NC State
BioResources
Su, M., and Li, N. (2024). "The influence of amino anchoring position on SnO2/biochar for electroreduction of CO2 to HCOOH," BioResources 19(4), 9049–9059.

Abstract

Biochar derived from biomass resources as a carrier to load SnO2 for electroreduction of CO2 not only can benefit carbon emissions, but it also can achieve waste utilization. However, the weak CO2 mass transfer and conductivity ability of SnO2/biochar limits its applications to such eCO2RR processes. This study focused on modifying SnO2/biochar with amino groups and investigated the effects of positioning of amino groups on the catalyst’s physicochemical properties and electrocatalytic behaviors. Elemental analysis revealed that anchoring amino groups on biochar (SnO2/C-NHx) is advantageous for increasing amounts of amino groups attached, thereby enhancing biochar adsorption energy and subsequently increasing the loading of Sn. The CO2 adsorption curve indicated that amino groups anchored on biochar facilitate CO2 adsorption due to high specific surface area of biochar. X-ray photoelectron spectroscopy showed that amino groups anchored on SnO2 (NHx-SnO2/C) resulted in highly electron-rich centers on Sn, which promoted electron transfer between the catalyst and CO2. Electrochemical tests demonstrated the improved performance of amino-modified SnO2/biochar. SnO2/C-NHx exhibited enhanced Faraday efficiency, whereas NHx-SnO2/C showed higher current density. The disparity in electrochemical performance can be mainly attributed to the different selectivity towards rate-controlling steps of electron transfer and mass transfer induced by the various positions of amino groups anchoring.


Download PDF

Full Article

The Influence of Amino Anchoring Position on SnO2/Biochar for Electroreduction of CO2 to HCOOH

Mingxue Su a,b,* and Ning Li a,b

Biochar derived from biomass resources as a carrier to load SnO2 for electroreduction of CO2 not only can benefit carbon emissions, but it also can achieve waste utilization. However, the weak CO2 mass transfer and conductivity ability of SnO2/biochar limits its applications to such eCO2RR processes. This study focused on modifying SnO2/biochar with amino groups and investigated the effects of positioning of amino groups on the catalyst’s physicochemical properties and electrocatalytic behaviors. Elemental analysis revealed that anchoring amino groups on biochar (SnO2/C-NHx) is advantageous for increasing amounts of amino groups attached, thereby enhancing biochar adsorption energy and subsequently increasing the loading of Sn. The CO2 adsorption curve indicated that amino groups anchored on biochar facilitate CO2 adsorption due to high specific surface area of biochar. X-ray photoelectron spectroscopy showed that amino groups anchored on SnO2 (NHx-SnO2/C) resulted in highly electron-rich centers on Sn, which promoted electron transfer between the catalyst and CO2. Electrochemical tests demonstrated the improved performance of amino-modified SnO2/biochar. SnO2/C-NHx exhibited enhanced Faraday efficiency, whereas NHx-SnO2/C showed higher current density. The disparity in electrochemical performance can be mainly attributed to the different selectivity towards rate-controlling steps of electron transfer and mass transfer induced by the various positions of amino groups anchoring.

DOI: 10.15376/biores.19.4.9049-9059

Keywords: Amino; Anchoring position; Tin oxide; Biochar; CO2 electroreduction

Contact information: a: Hefei Cement Research and Design Institute Corporation Ltd, Hefei, 230022, Anhui, China; b: Anhui Key Laboratory of Green and Low-carbon Technology in Cement Manufacturing, Hefei, 230022, Anhui, China; *Corresponding author: smx123@mail.ustc.edu.cn

 

GRAPHICAL ABSTRACT

INTRODUCTION

Compared to photo and thermal reduction methods, electrochemical CO2 reduction reaction (eCO2RR) is a more economical and environmentally friendly approach for CO2 conversion to organic chemicals (Han et al. 2019). Among the array of CO2 electroreduction products, formic acid (HCOOH) has garnered attention due to its versatility as a raw material for various chemicals and its role as a safe hydrogen storage carrier (Wang et al. 2022). Notably, HCOOH surpasses competitors such as carbon monoxide (CO) and methanol (CH3OH) in terms of electrochemical production cost (Jouny et al. 2018).

