Woo Seok
Cheon
a,
Su Geun
Ji
ab,
Jaehyun
Kim
a,
Sungkyun
Choi
a,
Jin Wook
Yang
a,
Sang Eon
Jun
a,
Changyeon
Kim
a,
Jeewon
Bu
a,
Sohyeon
Park
a,
Tae Hyung
Lee
a,
Jinghan
Wang
a,
Jae Young
Kim
a,
Sol A
Lee
ac,
Jin Young
Kim
*a and
Ho Won
Jang
*ad
aDepartment of Materials Science and Engineering, Research Institute of Advanced Materials, Seoul National University, Seoul 08826, Republic of Korea. E-mail: hwjang@snu.ac.kr; jykim.mse@snu.ac.kr
bChemistry and Nanoscience Center, National Renewable Energy Laboratory, Golden, Colorado 80401, USA
cLiquid Sunlight Alliance (LiSA), Department of Applied Physics and Materials Science, California Institute of Technology, Pasadena, CA 91106, USA
dAdvanced Institute of Convergence Technology, Seoul National University, Suwon 16229, Republic of Korea
First published on 19th November 2024
The sustainable electrocatalytic reduction of carbon dioxide into solar fuels offers a potential pathway to mitigate the impact of greenhouse gas-induced climate change. Here, we successfully achieved a high solar-to-fuel (STF) efficiency of 11.5% by integrating a low-cost tandem solar cell with robust, high-performance, non-precious metal-based electrocatalysts. The bismuth-based cathode exhibited a high formic acid selectivity of 97.2% at a potential of −1.1 VRHE, along with an outstanding partial current density of 32.5 mA cm−2. Furthermore, upon undergoing more than 24 hours of electrolysis, we observed an enhancement in the catalytic activity. Through comprehensive analysis including in situ Raman spectroscopy and density functional theory (DFT) calculations, we elucidated that the in situ transformation of bismuth into bismuth subcarbonate (BOC) induces multiple effects: (i) the formation of grain boundaries between phases with distinct lattice parameters, (ii) electronic modulation due to defect formation, and (iii) changes in the binding modes of key reaction intermediates on active sites, resulting in the stabilization of *OCHO species. The cause of these phase transformations was attributed to the structural similarity between the cathode template and BOC. The sustainability of the STF efficiency sets a new benchmark for all cost-effective photovoltaic-coupled electrochemical systems.
Broader contextAs climate change continues to impact human survival negatively, interest in achieving net-zero emissions has increased. To reduce the atmospheric concentration of the predominant greenhouse gas, CO2, renewable energy sources must replace fossil fuels and drive electrochemical CO2 reduction reactions. In this study, we combine non-noble metal-based electrocatalysts with perovskite-Si tandem cells to investigate the feasibility of a sustainable electrolyzer. Additionally, we explore the in situ transformation of Bi-based electrocatalysts, which exhibit exceptional catalytic activity and selectivity. The partial transition of the cathode material during the reduction reaction induces modulation of the electronic structure and, thus, stabilization of key reaction intermediates. The insights gained from this work could significantly contribute to the design of catalytic materials and integrated systems for CO2 reduction. |
According to Kuhl's research, there are 16 products obtainable from the CO2RR.6 Among these, formic acid is one of the most valuable chemicals due to its high market price and effectiveness in hydrogen storage and transportation.7 Moreover, the reaction requires only two electrons per molecule, making it kinetically favorable and cost-effective. While several metals, including Sn, In, Cd, and Pb, demonstrate selectivity for formate production, Bi outperforms them in terms of faradaic efficiency (FE) and energy efficiency. Furthermore, the low toxicity of bismuth compared to other metals makes it environmentally friendly. Several studies were conducted to maximize the catalytic activity of Bi-based electrocatalysts. Fan et al. designed Bi nanotubes with a large current density and high selectivity in a wide potential window.8 In Li's work, Bi nanoribbons with modified edge sites enhanced the robustness of the reaction.9 While exploring synergistic effects with other metals, Zeng's group developed Bi@Sn with a core–shell structure that exhibited a high HCOOH production rate of HCOOH, attributed to the compressive strain in the Sn shell.10
However, Bi is susceptible to oxidation and restructuring, and comprehensive studies regarding catalyst transformation during electrolysis are still lacking. In the case of Cu-based cathodes, extensive research has been conducted on in situ phase or morphological transitions. For instance, the electrochemical fragmentation of Cu oxide enhances hydrocarbon production selectivity, whereas potential-driven clustering of Cu nanocubes leads to diminished CO2RR performance.11 Therefore, a deeper understanding of in situ transformation mechanisms and their impacts is essential to improve the long-term stability of catalysts.
