c, showed the main peak of Ti3C2Tx MXene at 2θ of around 6°, corresponding to the (002) plane. Moreover, the Raman spectra of Ti3C2Tx MXene demonstrated several important regions that contained information about its local structure. The Raman shift at 150 cm−1 to 250 cm−1 corresponded to the flake region, a combination of Ti, C, and Tx surface group (–F, –O–, and –OH) vibrations, while the bands at 225 cm−1 to 500 cm−1 were solely attributed to the Tx vibrations. In addition, the carbon region, located at 550 cm−1 to 800 cm−1, was ascribed to the in-plane and out-of-plane vibrations of the carbon layer [26]. Two prominent bands were observed at 1375 cm−1 and 1575 cm−1, assigned to the D band (sp2 carbon) and G band (perturbed sp2 carbon), respectively [23].
The BDD (B/C 1%) was fabricated on a silicon wafer using MP-CVD technology, with carbon and boron precursors from methane and B(OCH3)3. Prior to modification, BDD was pre-treated to clear the BDD electrode of impurities of elements or other compounds on its surface. This pretreatment was carried out electrochemically by performing a CV for 40 cycles in 0.1 M H2SO4 solution, with a potential range of −2.5 V to +2.5 V (vs. Ag/AgCl) and with a scan rate of 1 V/s. The modified BDD electrode was created by depositing MXene (Ti3C2Tx) on the BDD surface using the drop-cast technique. MXene concentrations ranged from 0.5 mg/mL, to 1.0 mg/mL, and to 2.0 mg/mL; as much as 20 µL of each variation was deposited on the BDD surface and left to dry at room temperature.
Cyclic voltammetry (CV) was performed in a one-compartment cell with three electrodes: Ag/AgCl electrodes as reference electrodes, Pt spiral electrode as a counter electrode, and BDD electrodes modified with MXene compounds as working electrodes (MXene-BDD). CV was performed in 0.1 M H2SO4 solution at a scan rate of 100 mV/s within a potential range of −1.0 V to +1.5 V (vs. Ag/AgCl). As shown in Figure 2, the CV results of the modified BDD electrode were compared to those of the unmodified BDD electrode (bare BDD). The cathodic current started to decrease to a more negative current value at a potential of −0.4 V (vs. Ag/AgCl) when the concentration of the MXene compound deposited on the BDD surface increased, indicating the presence of the MXene compound. Moreover, the oxidation peak at a potential of around −0.5 V was observed. The linearity test of the oxidation peak at a potential of −0.5 V showed the value of R2 = 0.9925, which is considered a good linearity value. Finally, these reversible redox natures (at a potential of −1.0 V to −0.1 V) belonged to the protonation/deprotonation of the MXene surface by the changes in the Ti redox state [27].
Figure 3 shows the SEM images of the BDD surface before and after MXene was deposited. From this SEM image, we can see that the amount of MXene on the surface of BDD improved as the concentration of MXene increased. In the MXene-BDD 0.5, some of the BDD grain was still exposed, while on MXene-BDD 2.0, MXene was covering the entire surface of BDD. Further analysis was performed using EDX to identify the composition of the components distributed throughout the surface of the modified BDD. Table 1 shows the composition of the elements present on the MXene-BDD surface. This table also confirmed the presence of MXene on the surface of BDD, where the element of Ti increased as the concentration of MXene increased.

