Next Article in Journal
Hispolon Cyclodextrin Complexes and Their Inclusion in Liposomes for Enhanced Delivery in Melanoma Cell Lines
Next Article in Special Issue
Synthesis of Co3O4 Nanoparticles-Decorated Bi12O17Cl2 Hierarchical Microspheres for Enhanced Photocatalytic Degradation of RhB and BPA
Previous Article in Journal
Nanostrategies: The Future Medicine for Fighting Cancer Progression and Drug Resistance 2.0
Previous Article in Special Issue
Rhombohedral/Cubic In2O3 Phase Junction Hybridized with Polymeric Carbon Nitride for Photodegradation of Organic Pollutants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nitrogen-Doped Bismuth Nanosheet as an Efficient Electrocatalyst to CO2 Reduction for Production of Formate

1
State Key Laboratory of High-Efficiency Utilization of Coal and Green Chemical Engineering, College of Chemistry and Chemical Engineering, Ningxia University, Yinchuan 750021, China
2
Key Laboratory of Ministry of Education for Protection and Utilization of Special Biological Resources in Western China, School of Life Sciences, Ningxia University, Yinchuan 750021, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2022, 23(22), 14485; https://doi.org/10.3390/ijms232214485
Submission received: 3 November 2022 / Revised: 17 November 2022 / Accepted: 18 November 2022 / Published: 21 November 2022

Abstract

:
Electrochemical CO2 reduction (CO2RR) to produce high value-added chemicals or fuels is a promising technology to address the greenhouse effect and energy challenges. Formate is a desirable product of CO2RR with great economic value. Here, nitrogen-doped bismuth nanosheets (N-BiNSs) were prepared by a facile one-step method. The N-BiNSs were used as efficient electrocatalysts for CO2RR with selective formate production. The N-BiNSs exhibited a high formate Faradic efficiency (FEformate) of 95.25% at −0.95 V (vs. RHE) with a stable current density of 33.63 mA cm−2 in 0.5 M KHCO3. Moreover, the N-BiNSs for CO2RR yielded a large current density (300 mA cm−2) for formate production in a flow-cell measurement, achieving the commercial requirement. The FEformate of 90% can maintain stability for 14 h of electrolysis. Nitrogen doping could induce charge transfer from the N atom to the Bi atom, thus modulating the electronic structure of N-Bi nanosheets. DFT results demonstrated the N-BiNSs reduced the adsorption energy of the *OCHO intermediate and promoted the mass transfer of charges, thereby improving the CO2RR with high FEformate. This study provides a valuable strategy to enhance the catalytic performance of bismuth-based catalysts for CO2RR by using a nitrogen-doping strategy.

1. Introduction

With rapid economic development and the extensive use of fossil fuels, carbon dioxide (CO2) in the atmosphere continues to increase. Excessive amounts of CO2 emissions will cause severe global warming and sea levels to rise. Therefore, the necessary measures to reduce the environmental impact of CO2 are of great importance. Among the various existing CO2 conversion technologies, the electrochemical CO2 reduction reaction (CO2RR) is a prospective strategy to produce fuels or value-added chemicals. Electrochemical CO2RR can convert CO2 into CO [1], formate [2], methanol [3], ethanol [4,5], and hydrocarbons [6,7,8] under room temperature and atmospheric pressure. Among the CO2 reduction products, formate has attracted great interest because of its commercial value. Mainly, formate is an essential feedstock for the pharmaceutical and chemical industries. Therefore, it is significant to develop an efficient way to produce formate.
The practical application of electrochemical CO2RR to formate production is constrained owing to the following factors, such as high overpotential, colossal cost, low selectivity, and poor stability [9]. In addition, since electrochemical CO2RR usually occurs in an aqueous solution, it will be accompanied by a hydrogen evolution reaction (HER), resulting in low Faradaic efficiency (FE) of the product. So far, many metal-based catalysts, including Pb [10], In [11], Sb [12,13], Sn [14], Co [15], Cu [16,17], Bi [18], and Pd [19] were designed and developed to improve the efficiency and selectivity of CO2RR to produce the formate in an aqueous solution. Bi-based catalysts are popular because of the selective production of formate, low cost, and ability to inhibit the occurrence of adverse reaction HER in an aqueous solution [20]. In addition, Bi-based catalysts tend to stabilize *OCHO, an essential intermediate in formate formation [21]. However, the catalytic efficiency and selectivity of Bi-based nanomaterials are far from meeting the requirements of CO2RR [22]. Therefore, there is an urgent need to develop effective and straightforward pathways to improve catalytic efficiency and selectivity. As demonstrated recently, plenty of studies have exhibited that the properties of CO2RR were improved by changing the structure of the catalyst, such as size and morphology [23,24,25]. Two-dimensional (2D) nanostructures with high conductivity and abundant active sites have been extensively studied as potential electrocatalysts, which hold a key to enhanced CO2RR activity [26,27]. Meanwhile, the FE of target products was improved via surface modification, doping hetero-atoms, and defects [28,29]. Especially for doping hetero-atoms, such as N, S, or O, this tactic would optimize the electronic structure of the catalyst and provide more active sites. For instance, Wu et al. [30] used this strategy to prepare the nitrogen-doped Sn(S) nanosheets. It can reach the highest FE of formate of 93.3% at −0.7 V versus reversible hydrogen electrode (vs. RHE). Liu et al. [31] prepared a Bi2O3-NGQDs catalyst and achieved the FE of formate was nearly 100% at −0.71 V (vs. RHE). Besides, Kang et al. [32] prepared nitrogen-doped SnO2/C material that could enhance electrocatalytic CO2RR activity, showing high FE (90%) for formate at –0.65 V vs. RHE. Therefore, doping hetero-atoms can increase the amounts of active sites and optimize the electronic structure to achieve high formate current density and selectivity.
Herein, we developed a one-step activation and nitrogen-doping combination method using bismuth citrate as the Bi precursor, and the combination of Ca(OH)2 and NH4Cl as an activator and a nitrogen source to prepare the nitrogen-doped bismuth nanosheets (N-BiNSs). The N-BiNSs can be used to electrocatalyze CO2RR to produce formate with high efficiency. N-doping improved the electrical conductivity, which will make up for the poor conductivity of bismuth-based materials. The theoretical study shows that the outstanding selectivity could be attributed to the change in the electronic structure of bismuth after doping N atoms. At the same time, N-BiNSs reduced the adsorption energy of *OCHO intermediate and promoted the mass transfer of charge. The optimal adsorption *OCHO intermediate promoted formate formation while inhibiting the CO product pathway, thereby enhancing the selectivity of CO2RR for formate.