In this study, we integrated cost-effective, high-efficiency perovskite-Si tandem cells with non-precious metal electrocatalysts. The unassisted PV-EC CO2RR system exhibited unmatched solar-to-fuel (STF) efficiency and demonstrated the capability of the solar cells and the electrocatalysts and the effectiveness of their integration. BiOI-derived cathodes, featuring abundant edge sites and a nanostructure favorable for mass transfer, exhibited a high HCOOH production rate of 606.3 μmol cm−2 h−1 and 97.2% selectivity. We also examined the in situ transformation of Bi during long-term electrolysis and its effect on catalytic activity. Utilizing first-principles calculations, we elucidate how the phase transformation of a Bi-based catalyst influences the binding mode of key reaction intermediates on the surface. We believe that these findings will significantly advance catalyst and solar-driven system designs for effective greenhouse gas reduction.
For comparison, Bi film with a thickness of 100 nm was deposited onto electropolished Cu foil by an e-beam evaporator at a deposition rate of 0.3 Å s−1. The base pressure of the vacuum chamber was 1.5 × 10−6 torr.
For the fabrication of the perovskite-Si tandem solar cells, a-Si:H layers were deposited on both sides of n-type Si wafers using plasma-enhanced chemical vapor deposition (PE-CVD). ITO/poly(triarylamine) (PTAA)/perovskite/C60/PEIE/ITO/Ag grid/MgF2 layers were subsequently deposited, with their thickness varying from 15 nm to 500 nm. For more details, please refer to Ji et al.'s work.12
FEgas (%) = Qgas/Qtotal × 100 (%) = (z × ngas × F)/Qtotal × 100 (%) |
= (z × P × V × Cgas × F)/R × T × Qtotal × 100 (%) |
The complete transformation from BiOI to Bi was verified through XRD patterns and Raman spectra (Fig. 1a and Fig. S6, ESI†). Comparison of the diffraction patterns before and after 5 minutes of electrochemical reduction reveals that the XRD peaks corresponding to BiOI (JCPDS No. 04-012-5693) completely disappear, and new peaks associated with rhombohedral Bi (JCPDS No. 00-005-0519) are newly formed. However, the nanosheet morphology of BiOI is overall maintained, and the surface becomes more porous, as evidenced by SEM images (Fig. 1b and c). An increase in the surface area leads to improved exposure of active sites and the mass transport of reactants. The thickness of a single vertical nanosheet was approximately 30 nm, and the height measured from the substrate was approximately 1 μm. Energy-dispersive X-ray spectroscopy (EDS) mapping on the bismuth vertical nanosheet (denoted as Bi-VNS) confirmed the uniform distribution of bismuth and oxygen, indicating the generation of native oxide on the Bi surface after exposure to the air (Fig. 1d).19 The lattice distances of 0.227 nm and 0.328 nm, corresponding to Bi (110) and (012) planes, were discerned from the high-resolution transmission electron microscopy (HRTEM) image and corresponding fast Fourier transform (FFT) pattern (Fig. 1e).