3.2. Electrochemical Reduction of CO2

3.2.1. Determining the Optimal CO2 Bubbling Time

The CO2 gas was aerated and dissolved into the 0.5 M KCl electrolyte. CO2 gas contained in a solution largely consisting of water can experience carbonic acid equilibrium, notably in Equations (1)–(3), where carbonic acid can be degraded to create H+ ions and HCO3 species, resulting in a pH decrease. Because of this pH decrease, the optimal CO2 aeration period can be determined using the pH parameters of the electrolyte solution.
CO2 (aq) + H2O (l) ⇄ H2CO3 (aq)  pKa = 2.77
H2CO3 (aq) ⇄ H+ (aq) + HCO3 (aq)  pKa = 6.35
HCO3 (aq) ⇄ H+ (aq) + CO32− (aq)  pKa = 10.33
The pH parameter indicates which carbon species are more abundant in the solution, based on the distribution of carbon species in aqueous solutions. Dissolved CO2 is the most common species at a pH less than 5, HCO3 (bicarbonate ion) is the most abundant species at a pH of 7.5 to 9, and CO32− (carbonate ion) is the most abundant species at a pH greater than 12. As a result, the electrolyte solution employed in this study had the most dissolved CO2 species, leading to a reduction in dissolved CO2 [27].
Figure 4A shows the results of pH measurements, which showed that the 0.5 M KCl electrolyte had an initial pH of 6.6. The KCl electrolyte was then aerated for 10 min with N2 gas to remove dissolved oxygen gas from the solution. The solution’s pH increased to 6.9 as a result of the N2 gas aeration. The solution was then aerated for 5 min with CO2 gas, which resulted in a pH decrease to 3.9, due to an increase in the amount of H+ ions, causing the pH of the solution to decrease. Meanwhile, after 10 and 15 min of aeration, the pH decreased, but the solution’s pH remained constant at 3.7 as the aeration time increased. This was due to the fact that the CO2 gas that was converted into dissolved CO2 was already saturated when it was aerated into the KCl electrolyte solution, and therefore, the pH of the electrolyte solution would remain stable at 3.7. According to the equilibrium reactions from Equations (1)–(3), the higher the concentration of CO2 dissolved in the solution, the higher the concentration of H+ generated, and therefore, the pH value decreased. In addition, an electrochemical investigation was conducted with linear sweep voltammetry (LSV) to identify the optimal CO2 aeration time by monitoring the peak reduction current of the dissolved CO2 species. LSV was performed at a scan rate of 100 mV/s over a potential range of 0 V to −1.7 V (vs. Ag/AgCl). The LSV in Figure 4B showed no peaks after 10 min of N2 gas aeration. Because the purpose of N2 gas aeration is to remove dissolved oxygen, this indicates that the electrolyte was cleared of dissolved oxygen. After aerating CO2 gas for 10 and 15 min, it showed a reduction current peak at roughly −1.5 V (vs. Ag/AgCl), suggesting the existence of the dissolved CO2 species in the electrolyte. There was no reduction in the present CO2 peak after 5 min of aeration. This was possible, since the amount of dissolved CO2 was still low and could not cover the surface of the working electrode. The largest shift in CO2 reduction current occurred after 15 min of aeration, compared to 5 min of aeration, and was only slightly different from 10 min of aeration. This indicated that 15 min of CO2 aeration generated the largest quantity of CO2 dissolved in the 0.5 M KCl electrolyte. Following that, 10 min of N2 aeration and 15 min of CO2 aeration were employed to begin CO2 electroreduction applications.

3.2.2. Preliminary Study of the Electrochemical Reduction of CO2

The first study of CO2 electroreduction through LSV was carried out by comparing the reduction activity of each variant of the working electrode, including bare BDD and modified BDD electrodes. The CO2 reduction activity experiment was conducted using the LSV across a potential range of 0 V to −1.7 V, with a scan rate of 100 mV/s after N2 and CO2 gas aerations.
The reduction activities of the bare BDD and the modified BDD electrodes with varied MXene concentrations are shown in Figure 5. There was a reduction signal at roughly −1.5 V for the bare BDD electrode, showing that there was reduction activity of dissolved CO2 to its reduced products, which could not be seen at the LSV once nitrogen gas was added. Furthermore, the presence of CO2 gas caused a shift in hydrogen evolution activity into a more negative potential, since CO2 reduction activity released reduced CO2 products, which covered the surface of the BDD, lowering hydrogen evolution activity [28]. There was a slight reduction signal of CO2 at a potential of about −1.4 V for MXene-BDD 0.5, a reduction signal from CO2 at a potential of around −1.2 V for MXene-BDD 1.0, and a reduction signal for CO2 gas at a potential of around −1.0 V for MXene-BDD 2.0. Based on these results, all MXene-BDD electrodes at different concentrations showed that as the deposited MXene concentration increased, the CO2 reduction activity shifted to a positive potential; it was suggested that the overpotential that occurred in the CO2 electroreduction was lower. Furthermore, the intensity of the signals rose with increasing concentrations of deposited MXene, indicating an increase in the quantity of CO2 reduction products formed at the surface of the BDD electrode.