Numerous metallic elements, particularly their oxides, can serve as catalysts for CO2 electroreduction to HCOOH. SnO2 has garnered attention due to its affordability, non-toxicity, and high Faraday efficiency (FE) in HCOOH production (Kim et al. 2019), and it has been studied from various aspects such as particle size, morphology and content. Zhang et al. (2014) reported the electrocatalytic reduction of CO2 to HCOOH using carbon black loaded with small-sized nano SnO2 (3 to 6 nm). The FE was 90%, but the current density was only 5 mA/cm2. Liu et al. (2017) investigated the urchin-like nanostructured SnO2 and the FE of 62% was achieved. Yu et al. (2017) found that low SnO2 content led to insufficient catalytic sites, and high SnO2 content led to aggregation and blockage of catalytic sites, which are unfavorable for CO2 reduction. Nevertheless, studies have revealed that the FEHCOOH of SnO2 plateaus at 60 to 70% (Hu et al. 2018; Tay et al. 2021; Yuan et al. 2021; Yue et al. 2023). To enhance SnO2’s electrocatalytic efficiency, researchers have explored the integration of carbon materials and amino functional groups (Lu et al. 2017; Meng et al. 2019a; Cheng et al. 2021; Zhang et al. 2021). Biochar, a carbon material obtained through pyrolysis of waste biomass, could act as the carrier and boost conductivity and specific surface area of catalyst, thereby facilitating contact with CO2 and charge transfer (Fu et al. 2023; Hu et al. 2024). Tan et al. (2023) studied coconut husk-derived biochar for enhancing electrochemical conversion of CO2, and 140 mA/cm2 current density was achieved. Zhang et al. (2023) engineered the porosity of biochar to promote the electroreduction performance of single-atom Ni. Grafting amines onto catalysts is a strategy to further improve the electroreduction of catalysts because amino functional groups, serving as Lewis bases, can accelerate the adsorption and mass transfer of CO2 on SnO2 (Petrovic et al. 2021). Besides, the N atom in the amino group can also improve the chemical environment of SnO2 (Lin et al. 2022). Optimizing the composition and structure of tin oxide, biochar, and amino groups is crucial for enhancing SnO2’s electrocatalytic efficiency. Among these factors, the anchoring position of amino groups is a pressing concern. Various studies have shown that the role of surface modification functional groups varies with the anchoring position of the catalyst, especially for catalysts with complex structures (Sun et al. 2017; Meng et al. 2019b; Mulik et al. 2021). For instance, Dou et al. (2022) demonstrated that Co-N4 sites anchored on the edge of hierarchical porous structures exhibit higher selectivity in electroreduction of O2 compared to those anchored on the basal plane. Given that SnO2 with biochar forms a composite catalyst, both the biochar carrier and the SnO2 active sites can anchor amino groups. The anchoring points of amino groups inevitably exert a significant impact on the physicochemical properties of the overall catalyst and the characteristics of CO2 electroreduction. However, there is currently a lack of clear and systematic research on this aspect.

In this study, diethanolamine and rice husk serve as a representative amino functional group and a raw material for biochar, respectively. Biochar was prepared by pyrolysis and hydrothermal methods, and amino groups were anchored onto the biochar carrier and nano tin oxide active sites. The impact of various amino group anchoring points on the physicochemical properties, CO2 adsorption, and electroactivity of tin oxide was investigated.

EXPERIMENTAL

Materials

Diethanolamine (AR, DEA), dicyclohexylcarbodiimide (AR, DCC), dimethyl formamide (AR, DMF), ethanol (AR), tin (II) chloride (AR, SnCl2), potassium bicarbonate (AR, KHCO3), sulfuric acid (AR, 98 wt%, H2SO4), ammonium hydroxide solution (AR, NH4OH), and nitric acid (AR, 69.5 wt%, HNO3) were procured from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Rice husk was sourced from local farmland (Hefei, China). Carbon paper sheets were supplied by Toray (Japan). Nafion solution (5 wt%) and Nafion 117 membranes were obtained from DuPont (USA). All reagents and materials were used directly.