2. Results and Discussion

2.1. Morphology and Structure Analysis

In this work, nitrogen-doped bismuth nanosheets (N-BiNSs) were prepared by a simple and rapid one-pot method. A series of N-BiNSs-X was successfully constructed by adjusting the amount of Ca(OH)2 and NH4Cl in different proportions to provide an activating agent and a nitrogen source (Figure 1). To study the phase composition of N-BiNSs, the catalyst was determined by X-ray diffraction (XRD) measurement. It can be drawn from the XRD pattern in Figure 2a, the N-BiNSs exhibited different peaks at 27.12°, 37.95°, and 39.62°, which were indexed to the (012), (104), and (110) planes of tripartite crystal system Bi (PDF#85-1329). Compared with BiNSs catalyst, the N-BiNSs had stronger (110) and (012) crystal planes. The morphologies of the BiNSs and N-BiNSs were studied by scanning electron microscopy (SEM), as shown in Supplementary Figure S1a,c and Figure 2b. It can be seen that the BiNSs showed the stacked flakiness morphology of bulk Bi (Supplementary Figure S1c), while the N-BiNSs showed a uniform 2D nanosheet structure with a large thin layer area (Figure 2b). This 2D nanostructure was very beneficial for increasing surface area and abundant active sites. Additionally, the transmission electron microscopy (TEM) observations (Figure 2c) revealed the ultrathin feature of N-BiNSs. These nanosheets were ubiquitously present. Besides, the high-resolution transmission microscopic (HRTEM) images of N-BiNSs display a lattice spacing of about 0.272 nm that corresponds to the (110) crystal plane of the tripartite crystal system Bi. The lattice band gap of the material was very uniform. As shown in Figure 2e, the high-angle annular dark field scanning TEM (HAADF-STEM) images also exhibited the morphology of nanosheets. Moreover, the N-BiNSs structure was confirmed again by the dispersive energy X-ray (EDX) elemental mapping. The Bi (red) and N (green) elements were well distributed on the whole surface of N-BiNSs (Figure 2f–h). Meanwhile, the EDS elemental mapping and line scan (Supplementary Figure S1b) were first recorded to demonstrate Bi and N distribution on N-Bi nanosheets accompanied by a disparate atomic ratio of 97.85% (Bi) and 2.85% (N). The above results suggested that nitrogen had been triumphantly incorporated into the BiNSs catalyst.
X-ray photoelectron spectroscopy (XPS) was also used to characterize the surface chemical constituents of the specimens and the valence state of the nanomaterials. The position of the C(sp2) peak in the C1s spectrum was taken as the reference value of 284.8 eV, and the obtained XPS spectrum was calibrated. Figure 3a,b reveals the detailed Bi 4f spectra by comparing the XPS spectra of BiNSs and N-BiNSs. For the BiNSs catalyst, binding energies at 165 eV and 159.6 eV were belong to the Bi3+ 4f5/2 and 4f7/2 [23], respectively. However, binding energies for Bi3+ 4f5/2 and 4f7/2 in N-BiNSs shifted to 164.5 and 159.11 eV, respectively. These could be caused by the charge transfer from the N to the Bi atom, thus optimizing the electronic structure of N-BiNSs. Furthermore, the N1s peaks of N-BiNSs can be split into three peaks (Figure 3c), pyridinic-N (397.6 eV), N-oxidized (404.6 eV), and Nitrate (405.4 eV) [12,33,34,35], respectively. XPS analyses showed that the atomic ratio of N was 0.41% and pyridinic N accounted for the largest proportion, it indicated that the N element was doped successfully in BiNSs.