Nanostructures influence not only activity but also selectivity. In addition to the FIRC effect, the plastron effect may also occur due to the nanostructure. When contact angles were measured using a 10 μL water droplet, the angles were 80.4° for the Bi film and 135.0° for Bi-VNS (Fig. S9, ESI†). It is well known that greater hydrophobicity of the cathode leads to enhanced CO2RR selectivity. Gas chromatography (GC) and nuclear magnetic resonance (NMR) spectroscopy were used to quantitatively analyze the gas and liquid products formed during the reaction. Measurements of faradaic efficiency at various potentials revealed that CO2RR's selectivity was predominant over the hydrogen evolution reaction (HER) across a wide voltage window, recording a high formate/formic acid FE (FEHCOOH) of 97.2% at −1.1 VRHE (Fig. 2b). As formic acid (HCOOH) exists in the form of formate in the electrolyte with a neutral pH, these terms are used interchangeably in this paper. The FE value for CO remained almost constant at about 3%. The FE of H2 gas produced due to the competing HER gradually decreased as the voltage increased. While the control sample, Bi film, generally favored the CO2RR over the HER and exhibited an increase in FEHCOOH as the voltage rose, its liquid FE and partial current density were significantly lower than those of Bi-VNS (Fig. S10, ESI†). In the partial current density plot, the portions attributed to gaseous products were kept low, while the formate partial current density (jHCOOH) continued to increase with the applied potential, exceeding −32 mA cm−2 at −1.1 VRHE (Fig. 2c). This value is among the very superior performances measured in a gas-tight H-cell (Table S1, ESI†).
EIS analysis was conducted to evaluate the charge transfer capability at the interface. Upon analysis of the Nyquist plot, the charge transfer resistance (Rct) of Bi-VNS was significantly lower, thus facilitating the electron transfer to the adsorbed reactant (Fig. 2d).27,28 The simulated Rct values are listed in Table S2 (ESI†).
The commercialization of CO2 electrolyzers necessitates resolving the stability issue, a paramount concern.29–31 Many preceding studies on the CO2RR have observed a degradation in catalyst performance during measurements. To assess the stability of Bi-VNS, long-term electrolysis was conducted at −0.9 VRHE (Fig. 2e). Surprisingly, aside from an abrupt decrease in the initial three minutes, the current density continued to increase until saturation occurred around 24 h. The initial sharp decline in current density is observed in chronoamperometric electrolysis (CA) over a wide voltage range (Fig. S11, ESI†). The current density value after approximately 3 minutes followed the V–j relationship depicted in the LSV. As the CA at −0.9 VRHE proceeded, the current density increased while the FEHCOOH remained stable at 80%. Calculating the partial current density revealed that jHCOOH increased from −16.3 mA cm−2 to −24 mA cm−2. jCO was maintained below −1 mA cm−2 without significant changes, while jH2 increased from −1.8 mA cm−2 to −5.3 mA cm−2 (Fig. 2f). At first glance, the increased HER activity may seem to affect the selectivity of the CO2RR negatively. However, since the main product, formate, exists in liquid form and hydrogen is a gaseous product, their separation is facile. Moreover, the addition of the total quantity of profitable chemical substances contributes to the economic viability of the CO2 electrolyzer.14 The generation of CO, another gaseous product, remained suppressed, maintaining the purity of hydrogen gas. To the best of our knowledge, no reports have detailed the activation behaviour of Bi catalysts for both the CO2RR and HER. Additionally, such phenomena were not observed in the control sample of Bi film (Fig. S12, ESI†). We conducted ex situ characterization studies to gain a comprehensive understanding of this extraordinary phenomenon.
Bi3+ (in BiOI) + 3e− ![]() | (1) |
CO2 + H2O + 2e− → HCOO− + OH− | (2) |
2H2O + 4e− → H2 + 2OH− | (3) |
Bi + 3OH− ![]() | (4) |
Bi(OH)3 + CO2 ![]() | (5) |
Also, strain-induced lattice contraction appeared to have occurred due to the creation of a heterojunction between Bi and BOC, which have different lattice parameters. The co-existence of Bi and BOC results in a higher current density, an increased faradaic efficiency, and reduced overpotential compared to a single phase.33,34
The existence of a heterojunction between Bi and BOC in BOC-incorporated Bi-VNS (BOC@Bi-VNS) has been confirmed through HRTEM. The inset in Fig. 3b demonstrates that the lattice fringe distances of 0.372 and 0.296 nm correspond to the (011) and (013) interplanar spacing of BOC.23 The elemental mapping reveals a uniform distribution of Bi, O, and C across the nanosheet (Fig. 3c). This observation further confirms the formation of BOC. It is well known that a heterojunction between metal and metal oxide or other metal compounds can enhance the catalytic activity.35 As a consequence of junction formation, surface wettability may be enhanced,36 and charge transfer across the interface could become more efficient. Furthermore, it should be emphasized that lattice strain-induced distortions or built-in electric fields can significantly alter the electronic structure of surface atoms.37–40 As observed in the diffraction pattern in Fig. 3a, lattice contraction in BOC occurred due to the lattice mismatch as the CA duration extended. Cho et al. elucidated through DFT calculations that interfacial strain lowered the thermodynamic barrier of formate production.37 By disrupting the linear scaling relationship, lattice strain can potentially improve both the activity and selectivity of electrochemical processes.38 After Bi's in situ partial transformation into BOC, the charge transfer resistance decreased, as can be seen in the EIS analysis (Fig. S13, ESI†). The enhancement in electron transfer might have benefited from the newly formed BOC/Bi junction.