Catalyst Preparation

Biochar was produced by pyrolysis of rice husk at 550 °C in a nitrogen atmosphere and was labeled as C. Amination of biochar was conducted following procedures outlined in prior literature (Cheng et al. 2017). Initially, biochar underwent oxidation using a mixture of concentrated H2SO4 and HNO3. Subsequently, a specific amount of DEA was fully dispersed in DMF after ultrasonic treatment, followed by the addition of DCC and the oxidized biochar. The mixture was stirred for 24 h at 135 °C, and the resulting solution was filtered. The filtered mixtures were then washed with ethanol and dried overnight at 90 °C. The resulting product was designated as C-NHx.

A series of SnO2-supported biochar catalysts modified by amino groups were synthesized using a straightforward hydrothermal method (Jiang et al. 2005). Typically, 1 g of SnCl2 was dissolved in 80 mL mixture of ethanol and DIW, and then 200 mg of C was added to the solution. After stirring, NH4OH was added to adjust pH, and then the mixture was refluxed at 100 °C, followed by centrifugation, washing with ethanol, and drying. The resulting product was labeled as SnO2/C. Alternatively, by substituting C with C-NHx, the resulting product was designated as SnO2/C-NHx, indicating that the amino groups are anchored on the biochar carrier. To anchor amino groups onto nano SnO2, 0.5 g of DEA was added before the reflux process, and the obtained product was labeled as NHx-SnO2/C, indicating that the amino groups are located on the nano SnO2. The material obtained by impregnating SnO2/C into DEA-ethanol solution was named NHx-loaded SnO2/C.

Identification of Catalysts

Powder X-ray diffraction (XRD) patterns were obtained using a Smart Lab X-ray diffractometer (XRD, Japan) equipped with a Cu-Kα source. Fourier transform infrared spectroscopy (FTIR) analysis of the functional chemicals was performed using a ThermoFisher Nicolet IS20 instrument (USA) with KBr disks. X-ray photoelectron spectra (XPS) were collected using a Thermo ESCALAB 250 instrument (USA) with an Al K𝛼 X-ray source. CO2 and N2 adsorption-desorption isotherms were measured using a ChemStar TPx chemisorption analyzer (Quantachrome Instrument, USA). The content of Sn was analyzed using inductively coupled plasma mass spectrometry (ICP-MS, Thermo Fisher iCAP RQ, USA), while the nitrogen content was determined using an elemental analyzer (Vario EL cube, Germany). Transmission electron microscopy (TEM) on a JEM-2010 (HR) was operated at an accelerating voltage of 300 kV.

Electrochemical Experiments

Initially, 10 mg of catalyst was dispersed in 1 mL of ethanol containing 100 μL of NafionTM solution via ultrasonic dispersion. The resulting mixture was then sprayed onto carbon paper (1 cm2) repeatedly until the desired loading amount of 1 mg was achieved. The prepared working electrode and a standard electrode (Ag/AgCl) were placed in the cathodic chamber, with a counter electrode (Pt sheet) in the anodic chamber. The cathodic and anodic chambers were separated by a Nafion 117 membrane, and a 0.5 M KHCO3 solution was used as the electrolyte. High-purity CO2 was introduced for at least 30 min before performing linear sweep voltammetry (LSV) and chronoamperometry (i-t) studies. Liquid products were analyzed using ion chromatography (IC, Metrohm Eco, Switzerland), while gaseous products were analyzed using gas chromatography (GC, Thermo Trace 1310, USA) equipped with a flame ionization detector (FID) and a thermal conductivity detector (TCD). The FE for formic acid (FEHCOOH) was calculated by Eq. 1,

(1)

where n, F, and Q represent the amount of HCOOH per mole, Faraday’s constant (96485 C/mol), and charge (C), respectively.

The FE for CO and H2 were calculated using Eq. 2,

(2)

where v%, G, R, T, and I represent the concentration of CO or H2, CO2 flow rate (10 mL/min), gas constant (8.314 J/mol/K), temperature (298 K), and constant current (A), respectively.