2.2. Electrocatalytic Perfomance

To estimate the CO2 reduction performance of BiNSs, N-BiNSs, N-BiNSs-1, and N-BiNSs-2, the catalytic reactions are applied in a proton exchange membrane (PEM) separated two-compartment cell. Linear sweep voltammetry (LSV) curves (Supplementary Figure S3) of N-BiNSs were measured in both CO2-saturated and Ar-saturated 0.5 M KHCO3 electrolytes. The N-BiNSs catalyst exhibited a larger current in the CO2-saturated electrolyte than in N2. The current density increases sharply from −0.6 V (vs. RHE), reaching about 30 mA cm−2 at −0.8 V (vs. RHE) to about 50 mA cm−2 at −1.2 V (vs. RHE). The polarization curves stated that the catalyst has higher activity to CO2RR. Meanwhile, the LSV curves (Figure 4a) were tested in the CO2-saturated 0.5 M KHCO3 electrolyte. Strikingly, for the N-BiNSs catalyst, the current density in the CO2-saturated atmosphere was higher than the BiNSs catalysts, indicating the introduction of N could improve the electrocatalytic activity of Bi on CO2RR. Moreover, comparing the three samples with different N doping amounts, the N-BiNSs show the highest current with the optimized Nitrogen amount of 2.85%.
To evaluate the selectivity of CO2RR, electrolysis was performed in a CO2-saturated 0.5 M KHCO3 aqueous solution at various applied potentials. The gaseous and liquid products were quantitatively analyzed via gas chromatography (GC) and ion chromatography (IC) (Supplementary Figure S2), respectively. As shown in Figure 4b, the FEformate of N-BiNSs was higher than BiNSs, N-BiNSs-1, and N-BiNSs-2 at all applied potentials. To investigate the catalytic activity of N-BiNSs, the CO2RR was carried out at different potentials between −0.4 V and −1.3 V (vs. RHE). As the applied potential changes, the FE of the CO2RR products was displayed in Figure 4c. The N-BiNSs also showed that formate was the main product of CO2RR. The FE value of CO was lower than 20% in the whole potential window, and the FE value of H2 decreased significantly from 60% at −0.5 V to 2–3% at −1.0 V (vs. RHE), the FEformate at −0.95 V (vs. RHE) was 95.25%. In contrast, for the BiNSs without nitrogen element, as shown in Supplementary Figure S4a, the FE values of formate, CO, and H2 at −0.95 V (vs. RHE) were 81.54%, 3.09%, and 10.56%, respectively. This indicated that the nitrogen-doped bismuth catalyst was beneficial to improve the selectivity of formate. Meanwhile, in the N-BiNSs-1 catalyst with a small amount of nitrogen, the FEformate could reach 88.94% at −0.95 V (vs. RHE) in Supplementary Figure S4b. For the catalyst N-BiNSs-2 with a higher amount of doping nitrogen, the maximum FE of formate in the H-shaped cell reached 78.63% in Supplementary Figure S4c. It could be concluded that varying N doping amounts lead to the formation of different catalyst morphologies, resulting in different activities. Supplementary Figure S4d showed the constant potential electrolysis of CO2 under a series of potentials. The stable current density showed that the N-BiNSs catalyst had good electrochemical stability in the CO2RR test. Furthermore, the formate partial current densities (jformate) of the N-BiNSs were measured in the whole potential region in Figure 4d, with a maximum value of jformate = 45 mA cm−2 at −1.2 V (vs. RHE). To better understand the activity and kinetics of N-BiNSs materials for CO2RR, the Tafel diagram of the catalyst was obtained in the low current density area. In Supplementary Figure S5a, the slope value obtained by the N-BiNSs catalyst was 184.13 mV dec−1, indicating that the electron transfer rate of this catalyst was relatively fast, thus facilitating the adsorption and desorption of CO2* intermediate on the N-BiNSs catalyst surface [36].
More importantly, the CO2RR activity was also related to the electrochemical active surface area (ECSA) of the catalyst. To evaluate the ECSA of BiNSs, N-BiNSs, N-BiNSs-1, and N-BiNSs-2, the double-layer capacitance (Cdl) was calculated. According to the cyclic voltammograms (CVs) of BiNSs, N-BiNSs, N-BiNSs-1, and N-BiNSs-2 at different sweep speeds in the potential region of 0.12 V–0.22 V (vs. RHE) (Supplementary Figure S6a–d). It can be seen from Figure 4e and Supplementary Figure S5c that the capacitance values of BiNSs, N-BiNSs, N-BiNSs-1, and N-BiNSs-2 were 0.43 mF cm−2, 22.2 mF cm−2, 0.56 mF cm−2, and 0.075 mF cm−2, respectively. These results indicated that the Cdl of N-BiNSs electrocatalyst was highest, which can provide abundant catalytic active sites for increasing the electrocatalytic performance of CO2RR. At the same time, we measured the impedance of some different catalysts at open circuit voltage and obtained the Nyquist diagram (Supplementary Figure S5b) to explore the kinetic processes among the catalyst interfaces. The N-BiNSs material corresponds to the most minor semicircle among the three. The results showed that the interfacial charge could be transferred rapidly to improve the catalytic activity in the reaction process, which was consistent with our inferred results.
We further studied the long-term durability of the material for CO2RR, as shown in Figure 4f. The material was tested for electrolysis at −0.95 V (vs. RHE) for about 20 h, and the FE of formate remained very stable, basically around 95%, indicating that the N-BiNSs material had significant durability to CO2RR. A comparison of the performance of N-Bi nanosheets in electrocatalytic CO2RR with other representative electrocatalysts is in the literature (Supplementary Table S1). It was noteworthy that the N-BiNSs catalyst showed the morphology of the nanosheets was maintained and became thinner after a period of electrolysis process (Supplementary Figure S7a). The crystal shape of the nanosheets remained after prolonged electrolysis (Supplementary Figure S7b). It was proved that the catalyst had excellent morphology stability.
To eliminate CO2 mass transfer constraints in the H-cell and achieve a commercially viable high current density (≈200 mA cm−2), a flow cell reactor was assembled using catalysts coated on the gas diffusion layer, carbon paper of 2 × 3 cm size, and commercial mercury oxide anion exchange membrane (Supplementary Figure S8). In this unit, carbon dioxide gas can react at the gas-liquid-solid three-phase boundary. Peristaltic pumps and gas-liquid mixed flow pumps are installed on the cathode and anode to remove liquid or gas products, keeping the pH of the electrolyte constant and fully contacting the electrode surface. Then, we systematically evaluated the CO2RR performance of the N-BiNSs catalyst on a carbon-based gas diffusion layer (C-GDL) substrate, as shown in Figure 5a,b. 1.0 M KOH aqueous solution was used as a flow-electrolyte. The alkaline electrolyte not only effectively inhibits HER, but also effectively reduces the activation energy barrier. Through the LSV curve, it can be displayed very clearly that the current density of the catalyst has met the commercial requirements, as shown in Figure 5a. Notably, the catalyst was maintained for more than 14 h at a high current density of ~300 mA cm−2 at a potential of −1.2 V (vs. RHE) in Figure 5b. The catalyst had significant durability at high current density, with the average selectivity of the catalyst being about 89.30%. After 14 h of electrolysis, due to the existing structure of the flow tank, flooding, seepage, and other problems occurred. Therefore, the optimization of the flow tank structure is still an urgent problem to be solved.
The two-electrode electrolyzer was assembled by utilizing the N-BiNSs loaded carbon paper as a cathode for CO2RR and IrO2 loaded on carbon paper as an anode for the oxygen evolution reaction (OER). As shown in Figure 5c, the polarization curve of the N-BiNSs||IrO2 cell exhibits electrocatalytic performance, with a current density of 6.19 mA cm−2 at 3.0 V, capable of delivering 10 mA cm−2 at 3.4V. It was worth noting that the initial current density was maintained at 6 mA cm−2. After 11 h of continuous reaction, the current density was maintained at 5.4 mA cm−2, the current density decreased by about 9.3%, and the stability was pretty, as shown in Figure 5d. The OER electrocatalyst IrO2 may also be replaced by non-noble metal-based materials [37].

2.3. Catalytic Mechanisms Revealed by DFT Computations

All the above electrochemical performance studies proved that the N-BiNSs catalysts had excellent activity and selectivity compared with pure BiNSs. In order to further explore the reasons why the catalyst improves CO2RR activity and selectivity, and determine the reaction mechanism of formate, the density functional theory (DFT) calculation was used to simulate and compare the CO2RR pathway on N-BiNSs and BiNSs surfaces. Figure 6a,b depicts the optimized adsorption geometries and their energy distributions for *OCHO (intermediate to formate) on N-BiNSs and BiNSs surfaces, respectively. As shown in Figure 6b, the total reaction pathway for the electrochemical reduction of CO2 to formate was two protons, and two electrons were transferred through the *OCHO intermediate and adsorbed by HCOOH (aq) [38]. Both N-BiNSs and BiNSs displayed a great energy barrier for *OCHO formation, confirming the original proton-coupled electron transfer is the potential limiting procedure [31] and that the optimally adsorbed *OCHO intermediate promotes formate production. Obviously, on the N-BiNSs (012) surface, the calculated Gibbs free energy ΔG for the formation of *OCHO was +0.36 eV, and the ΔG for the formation of *OCHO from BiNSs was +0.65 eV. Thus, the N-Bi nanosheets had a more negative Gibbs free energy than that of the Bi nanosheets site, indicating that *OCHO formation and protonation were more spontaneous.