The possibility of electronic modulation of Bi species due to the BOC formation reaction near the surface was verified through XPS analysis (Fig. 3d). The high-resolution spectrum of Bi 4f exhibits two pairs of doublets, corresponding to Bi0 and Bi3+. In BOC@Bi-VNS, the ratios of the Bi3+ peaks (158.66, 163.96 eV) to the Bi0 peaks are larger than in Bi-VNS, a result of an increased quantity of oxidized Bi due to BOC formation after 24 h of electrolysis. Comparing this with Fig. 3a, where the Bi metal phase is dominant, it can be inferred that BOC is concentrated near the surface. Bi-VNS also exhibits Bi3+ peaks (158.95, 164.25 eV), which is attributed to the native oxide formation by exposure to air.41 Unlike the position of metallic Bi, which remains at 156.90 eV for Bi 4f7/2 and 162.22 eV for Bi 4f5/2, the position of the Bi3+ peak in BOC@Bi-VNS shifted by 0.29 eV to a lower binding energy. The reduction in binding energy could indicate the formation of oxygen vacancies in BOC@Bi-VNS, and this may have redistributed the negative charge to Bi atoms, benefitting the stabilization of *OCHO.20,42 The presence of CO32− within BOC was also confirmed in the O 1s and C 1s spectra (Fig. 3e and f).
To provide evidence for the proposed reaction pathway and identify the state of surface adsorbates during the reaction, in situ Raman spectroscopy analysis was conducted in the same electrolyte, 0.2 M KHCO3 solution. The changes were closely monitored by incrementally adjusting the reductive potential in 0.1 V steps, with spectra recorded at the open circuit potential (OCP) both before and after the reaction. At OCP, ions from the electrolyte near the electrode surface, CO32− and HCO3−, were observed at 1000 and 1064 cm−1, respectively.37,43–45 Additionally, adsorbed species such as *OCOOH and *O2CO were detected at 1347 and 1583 cm−1,46 respectively. Given that the focus was on the catalyst surface, the intensities of these surface adsorbates were pronounced. Furthermore, peaks corresponding to the stretching and bending modes of water molecules were located at 3223/3403 cm−1 and 1648 cm−1, respectively.45,47,48 Upon applying the reductive potential, the peaks for adsorbates like HCO3− and *OCOOH, located near or on the surface, disappeared, and the presence of the key reaction intermediate *OCHO was confirmed at 1360 cm−1.42,45,49,50 The disappearance of these adsorbate peaks indicates that, under the influence of the applied negative potential, the adsorbates were consumed in the reduction reaction. The *OCHO peak first appeared at an applied potential of −0.2 VRHE, gradually decreased in intensity as the potential was increased to −0.7 VRHE, and ultimately disappeared at −0.8 VRHE. This behavior can be attributed to the conversion of *OCHO to *OCHOH via proton-coupled charge transfer on the catalyst surface, a process that requires overcoming an activation barrier. As the applied voltage increases, a greater number of molecules can overcome this barrier.50,51 The complete disappearance of the *OCHO peak at −0.8 VRHE aligns well with the sharp increase in the faradaic efficiency for formic acid at this potential. Moreover, the peak intensity of water, the reactant molecule required for the CO2 reduction reaction, also varied with increasing voltage. Similar to *OCHO, the water peaks sharply decreased at −0.8 VRHE, which can be understood in the same context as the disappearance of the *OCHO peak. However, unlike *OCHO, the water peak partially recovered at OCP. This reappearance occurs because, unlike *OCHO, which forms only under a reductive potential, water is always present in the electrolyte, allowing its concentration near the surface to recover after the potential is removed.