RESULTS AND DISCUSSION

XRD spectra were utilized to analyze the crystal phase of each catalyst. As depicted in Fig. 1, all catalysts exhibited similar diffraction peaks at 2θ = 27° and 52°, corresponding to the (110) and (211) planes in the crystalline phase of SnO2, respectively. Additionally, weak diffraction peaks at 2θ = 43° were attributed to the characteristic peaks of carbon (Yua et al. 2017). This suggests that SnO2 is loaded onto biochar, and the different anchoring points of amino groups do not affect the crystal phase of SnO2 active sites.

Fig. 1. XRD patterns of SnO2/C, SnO2/C-NHx, and NHx-SnO2/C

The FTIR spectrum in Fig. 2 reveals the presence of functional groups in all catalysts. The characteristic peak observed in the range of 1500 to 1700 cm-1 was attributed to the bending vibration characteristic peaks of -NH, indicating the presence of amino functional groups anchored on all catalysts.

Fig. 2. FTIR spectra of SnO2/C-NHx, NHx-SnO2/C, and NHx-loaded SnO2/C

Table 1 summarizes the textural properties and elemental content of all catalysts. The Sn and N contents of SnO2/C-NHx were found to be 27.9% and 0.99%, respectively, which were higher than those of SnO2/C and NHx-SnO2/C. Based on the catalyst preparation process, it can be inferred that biochar with a high specific surface area (SBET) can anchor more amino functional groups compared to nano SnO2, and amino groups anchored on biochar also can enhance the adsorption energy of biochar, thereby increasing the loading amount of SnO2. The specific surface area (SBET) of SnO2/C was 63.1 m2/g, which was lower than 99.2 m2/g of NHx-SnO2/C and 67.5 m2/g of SnO2/C-NHx. This phenomenon indicated that amino groups are beneficial for the dispersion of tin oxide on biochar, as the Sn content of SnO2/C was significantly lower than NHx-SnO2/C and SnO2/C-NHx and theoretically SBET of SnO2/C should be higher. Simultaneously, it has been reported that amine substances can reduce the aggregation of metallic oxides during the formation process (Cheng et al. 2021), which also supported this phenomenon and inference. Furthermore, the order of decrease of SBET was NHx-SnO2/C < SnO2/C-NHx, meaning that amino groups anchored on SnO2 were more effective in delaying the aggregation of SnO2 compared to amino groups anchored on biochar.

Table 1. Textural Properties and Content of N and Sn for SnO2/C, SnO2/C-NHx, and NHx-SnO2/C

aDetermined by N2 adsorption-desorption isotherms at 77 K

The TEM images also clearly revealed the influence of different amino anchor sites on the size of tin oxide. As shown in Fig. 3, tin oxide aggregates to particles of about 20 nm, but the agglomeration phenomenon of NHx-SnO2/C was significantly lighter than that of SnO2/C-NHx. This result also supported the inference that amino groups anchored on SnO2 are more effective in delaying the aggregation.

Fig. 3. TEM images of SnO2/C-NHx (left) and NHx-SnO2/C (right)

All investigated catalysts were comprehensively analyzed for their compositions and chemical states using profiles obtained from XPS. As depicted in Fig. 4, compared to SnO2/C and SnO2/C-NHx, the O 1s peak of NHx-SnO2/C shifted to a higher binding energy, suggesting the presence of more oxygen deficit sites. Additionally, the N 1s and Sn 3d signals of NHx-SnO2/C shifted to lower binding energies, indicating a possible transfer of electrons from N to Sn. Consequently, the highly electron-rich centers of Sn species in NHx-SnO2/C facilitate CO2 activation. Moreover, the generated electrostatic field accelerates charge transfer (Lu et al. 2017). These findings suggest that the amino group anchored on nano SnO2 had a more pronounced effect on the chemical state of SnO2 compared to the amino group anchored on biochar.