3.3. CO2 Electrochemical Reduction Products

All products of CO2 electrochemical reduction in the cathode side were collected and analyzed using an HPLC instrument. In this report, we focused on the HCOOH products, given that HCOOH is a common byproduct of CO2 electroreduction on the BDD-based electrode. To explain the mechanism of CO2 reduction on the MXene-BDD electrode, knowing the active reduced species in the solution is important. The pH of the electrolyte solution after 10 min of CO2 aeration was 3.7. The pH parameter indicates which carbon species are more abundant in the solution, based on the distribution of carbon species in aqueous solutions [29]. Dissolved CO2 was the most common species at a pH less than 5, HCO3 (bicarbonate ion) was the most abundant species at a pH of 7.5 to 9, and CO32− (carbonate ion) was the most abundant species at a pH greater than 12. As a result, the electrolyte solution employed in this study had the most dissolved CO2 species, leading to a reduction in dissolved CO2. The HCOOH can be produced using three methods. The first pathway, the CO2•− anion radical, is created initially by giving one electron to CO2, where the oxygen in the CO2•− (i) anion radical is bound to the electrode surface. In this case, protonation will occur on the carbon atom to generate the HCOO (ii) intermediate, followed by the second electron transfer and the protonation step to produce the product HCOOH. The second method is not similar to the first, according to theoretical calculations. After the production of HCOO• (ii) intermediates, •OCHO (iii) intermediates can be generated by giving up one electron. Following that, HCOOH is produced by the protonation of •OCHO (iii). The third pathway is when the carbon in CO2•− attaches to the surface of the electrode (iv); the CO2•− intermediate may be reduced by the protonation of its own oxygen atom, resulting in the creation of •COOH (v). This intermediate can be converted to HCOOH or can lose H2O to generate CO [28]. The reaction’s equation is the following reaction: CO2 + 2H+ + 2e ⇄ HCOOH [30].

3.3.1. The Effect of MXene Concentration Variation

Using each of the bare BDD and MXene-BDD electrodes at varied concentrations, the quantification of CO2 electroreduction products at a potential of −1.7 V (vs. Ag/AgCl) was compared to the data given in Figure 6. As the variety of MXene deposited on the surface of the BDD electrode increased, so did the quantity of HCOOH products produced by the MXene-BDD electrode. The MXene-BDD 2.0 electrode produced the greatest HCOOH concentration, with a HCOOH concentration of 11.41 ppm. This could be explained by looking at the SEM EDS data, which revealed that the higher the concentration of MXene deposited on the electrode’s surface, the higher the mass percentage of MXene particles deposited. In addition, it showed that MXene catalyzed the production of HCOOH from the electrochemical reduction of CO2, where MXene had a high surface area and high electrical conductivity. Besides that, the EDX data (Table 1) showed a higher amount of element O, which was from the –O and –OH functional groups on the MXene surface. This promoted higher electronegativity on the surface of BDD with a higher MXene concentration, which led CO2 to easily come closer to the electrode surface and become further reduced.

3.3.2. The Effect of Reduction Potentials Variation

CO2 electrochemical reduction was performed for 1 h using the chronoamperometric technique under predetermined optimal conditions. The applied potential for electrochemical CO2 reduction was varied, specifically at −1.3 V, −1.5 V, and −1.7 V (vs. Ag/AgCl) for each working electrode of bare BDD and MXene-BDD. The product results and the faradaic efficiency for CO2 reduction are shown in Figure 7. It is suggested that the use of MXene-BDD 2.0 is able to produce large amounts of HCOOH. This may due to the catalytic effect of MXene on the CO2 reduction process, as explained in the previous section. The HCOOH production decreased as more negative potential reduction was applied. This loss in product yield at an increasingly negative potential might be caused by a huge shift in current density and the high production of H2 gas, which caused the release of MXene deposited on the BDD surface. As shown in the Figure 7, the bare BDD electrode generated the most HCOOH at a reduction potential of −1.7 V (vs. Ag/AgCl), the MXene-BDD 0.5 and MXene-BDD 1.0 electrodes produced the most at a potential of −1.5 V (vs. Ag/AgCl), and MXene-BDD electrode 2.0 produced the most at a reduction potential of −1.3 V (vs. Ag/AgCl). This data trend is supported by the previous report, in which the high yield of HCOOH was produced at a more negative potential (~2.2 V) [17], and thus, the modified MXene could suppress this higher potential up to a potential of −1.3 V. Moreover, the faradaic efficiency evaluation of BDD and all modified MXene-BDD electrodes showed a similar trend. MXene-BDD 2.0 could produce the highest efficiency of HCOOH of up to ~97% at a potential of −1.3 V. This suggests that MXene could suppress the energy used in the CO2 electrochemical reduction process.