3. Materials and Methods

3.1. Synthesis of Nitrogen-Doped Bismuth Nanosheets (N-BiNSs)

The synthesis method was improved according to the literature [39], and the synthesis of the specific process was as follows. Firstly, 5 g of C6H5BiO7, 3.7 g of Ca(OH)2, and 5.4 g of NH4Cl were weighed and placed in a sample vial the reagents were thoroughly mixed. Then the mixed reagents were transferred to a porcelain boat and placed in a tube furnace. Subsequently, the tube furnace was vacuumed, and N2 was pumped into it. The tube furnace was activated at 450 °C for 1 h in the N2 atmosphere, and then the temperature was increased to 800 °C for 2 h. During the whole synthesis process, the heating rate was 5 °C min−1 in the tube furnace. At the end of the reaction, the samples were cooled to room temperature, the samples were first repeatedly cleared with 2.0 M hydrochloric acid (HCl), and the inorganic salts of the samples were removed entirely. The samples were repeatedly cleared with deionized water until the pH value was neutral. Finally, the samples were dried under vacuum at 60 °C for 24 h. The samples were named nitrogen-doped bismuth nanosheets (N-BiNSs). For comparison, samples prepared at different Ca(OH)2 and NH4Cl were denoted as N-BiNSs-1 and N-BiNSs-2. The N-BiNSs-1 catalyst was prepared by adding Ca(OH)2 (1.85 g) and NH4Cl (2.7 g), and the N-BiNSs-2 was prepared by adding Ca(OH)2 (5.55 g) and NH4Cl (8.1 g). Moreover, the bismuth nanosheets (BiNSs) sample was prepared on the basis of the same procedure as introduced above, but without added Ca(OH)2 and NH4Cl.

3.2. Preparation of Working Electrode

Generally, 10 mg of as-prepared N-BiNSs-X were mixed with 600 μL of deionized water, 350 μL ethanol, and 50 μL of Nafion D-521 dispersion within a 2 mL vial in an ultrasonic bath for 2 h to become a homogeneous catalyst ink suspension; 10 μL of the catalyst ink was dropped onto a carbon paper (1 mg cm−2), and the catalyst-covered electrode was dried in a desiccator before use.

3.3. Electrochemical Measurements

In this work, all electrochemical properties were tested at an electrochemical workstation (Shanghai Chenhua, CHI760E). The electrocatalytic CO2RR was performed in an H-type electrolytic cell with the two chambers isolated by a Nafion 117 proton exchange membrane to stop reoxidation of the cathode generated to the anode. The prepared catalyst, the Ag/AgCl(saturated KCl), and the platinum sheet were used as the working electrodes, reference electrodes, and counter electrodes, respectively. All potential values were measured with Ag/AgCl and then converted to RHE. All electrode potentials were converted to electrode potentials relative to RHE by the Nernst equation: E (vs. RHE) = E (vs. Ag/AgCl) + 0.0591 × pH + 0.222 V. Each electrode chamber contained 35 mL of 0.5 M KHCO3 electrolyte. Before the test, the above electrolyte was continuously bubbled with high-purity CO2 or N2 for at least 30 min to saturate the electrolyte with N2 (pH = 8.5) or CO2 (pH = 7.2). The electrochemical reduction of CO2 on BiNSs and N-BiNSs-X electrodes were performed in N2-saturated 0.5 M KHCO3 or CO2-saturated 0.5 M KHCO3 at ambient temperature and pressure. Double-layer capacitance (Cdl) was obtained from cyclic voltammetry (CV) curves surveyed at unlike scan rates (10, 30, 50, 70, 90, 110, and 130 mV s−1) in the scope of +0.12 V to +0.22 V (vs. RHE). Electrochemical impedance spectroscopy (EIS) was employed to understand the charge mass transfer resistance in the electrocatalytic CO2RR course. The frequency range was 10 kHz–1.0 Hz and the amplitude was 5 mV s−1. The durability of the catalyst was obtained by using the current-time curve method (i-t). At the end of each potential test, the liquid products generated at each potential were collected and detected by ion chromatography.

3.4. Product Analysis

Faraday efficiency test method: Control potential electrolytic Coulomb method was used for the CO2 saturated solution, and the electrolytic reduction products were analyzed and calculated 0.5 h later. The flow rate of CO2 was mastered at 20 mL min−1 during electrolysis. The liquid products after electrolysis were detected by ion chromatography (AS-DV, Thermo Scientific, New York City, MA, USA). The Faraday efficiency (FE) of liquid-phase products was calculated as follows:
FE = N n F Q × 100 %
where N is the number of electrons transferred, n is the total mole fraction of the gas measured by gas chromatography, F is the Faraday constant (96,485 C mol−1), and Q is the total electric charge passed through the electrode.
The gaseous products were quantified by gas chromatography (GC7900, Tianmei, China), and the H2 and CO were detected by a thermal conductivity detector (TCD), and flame ionization detector (FID), respectively, and ultra-pure N2 (>99.99%) was used as carrier gas. The calculation method of Faraday efficiency (FE) was as follows:
FE = n × C × v × F V m × j × 100 %
where ν is the gas flow rate of supplied CO2, C is the concentration of the gaseous product, n is the number of electrons for producing a molecule of CO or H2, F is the Faraday constant (96,485 C mol−1), Vm is the molar volume of gas at 298 K, j is the recorded current (A).

4. Conclusions

In summary, we successfully prepared nitrogen-doped bismuth nanosheets (N-BiNSs) through a simple one-step activation and nitrogen-doping connective way. The N-Bi nanosheets exhibited very high activity, selectivity, and stability for formate production to BiNSs. The results showed that the selectivity of formate was 95.25% at −0.95 V (vs. RHE). Meanwhile, the N-BiNSs catalyst also showed excellent stability, and no apparent catalyst deactivation occurred after 20 h of electrolysis. Importantly, we also optimize the CO2RR activity by flow-cell. The N-BiNSs catalyst possessed high durability at a current density of 300 mA cm−2 at a potential of -1.2 V (vs. RHE), and the average FE of formate was about 90.30%. The DFT simulation suggested that compared with other catalysts, N-BiNSs had the active site to adsorb CO2 more easily. The free energy barrier for forming the critical intermediate *OCHO was reduced by nitrogen doping. This electrocatalyst with high catalytic activity, durability, and selectivity has great potential to improve the technical and economic feasibility of CO2 to formate conversion. In addition, this study provides a reasonable catalyst design for CO2RR studies by proposing a simple method for synthesizing nanostructured catalysts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms232214485/s1. References [40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59] are cited in Supplementary Materials.