Fig. 4. XPS spectra of Sn 3d (left), O 1s (middle) and N 1s (right) for SnO2/C, SnO2/C-NHx, and NHx-SnO2/C

The CO2 adsorption capacity of all catalysts was investigated using CO2 isothermal adsorption curves. As depicted in Fig. 5, the highest CO2 adsorption capacity was observed for SnO2/C-NHx at 5.7 cm3/g. In contrast, NHx-SnO2/C exhibited a CO2 adsorption capacity of 2.2 cm3/g, which was slightly higher than 1.9 cm3/g observed for SnO2/C. NHx-loaded SnO2/C also demonstrated a CO2 adsorption capacity of 4.5 cm3/g, despite having a similar N content to NHx-SnO2/C. Furthermore, the SBET of SnO2/C was lower than that of NHx-SnO2/C, theoretically leading to a higher CO2 adsorption capacity for NHx-SnO2/C. Therefore, it can be inferred that amino groups anchored on biochar were more conducive to CO2 adsorption than those anchored on SnO2. This may be attributed to the high SBET of biochar, facilitating the interaction of amino groups anchored on biochar with CO2.

Fig. 5. CO2 adsorption isotherms for SnO2/C, SnO2/C-NHx, NHx-loaded SnO2/C, and NHx-SnO2/C

A series of LSV and chronoamperometry (i-t) experiments were conducted to assess the electrochemical performances of all catalysts. As shown in Fig. 6(a), compared to the N2-saturated solution, all catalysts exhibited increased current levels in the CO2-saturated solution across the investigated potential range, indicating that eCO2RR was favored over the hydrogen evolution reaction (HER). Tafel slopes evaluated from the LSV curves of all catalysts in the CO2-saturated electrolyte are displayed in Fig. 6(c), representing the rate-controlling step of an initial single electron shift from the catalysts to CO2 generally (Jeong et al. 2021).

Fig. 6. (a) LSV curves of catalysts in N2– and CO2-saturated 0.5 M KHCO3 electrolytes at a scan rate of 50 mV/s, (b) i-t curves of catalysts at the potential -1.5 V vs. Ag/AgCl, (c) Tafel slopes curves of catalysts and (d) FE of HCOOH at the potential -1.5 V vs. Ag/AgCl after 1 h of eCO2RR

I-t tests of various catalysts at -1.5 V (vs. Ag/AgCl) for 1 h were performed to determine the FE for formic acid (FEHCOOH) and current density. The average current densities were approximately 22, 17, 15, and 12 mA/cm2 for NHx-SnO2/C, SnO2/C, SnO2/C-NHx, and NHx-loaded SnO2/C, respectively. It was evident from the comparison that NHx-SnO2/C could accelerate the electron transfer from Sn to CO2, which was also corresponding to XPS analysis. This indicates that NHx-SnO2/C exhibited superior electrochemical performance compared to other catalysts. However, the FEHCOOH presented different results. As shown in Fig. 6(d), SnO2/C-NHx exhibited an FEHCOOH of 73.3%, higher than 67.4%, 64.1%, and 59.1% for NHx-loaded SnO2/C, NHx-SnO2/C, and SnO2/C, respectively. These results indicate that the influence of amino anchoring position on current density and FE was significantly different. The amino group anchored on biochar was beneficial for improving FE, while the amino group anchored on SnO2 was favorable for increasing current density.