3.4. MXene Stability Test against CO2 Electroreduction Application

The stability of the MXene compound deposited on the BDD surface needed to be evaluated on the modified BDD electrode utilized for CO2 electroreduction applications. The electrochemical approach using the CV method was performed for the characterization, as it is the fastest and most effective way to see the differences just after the CO2 reduction process finishes. The stability determination of the deposited MXene on the BDD surface was compared before and after being used for CO2 electrochemical reduction. The CV result of the MXene-BDD 2.0 electrode is shown in Figure 8, revealing the relatively instable MXene on the BDD surface after the reduction process at a potential of −1.7 V for 1 h. This was confirmed by the CV characterization results, which revealed the shift of the potential window of the electrode to a more negative potential. Furthermore, there was an oxidation peak of MXene before reduction at a potential of roughly −0.8 V (vs. Ag/AgCl) and a current of 142 µA, while the peak of MXene fell to 24 µA after reduction. This revealed that there was a 79.9% decrease in MXene on the BDD surface, indicating that the MXene compound was less stable on the BDD electrode surface. To overcome this problem, the MXene was re-deposited for each application. The future challenge is to optimize the unstable modification of MXene on the surface of BDD, which might be overcome by some chemical modifications, rather than let it be physically deposited.

4. Conclusions

MXene (Ti3C2Tx) was successfully deposited on the surface of BDD electrodes to be utilized for CO2 electrochemical reduction applications. The CO2 reduction result was HCOOH with the highest yield obtained on the MXene-BDD 2.0 electrode at a reduction potential of −1.3 V (vs. Ag/AgCl). The HCOOH formed had a concentration of 28.9 ppm. This result suggests that modifying MXene (Ti3C2Tx) on a bare BDD electrode might increase the catalytic effect and potentially solve the high potential reduction requirement for CO2 reduction on a bare BDD. The further challenge is to improve the stability of the modified MXene on the BDD surface. Additionally, improving the yield and selectivity of the MXene-BDD is currently attempted by modifying it with metal or alloy nanoparticles in the MXene nanolayers, which will be reported on in the near future. By improving the stability and obtaining a higher yield of HCOOH at a low overpotential using MXene modified on a chemically and physically stable BDD electrode, we could realize CO2 utilization technology by producing higher-valued products with low cost and clean technology.

Author Contributions

Conceptualization, P.K.J.; formal analysis, P.K.J., A.M.A. and G.T.M.K.; investigation, A.M.A., S.A.C.N. and F.S.; writing—original draft preparation, P.K.J., A.M.A. and G.T.M.K.; witting—review and editing, P.K.J. and G.T.M.K.; supervision, P.K.J., G.T.M.K., Y.E., A.P., I.A. and I.N.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research is funded by Universitas Airlangga under the International Research Collaboration Top#300 Scheme with the contract number of 348/UN3.15/PT/2023.