DFT calculations were conducted to simulate CO2RR and HER processes on Bi and BOC surfaces. The (110) plane was chosen as the model surface for BOC, corroborated by XRD observations, while the (111) plane was selected for Bi. Despite the thermodynamically preferential growth of the (012) plane,52 the (111) surface was preferred for comparison due to its recognized activity based on first-principles calculations and widespread utilization.35,53 The free energy diagrams for the CO2RR were derived and depicted along the reaction pathway, comprising 2 sequential proton-coupled electron transfer (PCET) steps followed by desorption from the catalyst surface (Fig. 5a). The generation of *OCHO, a key reaction intermediate in the proposed pathway, was previously confirmed by in situ Raman (Fig. 4). While the reaction Gibbs free energy for *OCHO formation on the BOC (110) surface was only −0.07 eV, it was as high as 0.95 eV on the Bi (110) surface, indicating a significant difference in the reaction energetics. On Bi (111), the considerable energy barrier for forming the key intermediate species emphasizes that the initial step of CO2 conversion was the rate-determining step (RDS). Conversely, on the BOC (110) surface, *OCHO was greatly stabilized. As a result, the second PCET step, which requires an uphill energy of 0.54 eV, was identified as the rate-determining step (RDS). In Fig. 5b and c, charge density redistribution and Bader charge analysis were used to compare *OCHO stabilization on BOC and Bi surfaces. During the CO2-to-formic acid reaction pathway, the key reaction intermediate, *OCHO, adheres to the catalyst surface via oxygen atoms in a bidentate mode.54 On the BOC surface, the oxygen atoms attach to individual Bi atoms, intensifying the interaction and stabilizing *OCHO.34 The cyan-colored depletion regions in the charge density map clearly illustrate this point. Comparing the Bader charge changes of C, O, and Bi in BOC, bismuth is the only element whose charge changes significantly on the catalyst surface before and after the adsorption of *OCHO, indicating that Bi atoms serve as the active site. The Bader charge values of Bi atoms with *OCHO bound on BOC (110) are +1.95, 1.88 e (Fig. 5b). Since electron depletion is more pronounced than in Bi (111), the nucleophilic attack of oxygen atoms is stronger, allowing the key reaction intermediate to remain stably bound on the surface. The high HCOOH FE of Bi-based cathodes is linked to the adsorption mode of CO2 molecules on the metal surface. For CO-selective metals like Au and Ag, a nucleophilic attack occurs at the carbon atom, leading to the formation of the *COOH intermediate. In contrast, metals such as Bi and Sn undergo a nucleophilic attack at the oxygen atom, producing *OCHO. *COOH can either become *CO via PCET, which is released as CO gas if its binding energy to the surface is weak or transform into the C2+ product via C–C coupling if the binding energy is appropriate. In the case of bismuth, the energy required for *COOH generation is significantly higher than for *OCHO, resulting in a low CO faradaic efficiency at all potentials. The DFT results are consistent with previous studies suggesting that BOC is a stable and efficient phase for formate production under CO2RR conditions.27,32,35,55,56 Similar to the CO2RR, the BOC (110) surface exhibited key intermediate stabilization in the free energy diagram for the HER. The increased partial current densities for both hydrogen and formate observed in Bi-VNS's CA tests were attributed to the incorporation of the BOC phase. As proposed by Chen et al., the primary limitation for formate production is sluggish water dissociation; hence, strategies to stabilize *H and enhance CO2RR activity are effective.42 Wang et al.'s research further supports the result, as the authors demonstrated experimentally and computationally that the increasing amount of Bi–O bonds promotes the electrochemical reduction of CO2.57