Author Contributions

Conceptualization, J.P.; methodology, S.L., Y.K. and J.P.; formal analysis, S.L., Y.K. and C.M.; investigation, S.L., Y.K. and C.M.; resources, H.M. and J.P.; data curation, S.L., and Y.K.; writing—original draft preparation, S.L., C.M., Y.P., H.M. and J.P.; writing—review and editing, S.L., Y.P. and J.P.; visualization, S.L. and K.Y; project administration, J.P.; funding acquisition, Y.P. and J.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the support from the National Natural Science Foundation of China (No. 22262027), Ningxia leading scientific and technological innovation talents projects (No. KJT2018002), Ningxia Natural Science Foundation (No. 2022AAC03103, 2022AAC03113), and National First-rate Discipline Construction Project of Ningxia (NXYLXK2017A04). This paper was also supported by College Students’ Innovative and Entrepreneurship Training Program of Ningxia University (No. S202210749054).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data in this study can be found in public data bases and Supplementary Materials, as described in the Material and Methods section (Section 3).

Conflicts of Interest

The authors declared that there was no competing financial interest.

References

  1. Kumar, B.; Asadi, M.; Pisasale, D.; Sinha-Ray, S.; Rosen, B.A.; Haasch, R.; Abiade, J.; Yarin, A.L.; Salehi-Khojin, A. Renewable and metal-free carbon nanofibre catalysts for carbon dioxide reduction. Nat. Commun. 2013, 4, 2819. [Google Scholar] [CrossRef] [Green Version]
  2. Feng, X.; Zou, H.; Zheng, R.; Wei, W.; Wang, R.; Zou, W.; Lim, G.; Hong, J.; Duan, L.; Chen, H. Bi2O3/BiO2 Nanoheterojunction for Highly Efficient Electrocatalytic CO2 Reduction to Formate. Nano Lett. 2022, 22, 1656–1664. [Google Scholar] [CrossRef] [PubMed]
  3. Boutin, E.; Wang, M.; Lin, J.C.; Mesnage, M.; Mendoza, D.; Lassalle-Kaiser, B.; Hahn, C.; Jaramillo, T.F.; Robert, M. Aqueous Electrochemical Reduction of Carbon Dioxide and Carbon Monoxide into Methanol with Cobalt Phthalocyanine. Angew. Chem. Int. Ed. 2019, 58, 16172–16176. [Google Scholar] [CrossRef] [PubMed]
  4. Guo, C.; Guo, Y.; Shi, Y.; Lan, X.; Wang, Y.; Yu, Y.; Zhang, B. Electrocatalytic Reduction of CO2 to Ethanol at Close to Theoretical Potential via Engineering Abundant Electron-Donating Cu(delta+) Species. Angew. Chem. Int. Ed. 2022, 134, e202205909. [Google Scholar] [CrossRef]
  5. Liu, B.; Yao, X.; Zhang, Z.; Li, C.; Zhang, J.; Wang, P.; Zhao, J.; Guo, Y.; Sun, J.; Zhao, C. Synthesis of Cu2O Nanostructures with Tunable Crystal Facets for Electrochemical CO2 Reduction to Alcohols. ACS Appl. Mater. Interfaces 2021, 13, 39165–39177. [Google Scholar] [CrossRef]
  6. Gao, Y.; Wu, Q.; Liang, X.; Wang, Z.; Zheng, Z.; Wang, P.; Liu, Y.; Dai, Y.; Whangbo, H.M.; Huang, B. Cu2O Nanoparticles with Both {100} and {111} Facets for Enhancing the Selectivity and Activity of CO2 Electroreduction to Ethylene. Adv. Sci. 2020, 7, 1902820. [Google Scholar] [CrossRef] [Green Version]
  7. Jung, H.; Lee, S.Y.; Lee, C.W.; Cho, M.K.; Won, D.H.; Kim, C.; Oh, H.S.; Min, B.K.; Hwang, Y.J. Electrochemical Fragmentation of Cu2O Nanoparticles Enhancing Selective C-C Coupling from CO2 Reduction Reaction. J. Am. Chem. Soc. 2019, 141, 4624–4633. [Google Scholar] [CrossRef]
  8. Tan, X.; Yu, C.; Zhao, C.; Huang, H.; Yao, X.; Han, X.; Guo, W.; Cui, S.; Huang, H.; Qiu, J. Restructuring of Cu2O to Cu2O@Cu-Metal-Organic Frameworks for Selective Electrochemical Reduction of CO2. ACS Appl. Mater. Interfaces 2019, 11, 9904–9910. [Google Scholar] [CrossRef]
  9. Wang, Y.-H.; Jiang, W.-J.; Yao, W.; Liu, Z.-L.; Liu, Z.; Yang, Y.; Gao, L.-Z. Advances in electrochemical reduction of carbon dioxide to formate over bismuth-based catalysts. Rare Met. 2021, 40, 2327–2353. [Google Scholar] [CrossRef]
  10. 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]
  11. Yang, Y.; Fu, J.J.; Tang, T.; Niu, S.; Zhang, L.B.; Zhang, J.N.; Hu, J.S. Regulating surface In–O in In@ InOx core-shell nanoparticles for boosting electrocatalytic CO2 reduction to formate. Chin. J. Catal. 2022, 43, 1674–1679. [Google Scholar] [CrossRef]
  12. Jiang, Z.; Wang, T.; Pei, J.; Shang, H.; Zhou, D.; Li, H.; Dong, J.; Wang, Y.; Cao, R.; Zhuang, Z.; et al. Discovery of main group single Sb–N4 active sites for CO2 electroreduction to formate with high efficiency. Energy Environ. Sci. 2020, 13, 2856–2863. [Google Scholar] [CrossRef]
  13. Wen, G.; Lee, D.U.; Ren, B.; Hassan, F.M.; Jiang, G.; Cano, Z.P.; Gostick, J.; Croiset, E.; Bai, Z.; Yang, L.; et al. Orbital Interactions in Bi-Sn Bimetallic Electrocatalysts for Highly Selective Electrochemical CO2 Reduction toward Formate Production. Adv. Energy Mater. 2018, 8, 1802427. [Google Scholar] [CrossRef]
  14. Cheng, Y.; Hou, J.; Kang, P. Integrated Capture and Electroreduction of Flue Gas CO2 to Formate Using Amine Functionalized SnOx Nanoparticles. ACS Energy Lett. 2021, 6, 3352–3358. [Google Scholar] [CrossRef]
  15. Gao, S.; Jiao, X.C.; Sun, Z.T.; Zhang, W.H.; Sun, Y.F.; Wang, C.M.; Hu, Q.T.; Zu, X.L.; Yang, F.; Yang, S.Y.; et al. Ultrathin Co3O4 layers realizing optimized CO2 electroreduction to formate. Angew. Chem. Int. Ed. 2016, 55, 698–702. [Google Scholar] [CrossRef] [PubMed]