To further elucidate this difference, Table 2 summarized the concentration of HCOOH over all of the catalysts. Although the current density of SnO2/C-NHx was lower than SnO2/C, CO2 was still converted into more HCOOH by SnO2/C-NHx. Actually, even in the CO2-saturated solution, current density is still co-determined by proton and CO2 reduction (Jeong et al. 2021). Therefore, the decrease in current density of SnO2/C-NHx might have been due to the suppression of HER because of enhanced CO2 adsorption capacity of SnO2/C-NHx based on the CO2 adsorption isotherm. More electricity was used to convert CO2, and this resulted in the highest FEHCOOH for SnO2/C-NHx. Conversely, NHx-SnO2/C exhibited a higher concentration of HCOOH (236 ppm vs. 184 ppm) and a larger current density (22 mA/cm2 vs. 15 mA/cm2) due to improved electron transfer between CO2 and highly electron-rich centers of Sn species caused by anchoring amino groups on SnO2 based on XPS results. This indicated that NHx-SnO2/C was superior in conductivity and electrochemical rate as confirmed by LSV, i-t, and Tafel slope tests. However, a high-speed conversion rate meant that CO2 was consumed rapidly on the catalyst surface. Due to the poor CO2 adsorption capacity, NHx-SnO2/C could not fully contact with the CO2 source during electroreduction. This resulted in ineffective inhibition of HER, which was reflected in decreased FEHCOOH compared to SnO2/C-NHx. It is well known that the rate-determining steps of eCO2RR are divided into electrochemical processes and mass transfer processes. Thereby, the amino group anchored on biochar exerted on CO2 mass transfer process to achieve high FEHCOOH, but the amino group anchored on SnO2 acted on electrochemical processes for increasing current density and high HCOOH concentration.

Table 2. Concentration of HCOOH Catalyzed by All Catalysts

CONCLUSIONS

  1. Amino functional groups can enhance the activity of SnO2/biochar catalyst for the electroreduction of CO2. However, the influence of amino anchoring position on the reaction process is significantly different.
  2. The amino group anchored on biochar is advantageous for the adsorption of CO2 due to the high specific surface area of biochar, improving the FE of HCOOH through enhanced mass transfer processes. Conversely, the amino group anchored on SnO2 is favorable for the charge transfer between CO2 and highly electron-rich centers of Sn active site caused by transfer of electrons from N to Sn, increasing the current density of eCO2RR via boosting electron transfer processes.
  3. From the perspective of product concentration, anchoring amino groups on nano tin oxide (NHx-SnO2/C) is a more suitable modification strategy.

ACKNOWLEDGMENTS

The Chinese Ministry of Sciences and Technology of the PRC provided a grant for this research under its National Key Development and Research Program (No. 2020YFC1908704).

REFERENCES CITED

Cheng, Y. L., Xia, B. B., Fang, C. Q., and Yang, L. (2017). “Structure, mechanical and thermal properties of BMI/E-44/CNTs ternary composites via amination method,” J Mater. Sci. Technol. 33, 1187-1194. DOI: 10.1016/j.jmst.2017.04.013

Cheng, Y. Y., Hou, J., and Kang, P. (2021). “Integrated capture and electroreduction of flue gas CO2 to formate using amine functionalized SnOx nanoparticles,” ACS Energy Lett. 6, 3352-3358. DOI: 10.1021/acsenergylett.1c01553

Fu, S. L., Li, M., de Jong, W., and Kortlever, R. (2023). “Tuning the properties of N-doped biochar for selective CO2 electroreduction to CO,” ACS Catal. 13(15), 10309-10323. DOI: 10.1021/acscatal.3c01773

Han, N., Ding, P., He, L., Li, Y., and Li, Y. (2019). “Promises of main group metal–based nanostructured materials for electrochemical CO2 reduction to formate,” Adv. Energy Mater. 10(11), article 1902338. DOI: 10.1002/aenm.201902338

Hu, X., Yang, H., Guo, M., Gao, M., Zhang, E., Tian, H., Liang, Z., and Liu, X. (2018). “Synthesis and characterization of (Cu, S) Co-doped SnO2 for electrocatalytic reduction of CO2 to formate at low overpotential,” ChemElectroChem. 5(9), 1330-1335. DOI: 10.1002/celc.201800104

Hu, K., Jia, S. N., Shen, B. X., Wang, Z. Q., Dong, Z. J., and Lyu, H. H. (2024). “Nickel phthalocyanine anchored onto N-doped biochar for efficient electrocatalytic carbon dioxide reduction,” CHEM ENG J. 497, article 154686. DOI: 10.1016/j.cej.2024.154686

Jeong, S., Ohto, T., Nishiuchi, T., Nagata, K., Fujita, J. I., and Ito, Y. (2021). “Polyethylene glycol covered Sn catalysts accelerate the formation rate of formate by carbon dioxide reduction,” ACS Catal. 11, 9962-9969. DOI: 10.1021/acscatal.1c02646