Data Availability Statement

The data presented in this study is available within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nurdiawansyah, N.; Lindrianasari, L. Carbon Emission Issues in Indonesia. Rev. Integr. Bus. Econ. Res. 2018, 7, 20–33. [Google Scholar]
  2. Friedlingstein, P.; Jones, M.W.; O’Sullivan, M.; Andrew, R.M.; Bakker, D.C.E. Global Carbon Budget 2021, Raport Earth System Science Data. Earth Syst. Sci. Data 2022, 14, 1917–2005. [Google Scholar] [CrossRef]
  3. Lau, L.C.; Lee, K.T.; Mohamed, A.R. Global Warming Mitigation and Renewable Energy Policy Development from the Kyoto Protocol to the Copenhagen Accord—A Comment. Renew. Sustain. Energy Rev. 2012, 16, 5280–5284. [Google Scholar] [CrossRef]
  4. Kleij, A.W.; North, M.; Urakawa, A. CO2 Catalysis. ChemSusChem 2017, 10, 1036–1038. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Friedlingstein, P.; O’Sullivan, M.; Jones, M.W.; Andrew, R.M.; Hauck, J.; Olsen, A.; Peters, G.P.; Peters, W.; Pongratz, J.; Sitch, S.; et al. Global Carbon Budget 2020. Earth Syst. Sci. Data 2020, 12, 3269–3340. [Google Scholar] [CrossRef]
  6. Wu, J.; Huang, Y.; Ye, W.; Li, Y. CO2 Reduction: From the Electrochemical to Photochemical Approach. Adv. Sci. 2017, 4, 1700194. [Google Scholar] [CrossRef]
  7. Janáky, C.; Hursán, D.; Endrödi, B.; Chanmanee, W.; Roy, D.; Liu, D.; De Tacconi, N.R.; Dennis, B.H.; Rajeshwar, K. Electro- and Photoreduction of Carbon Dioxide: The Twain Shall Meet at Copper Oxide/Copper Interfaces. ACS Energy Lett. 2016, 1, 332–338. [Google Scholar] [CrossRef] [Green Version]
  8. Sánchez, O.G.; Birdja, Y.Y.; Bulut, M.; Vaes, J.; Breugelmans, T.; Pant, D. Recent Advances in Industrial CO2 Electroreduction. Curr. Opin. Green Sustain. Chem. 2019, 16, 47–56. [Google Scholar] [CrossRef]
  9. Benson, E.E.; Kubiak, C.P.; Sathrum, A.J.; Smieja, J.M. Electrocatalytic and Homogeneous Approaches to Conversion of CO2 to Liquid Fuels. Chem. Soc. Rev. 2009, 38, 89–99. [Google Scholar] [CrossRef]
  10. Lu, Q.; Jiao, F. Electrochemical CO2 Reduction: Electrocatalyst, Reaction Mechanism, and Process Engineering. Nano Energy 2016, 29, 439–456. [Google Scholar] [CrossRef] [Green Version]
  11. Jiwanti, P.K.; Sultana, S.; Wicaksono, W.P.; Einaga, Y. Metal Modified Carbon-Based Electrode for CO2 Electrochemical Reduction: A Review. J. Electroanal. Chem. 2021, 898, 115634. [Google Scholar] [CrossRef]
  12. Muzyka, K.; Sun, J.; Fereja, T.H.; Lan, Y.; Zhang, W.; Xu, G. Boron-Doped Diamond: Current Progress and Challenges in View of Electroanalytical Applications. Anal. Methods 2019, 11, 397–414. [Google Scholar] [CrossRef]
  13. Xu, J.; Einaga, Y. Effect of Sp2 Species in a Boron-Doped Diamond Electrode on the Electrochemical Reduction of CO2. Electrochem. Commun. 2020, 115, 106731. [Google Scholar] [CrossRef]
  14. Ivandini, T.A.; Einaga, Y. Polycrystalline Boron-Doped Diamond Electrodes for Electrocatalytic and Electrosynthetic Applications. Chem. Commun. 2017, 53, 1338–1347. [Google Scholar] [CrossRef]
  15. Jiwanti, P.K.; Natsui, K.; Einaga, Y. The Utilization of Boron-Doped Diamond Electrodes for the Electrochemical Reduction of CO2: Toward the Production Compounds with a High Number of Carbon Atoms. Electrochemistry 2019, 2, 109–113. [Google Scholar] [CrossRef] [Green Version]
  16. Jiwanti, P.K.; Aritonang, R.P.; Abdullah, I.; Einaga, Y.; Ivandini, T.A. Copper-Nickel-Modified Boron-Doped Diamond Electrode for CO2 Electrochemical Reduction Application: A Preliminary Study. Makara J. Sci. 2019, 23, 204–209. [Google Scholar] [CrossRef] [Green Version]
  17. Ikemiya, N.; Natsui, K.; Nakata, K.; Einaga, Y. Long-Term Continuous Conversion of CO2 to Formic Acid Using Boron-Doped Diamond Electrodes. ACS Sustain. Chem. Eng. 2018, 6, 8108–8112. [Google Scholar] [CrossRef]