  16. Peng, L.; Wang, Y.; Wang, Y.; Xu, N.; Lou, W.; Liu, P.; Cai, D.; Huang, H.; Qiao, J. Separated growth of Bi-Cu bimetallic electrocatalysts on defective copper foam for highly converting CO2 to formate with alkaline anion-exchange membrane beyond KHCO3 electrolyte. Appl. Catal. B Environ. 2021, 288, 120003. [Google Scholar] [CrossRef]
  17. Tao, Z.X.; Wu, Z.S.; Yuan, X.L.; Wu, Y.S.; Wang, H.L. Copper–gold interactions enhancing formate production from electrochemical CO2 reduction. ACS Catal. 2019, 9, 10894–10898. [Google Scholar] [CrossRef]
  18. Fan, J.; Zhao, X.; Mao, X.; Xu, J.; Han, N.; Yang, H.; Pan, B.; Li, Y.; Wang, L.; Li, Y. Large-Area Vertically Aligned Bismuthene Nanosheet Arrays from Galvanic Replacement Reaction for Efficient Electrochemical CO2 Conversion. Adv. Mater. 2021, 33, e2100910. [Google Scholar] [CrossRef]
  19. Xie, L.; Liu, X.; Huang, F.; Liang, J.; Liu, J.; Wang, T.; Yang, L.; Cao, R.; Li, Q. Regulating Pd-catalysis for electrocatalytic CO2 reduction to formate via intermetallic PdBi nanosheets. Chin. J. Catal. 2022, 43, 1680–1686. [Google Scholar] [CrossRef]
  20. Han, N.; Wang, Y.; Yang, H.; Deng, J.; Wu, J.; Li, Y.; Li, Y. Ultrathin bismuth nanosheets from in situ topotactic transformation for selective electrocatalytic CO2 reduction to formate. Nat. Commun. 2018, 9, 1320. [Google Scholar] [CrossRef]
  21. Wu, J.; Yu, X.; He, H.; Yang, C.; Xia, D.; Wang, L.; Huang, J.; Zhao, N.; Tang, F.; Deng, L.; et al. Bismuth-Nanosheet-Based Catalysts with a Reconstituted Bi0 Atom for Promoting the Electrocatalytic Reduction of CO2 to Formate. Ind. Eng. Chem. Res. 2022, 61, 12383–12391. [Google Scholar] [CrossRef]
  22. Wu, Z.; Wu, H.; Cai, W.; Wen, Z.; Jia, B.; Wang, L.; Jin, W.; Ma, T. Engineering Bismuth-Tin Interface in Bimetallic Aerogel with a 3D Porous Structure for Highly Selective Electrocatalytic CO2 Reduction to HCOOH. Angew. Chem. Int. Ed. 2021, 60, 12554–12559. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, W.; Hu, Y.; Ma, L.; Zhu, G.; Zhao, P.; Xue, X.; Chen, R.; Yang, S.; Ma, J.; Liu, J.; et al. Liquid-phase exfoliated ultrathin Bi nanosheets: Uncovering the origins of enhanced electrocatalytic CO2 reduction on two-dimensional metal nanostructure. Nano Energy 2018, 53, 808–816. [Google Scholar] [CrossRef]
  24. Yi, L.; Chen, J.; Shao, P.; Huang, J.; Peng, X.; Li, J.; Wang, G.; Zhang, C.; Wen, Z. Molten-Salt-Assisted Synthesis of Bismuth Nanosheets for Long-term Continuous Electrocatalytic Conversion of CO2 to Formate. Angew. Chem. Int. Ed. 2020, 59, 20112–20119. [Google Scholar] [CrossRef] [PubMed]
  25. Xing, Z.; Hu, X.; Feng, X. Tuning the microenvironment in gas-diffusion electrodes enables high-rate CO2 electrolysis to formate. ACS Energy Lett. 2021, 6, 1694–1702. [Google Scholar] [CrossRef]
  26. Fan, K.; Jia, Y.; Jia, Y.F.; Ji, Y.F.; Kuang, Y.P.; Zhu, B.C.; Liu, X.Y.; Yu, J.G. Curved surface boosts electrochemical CO2 reduction to formate via bismuth nanotubes in a wide potential window. ACS Catal. 2019, 10, 358–364. [Google Scholar] [CrossRef]
  27. Huang, J.; Guo, X.; Yang, J.; Wang, L. Electrodeposited Bi dendrites/2D black phosphorus nanosheets composite used for boosting formic acid production from CO2 electroreduction. J. CO2 Util. 2020, 38, 32–38. [Google Scholar] [CrossRef]
  28. Li, W.; Bandosz, T.J. Analyzing the effect of nitrogen/sulfur groups’ density ratio in porous carbons on the efficiency of CO2 electrochemical reduction. Appl. Surf. Sci. 2021, 569, 151066. [Google Scholar] [CrossRef]
  29. Zhao, M.; Gu, Y.; Gao, W.; Cui, P.; Tang, H.; Wei, X.; Zhu, H.; Li, G.; Yan, S.; Zhang, X.; et al. Atom vacancies induced electron-rich surface of ultrathin Bi nanosheet for efficient electrochemical CO2 reduction. Appl. Catal. B: Environ. 2020, 266, 118625. [Google Scholar] [CrossRef]
  30. Cheng, H.; Liu, S.; Zhang, J.; Zhou, T.; Zhang, N.; Zheng, X.S.; Chu, W.; Hu, Z.; Wu, C.; Xie, Y. Surface Nitrogen-Injection Engineering for High Formation Rate of CO2 Reduction to Formate. Nano Lett. 2020, 20, 6097–6103. [Google Scholar] [CrossRef]
  31. Chen, Z.; Mou, K.; Wang, X.; Liu, L. Nitrogen-Doped Graphene Quantum Dots Enhance the Activity of Bi2O3 Nanosheets for Electrochemical Reduction of CO2 in a Wide Negative Potential Region. Angew. Chem. Int. Ed. 2018, 57, 12790–12794. [Google Scholar] [CrossRef] [PubMed]
  32. Li, Q.; Wang, Z.; Zhang, M.; Hou, P.; Kang, P. Nitrogen doped tin oxide nanostructured catalysts for selective electrochemical reduction of carbon dioxide to formate. J. Energy Chem. 2017, 26, 825–829. [Google Scholar] [CrossRef] [Green Version]
  33. Dong, W.; Zhang, N.; Li, S.; Min, S.; Peng, J.; Liu, W.; Zhan, D.; Bai, H. A Mn single atom catalyst with Mn–N2O2 sites integrated into carbon nanosheets for efficient electrocatalytic CO2 reduction. J. Mater. Chem. A 2022, 10, 10892–10901. [Google Scholar] [CrossRef]
  34. Guo, Y.; Yang, H.J.; Zhou, X.; Liu, K.L.; Zhang, C.; Zhou, Z.Y.; Wang, C.; Lin, W.B. Electrocatalytic reduction of CO2 to CO with 100% faradaic efficiency by using pyrolyzed zeolitic imidazolate frameworks supported on carbon nanotube networks. J. Mater. Chem. A 2017, 5, 24867–24873. [Google Scholar] [CrossRef]