Jiang, L. H., Sun, G. Q., Zhou, Z. H., Sun, S. G., Wang, Q., Yan, S. Y., Li, H. Q., Tian, J., Guo, J. S., Zhou, B., and Xin, Q. (2005). “Size-controllable synthesis of monodispersed SnO2 nanoparticles and application in electrocatalysts,” J. Phys. Chem. B. 109, 8774-8778. DOI: 10.1021/jp050334g

Jouny, M., Luc, W., and Jiao, F. (2018). “General techno-economic analysis of CO2 electrolysis systems,” Ind. Eng. Chem. Res. 57(6), 2165-2177. DOI: 10.1021/acs.iecr.0c01513

Kim, Y. E., Lee, W., Youn, M. H., Jeong, S. K., Kim, H. J., Park, J. C., and Park, K. T. (2019). “Leaching-resistant SnO2/γ-Al2O3 nano catalyst for stable electrochemical CO2 reduction into formate,” Ind Eng Chem Res. 78, 73-78. DOI: 10.1016/j.jiec.2019.05.042

Lin, C., Xu, Z. F., Kong, X. D., Zheng, H., Geng, Z. G., and Zeng, J. (2022). “Lysine functionalized SnO2 for efficient CO2 electroreduction into formate,” ChemNanomMat. 18, article e202200020. DOI: 10.1002/cnma.202200020

Liu, Y. Y., Fan, M. Y., Zhang, X., Zhang, Q., Guay, D., and Qiao, J. L. (2017). “Design and engineering of urchin-like nanostructured SnO2 catalysts via controlled facial hydrothermal synthesis for efficient electro-reduction of CO2,” Electrochim. Acta. 248, 123-132. DOI: 10.1016/j.electacta.2017.07.140

Lu, L., Sun, X., Ma, J., Zhu, Q., Wu, C., Yang, D., and Han, B. (2017). “Selective electroreduction of carbon dioxide to formic acid on electrodeposited SnO2@N-doped porous carbon catalysts,” Sci. China Chem. 61(2), 228-235. DOI: 10.1007/s11426-017-9118-6

Meng, N., Liu, C., Liu, Y., Yu, Y., and Zhang, B. (2019b). “Efficient electrosynthesis of syngas with tunable CO/H2 ratios over ZnxCd1‑xS-amine inorganic-organic hybrids,” Angew. Chem. Int. Edit. 58, 18908-18912. DOI: 10.1002/anie.201913003

Meng, Y., Jiang, J., Aihemaiti, A., Ju, T., Gao, Y., Liu, J., and Han, S. (2019a). “Feasibility of CO2 capture from O2-containing flue gas using a poly(ethylenimine)-functionalized sorbent: Oxidative stability in long-term operation,” ACS Appl. Mater. Interfaces 11, 33781-33791. DOI: 10.1021/acsami.9b08048

Mulik, B. B., Bankar, B. D., Munde, A. V., Chavan, P. P., Biradar, A. V., and Sathe, B. R. (2021). “Electrocatalytic and catalytic CO2 hydrogenation on ZnO/g-C3N4 hybrid nanoelectrodes,” Appl. Surf. Sci. 538, article 148120. DOI: 10.1016/j.apsusc.2020.148120

Petrovic, B., Gorbounov, M., and Soltani, S. M. (2021). “Influence of surface modification on selective CO2 adsorption: A technical review on mechanisms and methods,” Micropor. Mesopor. Mat. 312, article 110751. DOI: 10.1016/j.micromeso.2020.110751

Sun, Z., Fischer, J. M. T. A., Li, Q., Hu, J., Tang, Q., Wang, H., Wu, Z., Hankel, M., Searles, D. J., and Wang, L. (2017). “Enhanced CO2 photocatalytic reduction on alkali-decorated graphitic carbon nitride,” Appl. Catal. B-Environ. 216, 146-155. DOI: 10.1016/j.apcatb.2017.05.064