  18. Natsui, K.; Iwakawa, H.; Ikemiya, N.; Nakata, K.; Einaga, Y. Stable and Highly Efficient Electrochemical Production of Formic Acid from Carbon Dioxide Using Diamond Electrodes. Angew. Chem. Int. Ed. 2018, 57, 2639–2643. [Google Scholar] [CrossRef]
  19. Low, J.; Zhang, L.; Tong, T.; Shen, B.; Yu, J. TiO2/MXene Ti3C2 Composite with Excellent Photocatalytic CO2 Reduction Activity. J. Catal. 2018, 361, 255–266. [Google Scholar] [CrossRef]
  20. Qin, J.X.; Yang, X.G.; Lv, C.F.; Li, Y.Z.; Liu, K.K.; Zang, J.H.; Yang, X.; Dong, L.; Shan, C.X. Nanodiamonds: Synthesis, Properties, and Applications in Nanomedicine. Mater. Des. 2021, 210, 110091. [Google Scholar] [CrossRef]
  21. Li, N.; Chen, X.; Ong, W.J.; Macfarlane, D.R.; Zhao, X.; Cheetham, A.K.; Sun, C. Understanding of Electrochemical Mechanisms for CO2 Capture and Conversion into Hydrocarbon Fuels in Transition-Metal Carbides (MXenes). ACS Nano 2017, 11, 10825–10833. [Google Scholar] [CrossRef]
  22. Jiwanti, P.K.; Alfaza, A.M.; Kadja, G.T.M.; Khalil, M.; Einaga, Y. Electrochemical Study of CO2 Reduction on Ti3C2Tx Modified Boron-Doped Diamond Electrode. Inorg. Chem. Commun. 2022, 137, 109228. [Google Scholar] [CrossRef]
  23. Sagita, F.; Radiman, C.L.; Ledyastuti, M.; Khalil, M.; Kadja, G.T.M. Salt-Modified MXene Membrane for Ultrafast and Efficient Cationic and Anionic Dyes Removal. J. Water Process Eng. 2022, 49, 103133. [Google Scholar] [CrossRef]
  24. Jiwanti, P.K.; Natsui, K.; Nakata, K.; Einaga, Y. The electrochemical production of C2/C3 species from carbon dioxideon copper-modified boron-doped diamond electrodes. Electrochim. Acta 2018, 266, 414–419. [Google Scholar] [CrossRef]
  25. Kim, H.Y.; Choia, I.; Ahna, S.H.; Hwanga, S.J.; Yooa, S.J.; Hana, J.; Kimb, J.; Park, H.; Janga, J.H.; Kim, S.K. Analysis on the effect of operating conditions onelectrochemical conversion of carbon dioxide toformic acid. Int. J. Hydrogen Energy 2014, 339, 16506–16512. [Google Scholar] [CrossRef]
  26. Sarycheva, A.; Gogotsi, Y. Raman Spectroscopy Analysis of the Structure and Surface Chemistry of Ti3C2T XMXene. Chem. Mater. 2020, 32, 3480–3488. [Google Scholar] [CrossRef]
  27. Tang, J.; Mathis, T.S.; Kurra, N.; Sarycheva, A.; Xiao, X.; Hedhili, M.N.; Jiang, Q.; Alshareef, H.N.; Xu, B.; Pan, F.; et al. Tuning the Electrochemical Performance of Titanium Carbide MXene by Controllable In Situ Anodic Oxidation. Angew. Chem. 2019, 131, 18013–18019. [Google Scholar] [CrossRef] [Green Version]
  28. Jiwanti, P.K.; Natsui, K.; Nakata, K.; Einaga, Y. Selective Production of Methanol by the Electrochemical Reduction of CO2 on Boron-Doped Diamond Electrodes in Aqueous Ammonia Solution. RSC Adv. 2016, 6, 102214–102217. [Google Scholar] [CrossRef]
  29. 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 2015, 119, 55–61. [Google Scholar] [CrossRef]
  30. 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, 1–19. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a,b) SEM images, (c) X-ray diffractogram, and (d) Raman spectra of the as-prepared MXene (Ti3C2Tx).
Figure 1. (a,b) SEM images, (c) X-ray diffractogram, and (d) Raman spectra of the as-prepared MXene (Ti3C2Tx).
Energies 16 04537 g001
Figure 2. CV of the bare BDD and MXene-BDD electrodes in 0.1 M H2SO4 solution at a scan rate of 100 mV/s within a potential range of −1.0 V to +1.5 V (vs. Ag/AgCl).
Figure 2. CV of the bare BDD and MXene-BDD electrodes in 0.1 M H2SO4 solution at a scan rate of 100 mV/s within a potential range of −1.0 V to +1.5 V (vs. Ag/AgCl).
Energies 16 04537 g002
Figure 3. SEM image of (a) bare BDD and MXene-BDD with MXene concentrations of (b) 0.5, (c) 1.0, and (d) 2.0 mg/mL.