  35. Varela, A.S.; Sahraie, N.R.; Steinberg, J.; Ju, W.; Oh, H.S.; Strasser, P. Metal-doped nitrogenated carbon as an efficient catalyst for direct CO2 electroreduction to CO and hydrocarbons. Angew. Chem. Int. Ed. 2015, 54, 10758–10762. [Google Scholar] [CrossRef]
  36. Kwon, I.S.; Debela, T.T.; Kwak, I.H.; Seo, H.W.; Park, K.; Kim, D.; Yoo, S.J.; Kim, J.-G.; Park, J.; Kang, H.S. Selective electrochemical reduction of carbon dioxide to formic acid using indium–zinc bimetallic nanocrystals. J. Mater. Chem. A 2019, 7, 22879–22883. [Google Scholar] [CrossRef]
  37. Tian, W.L.; Zhang, J.; Feng, H.; Wen, H.; Sun, X.; Guan, X.; Zheng, D.C.; Liao, J.; Yan, M.L.; Yao, Y.D. Fuels, A hierarchical CoP@ NiCo-LDH nanoarray as an efficient and flexible catalyst electrode for the alkaline oxygen evolution reaction. Sustain. Energy Fuels 2021, 5, 391–395. [Google Scholar] [CrossRef]
  38. Yoo, J.S.; Christensen, R.; Vegge, T.; Norskov, J.K.; Studt, F. Theoretical Insight into the Trends that Guide the Electrochemical Reduction of Carbon Dioxide to Formic Acid. ChemSusChem 2016, 9, 358–363. [Google Scholar] [CrossRef]
  39. Peng, H.; Ma, G.; Sun, K.; Mu, J.; Lei, Z. One-step preparation of ultrathin nitrogen-doped carbon nanosheets with ultrahigh pore volume for high-performance supercapacitors. J. Mater. Chem. A 2014, 2, 17297–17301. [Google Scholar] [CrossRef]
  40. Kresse, G.; Furthmuller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B Condens Matter. 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  41. Kresse, G.; Hafner, J. Ab initio molecular-dynamics simulation of the liquid-metal-amorphous-semiconductor transition in germanium. Phys. Rev. B Condens Matter. 1994, 49, 14251–14269. [Google Scholar] [CrossRef] [PubMed]
  42. Blochl, P.E. Projector augmented-wave method. Phys. Rev. B Condens Matter. 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Zhang, Y.; Yang, W. Comment on “Generalized Gradient Approximation Made Simple”. Phys. Rev Lett. 1998, 80, 890. [Google Scholar] [CrossRef]
  45. Hammer, B.; Hansen, L.B.; Nørskov, J.K. Improved adsorption energetics within density-functional theory using revised Perdew-Burke-Ernzerh of functionals. Phys. Rev. B 1999, 59, 7413–7421. [Google Scholar] [CrossRef] [Green Version]
  46. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  47. Nørskov, J.K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J.R.; Bligaard, T.; Jónsson, H. Origin of the Overpotential for Oxygen Reduction at a Fuel-Cell Cathode. J. Phys. Chem. B 2004, 108, 17886–17892. [Google Scholar] [CrossRef]
  48. Xing, Y.; Kong, X.; Guo, X.; Liu, Y.; Li, Q.; Zhang, Y.; Sheng, Y.; Yang, X.; Geng, Z.; Zeng, J. Bi@Sn Core-Shell Structure with Compressive Strain Boosts the Electroreduction of CO2 into Formic Acid. Adv Sci. 2020, 7, 1902989. [Google Scholar] [CrossRef]
  49. Li, F.; Xue, M.; Li, J.; Ma, X.; Chen, L.; Zhang, X.; MacFarlane, D.R.; Zhang, J. Unlocking the electrocatalytic activity of antimony for CO2 reduction by two-dimensional engineering of the bulk material. Angew. Chem. Int. Ed. 2017, 129, 14910–14914. [Google Scholar] [CrossRef]
  50. Watanabe, M.; Shibata, M.; Kato, A.; Azuma, M.; Sakata, T. Design of alloy electrocatalysts for CO2 reduction: III. The selective and reversible reduction of on Cu alloy electrodes. J. Electrochem. Soc. 1991, 138, 3382. [Google Scholar] [CrossRef]
  51. Sreekanth, N.; Nazrulla, M.A.; Vineesh, T.V.; Sailaja, K.; Phani, K.L. Metal-free boron-doped graphene for selective electroreduction of carbon dioxide to formic acid/formate. Chem. Commun. 2015, 51, 16061–16064. [Google Scholar] [CrossRef] [PubMed]
  52. Won, D.H.; Choi, C.H.; Chung, J.; Chung, M.W.; Kim, E.H.; Woo, S.I. Rational design of a hierarchical tin dendrite electrode for efficient electrochemical reduction of CO2. ChemSusChem 2015, 8, 3092–3098. [Google Scholar] [CrossRef] [PubMed]
  53. Zheng, X.; De Luna, P.; García de Arquer, F.P.; Zhang, B.; Becknell, N.; Ross, M.B.; Li, Y.; Banis, M.N.; Li, Y.; Liu, M.; et al. Sulfur-modulated tin sites enable highly selective electrochemical reduction of CO2 to formate. Joule 2017, 1, 794–805. [Google Scholar] [CrossRef] [Green Version]
  54. He, S.; Ni, F.; Ji, Y.; Wang, L.; Wen, Y.; Bai, H.; Liu, G.; Zhang, Y.; Li, Y.; Zhang, B.; et al. The p-Orbital delocalization of main-group metals to boost CO2 electroreduction. Angew. Chem. Int. Ed. 2018, 130, 16346–16351. [Google Scholar] [CrossRef]
  55. Kumar, B.; Atla, V.; Brian, J.P.; Kumari, S.; Nguyen, T.Q.; Sunkara, M.; Spurgeon, J.M. Reduced SnO2 porous nanowires with a high density of grain boundaries as catalysts for efficient electrochemical CO2-into-HCOOH conversion. Angew. Chem. Int. Ed. 2017, 56, 3645–3649. [Google Scholar] [CrossRef]
  56. Liang, C.; Kim, B.; Yang, S.; Yang Liu, Y.L.; Francisco Woellner, C.; Li, Z.; Vajtai, R.; Yang, W.; Wu, J.; Kenis, P.J.A.; et al. High efficiency electrochemical reduction of CO2 beyond the two-electron transfer pathway on grain boundary rich ultra-small SnO2 nanoparticles. J. Mater. Chem. A 2018, 6, 10313–10319. [Google Scholar] [CrossRef]