Tan, Y. C., Jia, S., Tan, J., Leow, Y., Zheng, R., Tan, X. Y., Dolmanan, S. B., Zhang, M., Yew, P. Y. M., and Ni, X. P. (2023). “Coconut husk-derived biochar for enhancing electrochemical conversion of CO2,” Mater. Today Chem. 30, article 101595. DOI: 10.1016/j.mtchem.2023.101595

Tay, Y. F., Tan, Z. H., and Lum, Y. W. (2021). “Engineering Sn-based catalytic materials for efficient electrochemical CO2 reduction to formate,” ChemNanoMat. 7, 380-391. DOI: 10.1002/cnma.202100031

Tian, Y. H., Li, M., Wu, Z. Z., Sun, Q., Yuan, D., Johannessen, B., Xu, L., Wang, Y., Dou, Y. H., Zhao, H. J., and Zhang, S. Q. (2022). “Edge-hosted atomic Co-N4 sites on hierarchical porous carbon for highly selective two-electron oxygen reduction reaction,” Angew. Chem. Int. Edit. 61. DOI: 10.1002/anie.202213296.

Wang, G., Chen, J., Li, K., Huang, J., Huang, Y., Liu, Y., Hu, X., Zhao, B., Yi, L., Jones, T. W., and Wen, Z. (2022). “Cost-effective and durable electrocatalysts for Co-electrolysis of CO2 conversion and glycerol upgrading,” Nano Energy. 92, 106751-106760. DOI: 10.1016/j.nanoen.2021.106751

Yua, J. L., Liu, H. Y., Song, S. Q., Wang, Y., and Tsiakaras, P. (2017). “Electrochemical reduction of carbon dioxide at nanostructured SnO2/carbon aerogels: The effect of tin oxide content on the catalytic activity and formate selectivity,” Appl Catal A-Gen. 545, 159-166. DOI: 10.1016/j.apcata.2017.07.043

Yuan, Y. S., Sheng, K., Zhuang, G. L., Li, Q. Y., Dou, C., Fang, Q. J., Zhan, W. W., Gao, H., Sun, D., and Han, X. G. (2021). “Collaboratively boosting charge transfer and CO2 chemisorption of SnO2 to selectively reduce CO2 to HCOOH,” ChemCommun. 57, 8636-8639. DOI: 10.1039/d1cc01818c

Yu, J. L., Liu, H. Y., Song, S. Q., Wang, Y., and Tsiakaras, P. (2017). “Electrochemical reduction of carbon dioxide at nanostructured SnO2/carbon aerogels: The effect of tin oxide content on the catalytic activity and formate selectivity,” Appl. Catal., A. 545, 159-166. DOI: 10.1016/j.apcata.2017.07.043

Yue, Y., Zou, X. H., Shi, Y. D., Cai, J. N., Xiang, Y. X., Li, Z. S., and Lin, S. (2023). “A low crystallinity CuO-SnO2/C catalyst for efficient electrocatalytic reduction of CO2,” J Electroanal Chem. 928, article 117089. DOI: 10.1016/j.jelechem.2022.117089

Zhang, S., Kang, P., and J. Meyer, T. (2014). “Nanostructured tin catalysts for selective electrochemical reduction of carbon dioxide to formate,” J. Am. Chem. Soc. 136, 1734-1737. DOI: 10.1021/ja4113885

Zhang, B., Chen, S., Wulan, B., and Zhang, J. (2021). “Surface modification of SnO2 nanosheets via ultrathin N-doped carbon layers for improving CO2 electrocatalytic reduction,” Chem Eng J. 421, article 130003. DOI: 10.1016/j.cej.2021.130003

Zhang, B., Song, D. Z., Liu, Y. Z., Deng, Q. X., Yuan, Y., Liu, Q. Q., and Huo, J. (2023). “Porosity engineering of biochar-based single-atom Ni catalysts to promote CO2 electroreduction over a broad potential window,” AICHE J. 69, 18175. DOI: 10.1002/aic.18175

Article submitted: August 6, 2024; Peer review completed: September 7, 2024; Revised version received: September 19, 2024; Accepted: September 22, 2024; Published: October 10, 2024.

DOI: 10.15376/biores.19.4.9049-9059