To conclude, during the CO2RR, Bi-VNS underwent an in situ partial transformation, leading to the surface incorporation of BOC. The co-existence of BOC with Bi facilitated the stabilization of the key reaction intermediate, *OCHO, subsequently enhancing intrinsic activity and increasing the production rates of HCOOH on the cathode. The escalation in the catalytic performance of Bi-VNS also has roots in other factors, including increased roughness and active surface area, which led to improved active site exposure and mass transport (Fig. S14, ESI†). Additionally, ECSA analysis authenticated the growth in the specific surface area of BOC@Bi-VNS (Fig. S15, ESI†). However, this phenomenon was not observed in the control sample, Bi film. Fig. S12 (ESI†) shows the chronoamperometric result at 14 h. Other than small fluctuations, there was no increase in its catalytic activity, which is consistent with other literature in the field.8,19,23,35,43,58–61 XRD patterns indicate that the Bi film didn’t undergo phase transformation during the CO2 electrolysis (Fig. S16, ESI†). A plausible speculation related to this occurrence can be traced to the similarity found between layered structures of BiOX (X = Cl, Br, I) and BOC. BOC possesses a layered structure consisting of [Bi2O2]2+ and CO32− layers, a configuration strikingly analogous to that of BiOX, with the only difference being the type of interlayer anion.33 As the [Bi2O2]2+ slabs and iodide ion slabs in BiOI are stacked through weak van der Waals (vdW) force, the iodide ions are easily replaced with CO32− ions.62 Also, BiOX-derived Bi catalysts inherit the interlayer spacing of the precursor.63 Consequently, the easy formation of stable crystalline BOC through interlayer ion exchange in BiOI at the cathodic potential seems reasonable.34 The change of morphology and phase is well depicted in Fig. 6 and Fig. S14 (ESI†). While there have been studies starting from BOC to create phases that coexist with reduced Bi, there are only a few papers that have tried to create BOC incorporated Bi metal starting from BiOI and elucidate the conversion process and the mechanism of the high activity conversion process and the mechanism of high activity.
For unassisted CO2 reduction, low-cost, high-efficiency transition metal-based catalysts were employed at the cathode and anode. Synthesized through one-step electrodeposition, NiFe layered double hydroxide (NiFe-LDH) exhibited a porous structure suitable for robust oxygen evolution reaction (OER) catalytic activity (Fig. S17, ESI†). With the superior performance of an overpotential of 220 mV at a current density of 10 mA cm−2 and a very low Rct value, water splitting occurred vigorously.
The ratio between the solar cell's irradiated area and the catalyst area was determined based on LSV in a 2-electrode configuration employing Bi-VNS and NiFe-LDH. It was anticipated that a device operating point at VMP and IMP could be achieved when the catalyst's active area was set to 1.5 times the irradiated area. The faradaic efficiencies at various voltages were also measured in the 2-electrode configuration (Fig. 7c). FEHCOOH increases with higher voltage, reaching over 80% at 3 V. The sum of the faradaic efficiencies does not reach 100% at low voltages due to the low current values, resulting in inaccurate data when the product concentration was near or below the detecting limit.
After connecting all solar cells and electrodes in series, the CO2RR and water splitting occurred solely through electromotive force provided by the solar cell for an extended period of approximately 13 h (Fig. 7d). The experiment measured the longest stability in PV-EC CO2 electrolysis using low-cost PV cells, with the fuel production rates gradually increasing, as anticipated from Bi-VNS's activation. The sharp peaks and decreasing tails are attributed to bubble formation on the electrodes. The bubbles formed on the NiFe-LDH surface hindered reactant transport, and when the bubbles reached the threshold size, they detached from the surface. The current density was recovered upon each bubble's escape, showing an instant increase. This phenomenon was more prominent on the anode. Overall, STF was maintained above 10% and steadily increased. Unlike in Bi-VNS's 3-electrode measurement, a significant increase in current density was not observed, limited by the current that could flow through the connected PV module. The whole system's operation was confirmed through chopped-light measurement (Fig. 7e). As seen in Fig. 7f, the initial STF was only 10.8%, but it increased to 11.5% after 13 h as performance improvement occurred through Bi's in situ transformation. After 13 h of CO2 electrolysis measurement, the PV cell exhibited slight degradation. To our knowledge, maintaining and enhancing STF over such an extended period in a PV-EC system was unprecedented. This study achieved the highest STF efficiency in experiments using cost-effective PV cells (Fig. 7g and Table S3, ESI†).
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ey00209a |
This journal is © The Royal Society of Chemistry 2025 |