Figure 3. SEM image of (a) bare BDD and MXene-BDD with MXene concentrations of (b) 0.5, (c) 1.0, and (d) 2.0 mg/mL.
Energies 16 04537 g003
Figure 4. (A) pH measurement and (B) voltammograms after N2 and CO2 aerations in 0.5 M KCl, with a scan rate of 100 mV/s over a potential range of −1.7 to 0.7 V (vs. Ag/AgCl). Working electrode was bare BDD.
Figure 4. (A) pH measurement and (B) voltammograms after N2 and CO2 aerations in 0.5 M KCl, with a scan rate of 100 mV/s over a potential range of −1.7 to 0.7 V (vs. Ag/AgCl). Working electrode was bare BDD.
Energies 16 04537 g004
Figure 5. LSV of CO2 reduction activity on BDD and various MXene-BDD electrodes in 0.5 M KCl, with a scan rate of 100 mV/s.
Figure 5. LSV of CO2 reduction activity on BDD and various MXene-BDD electrodes in 0.5 M KCl, with a scan rate of 100 mV/s.
Energies 16 04537 g005
Figure 6. Concentration of HCOOH obtained from the CO2 electrochemical reduction with various MXene concentrations at a potential of −1.7 V (vs. Ag/AgCl). Reduction time was 1 h.
Figure 6. Concentration of HCOOH obtained from the CO2 electrochemical reduction with various MXene concentrations at a potential of −1.7 V (vs. Ag/AgCl). Reduction time was 1 h.
Energies 16 04537 g006
Figure 7. The HCOOH product concentration and faradaic efficiency of CO2 electrochemical reduction on bare BDD and MXene-BDD electrodes at various reduction potentials of −1.3 V, −1.5 V, and −1.7 V. The electrochemical reductions were performed in 0.5 M KCl for 1 h after N2 and CO2 gas aerations.
Figure 7. The HCOOH product concentration and faradaic efficiency of CO2 electrochemical reduction on bare BDD and MXene-BDD electrodes at various reduction potentials of −1.3 V, −1.5 V, and −1.7 V. The electrochemical reductions were performed in 0.5 M KCl for 1 h after N2 and CO2 gas aerations.
Energies 16 04537 g007
Figure 8. CV of MXene-BDD 2.0 before and after CO2 reduction applications in 0.5 M KCl. Scan rate of 100 mV/s.
Figure 8. CV of MXene-BDD 2.0 before and after CO2 reduction applications in 0.5 M KCl. Scan rate of 100 mV/s.
Energies 16 04537 g008
Table 1. Elemental analysis of various concentrations of MXene modified on the surface of the BDD electrode by EDX characterization.
Table 1. Elemental analysis of various concentrations of MXene modified on the surface of the BDD electrode by EDX characterization.
Electrode% Weight Load
COTi
Bare BDD99.770.230
MXene-BDD 0.594.405.470.12
MXene-BDD 1.074.8417.104.06
MXene-BDD 2.074.9717.897.14
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jiwanti, P.K.; Alfaza, A.M.; Kadja, G.T.M.; Natalya, S.A.C.; Sagita, F.; Einaga, Y.; Purwaningsih, A.; Amalina, I.; Rizki, I.N. Enhancement of the Catalytic Effect on the Electrochemical Conversion of CO2 to Formic Acid Using MXene (Ti3C2Tx)-Modified Boron-Doped Diamond Electrode. Energies 2023, 16, 4537. https://doi.org/10.3390/en16124537

AMA Style

Jiwanti PK, Alfaza AM, Kadja GTM, Natalya SAC, Sagita F, Einaga Y, Purwaningsih A, Amalina I, Rizki IN. Enhancement of the Catalytic Effect on the Electrochemical Conversion of CO2 to Formic Acid Using MXene (Ti3C2Tx)-Modified Boron-Doped Diamond Electrode. Energies. 2023; 16(12):4537. https://doi.org/10.3390/en16124537

Chicago/Turabian Style

Jiwanti, Prastika Krisma, Asmaul Mashad Alfaza, Grandprix T. M. Kadja, Suci A. C. Natalya, Fuja Sagita, Yasuaki Einaga, Aning Purwaningsih, Ilma Amalina, and Intan Nurul Rizki. 2023. "Enhancement of the Catalytic Effect on the Electrochemical Conversion of CO2 to Formic Acid Using MXene (Ti3C2Tx)-Modified Boron-Doped Diamond Electrode" Energies 16, no. 12: 4537. https://doi.org/10.3390/en16124537

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