  57. Lee, C.W.; Hong, J.S.; Yang, K.D.; Jin, K.; Lee, J.H.; Ahn, H.-Y.; Seo, H.; Sung, N.-E.; Nam, K.T. Selective Electrochemical Production of Formate from Carbon Dioxide with Bismuth-Based Catalysts in an Aqueous Electrolyte. ACS Catal. 2018, 8, 931–937. [Google Scholar] [CrossRef]
  58. Zhang, Y.; Li, F.; Zhang, X.; Williams, T.; Easton, C.D.; Bond, A.M.; Zhang, J. Electrochemical reduction of CO2 on defect-rich Bi derived from Bi2S3 with enhanced formate selectivity. J. Mater. Chem. A 2018, 6, 4714–4720. [Google Scholar] [CrossRef]
  59. Zhao, Y.; Liang, J.; Wang, C.; Ma, J.; Wallace, G.G. Tunable and efficient tin modified nitrogen-doped carbon nanofibers for electrochemical reduction of aqueous carbon dioxide. Adv. Energy Mater. 2018, 8, 1702524. [Google Scholar] [CrossRef]
Figure 1. The illustration scheme of N-BiNSs preparation.
Figure 1. The illustration scheme of N-BiNSs preparation.
Ijms 23 14485 g001
Figure 2. (a) XRD pattern of BiNSs and N-BiNSs. Characterization of N-BiNSs: (b) SEM image, (c) TEM image, (d) HR-TEM image, (e) HAADF-STEM and (fh) EDS mapping (green and red represent N, Bi element, respectively).
Figure 2. (a) XRD pattern of BiNSs and N-BiNSs. Characterization of N-BiNSs: (b) SEM image, (c) TEM image, (d) HR-TEM image, (e) HAADF-STEM and (fh) EDS mapping (green and red represent N, Bi element, respectively).
Ijms 23 14485 g002
Figure 3. (a) XPS Bi 4f spectrum of BiNSs and (b) N-BiNSs; (c) XPS N 1s spectrum of N-BiNSs.
Figure 3. (a) XPS Bi 4f spectrum of BiNSs and (b) N-BiNSs; (c) XPS N 1s spectrum of N-BiNSs.
Ijms 23 14485 g003
Figure 4. CO2RR performances of BiNSs, N-BiNSs, N-BiNSs-1, and N-BiNSs-2: (a) Linear sweep voltammetry curves in 0.5 M KHCO3 aqueous solutions with saturated gases CO2, sweeping speed of 5 mV s−1; (b) the product FEformate at different applied potentials; (c) corresponding FE of N-BiNSs at various potentials; (d) the N-BiNSs of jformate recorded at different potentials in 0.5 M KHCO3; (e) the BiNSs and N-BiNSs the relationship between charge current density difference (ΔJ) and scanning rate; (f) stability test of N-BiNSs in the H-type electrolytic cell and Faraday efficiency test of formate production.
Figure 4. CO2RR performances of BiNSs, N-BiNSs, N-BiNSs-1, and N-BiNSs-2: (a) Linear sweep voltammetry curves in 0.5 M KHCO3 aqueous solutions with saturated gases CO2, sweeping speed of 5 mV s−1; (b) the product FEformate at different applied potentials; (c) corresponding FE of N-BiNSs at various potentials; (d) the N-BiNSs of jformate recorded at different potentials in 0.5 M KHCO3; (e) the BiNSs and N-BiNSs the relationship between charge current density difference (ΔJ) and scanning rate; (f) stability test of N-BiNSs in the H-type electrolytic cell and Faraday efficiency test of formate production.
Ijms 23 14485 g004
Figure 5. (a) The LSV curve in a flow cell under CO2 atmosphere; (b) Stability test of the material at −1.2 V vs. RHE and corresponding Faraday efficiency test of formate production of N-BiNSs; (c) Polarization curve of N-BiNSs||IrO2 couple in the two-electrode system; (d) Chronoamperometry measurement of N-BiNSs||IrO2 couple at 3 V.
Figure 5. (a) The LSV curve in a flow cell under CO2 atmosphere; (b) Stability test of the material at −1.2 V vs. RHE and corresponding Faraday efficiency test of formate production of N-BiNSs; (c) Polarization curve of N-BiNSs||IrO2 couple in the two-electrode system; (d) Chronoamperometry measurement of N-BiNSs||IrO2 couple at 3 V.
Ijms 23 14485 g005
Figure 6. (a) Side and Top views of BiNSs (012) and N-BiNSs (012) configurations (the red, white, blue, purple and brown spheres represent: O, H, N, Bi, and C atoms, respectively); (b) Calculated free-energy diagram for *OCHO generation of N-BiNSs and BiNSs for the electrochemical reduction to formate process (where * represents the active site of the catalyst).
Figure 6. (a) Side and Top views of BiNSs (012) and N-BiNSs (012) configurations (the red, white, blue, purple and brown spheres represent: O, H, N, Bi, and C atoms, respectively); (b) Calculated free-energy diagram for *OCHO generation of N-BiNSs and BiNSs for the electrochemical reduction to formate process (where * represents the active site of the catalyst).
Ijms 23 14485 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, S.; Kang, Y.; Mo, C.; Peng, Y.; Ma, H.; Peng, J. Nitrogen-Doped Bismuth Nanosheet as an Efficient Electrocatalyst to CO2 Reduction for Production of Formate. Int. J. Mol. Sci. 2022, 23, 14485. https://doi.org/10.3390/ijms232214485

AMA Style

Li S, Kang Y, Mo C, Peng Y, Ma H, Peng J. Nitrogen-Doped Bismuth Nanosheet as an Efficient Electrocatalyst to CO2 Reduction for Production of Formate. International Journal of Molecular Sciences. 2022; 23(22):14485. https://doi.org/10.3390/ijms232214485

Chicago/Turabian Style

Li, Sanxiu, Yufei Kang, Chenyang Mo, Yage Peng, Haijun Ma, and Juan Peng. 2022. "Nitrogen-Doped Bismuth Nanosheet as an Efficient Electrocatalyst to CO2 Reduction for Production of Formate" International Journal of Molecular Sciences 23, no. 22: 14485. https://doi.org/10.3390/ijms232214485

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