Designing atomic Ni/Cu pairs on a reactive BiOCl surface for efficient photo-chemical HCO3-to-CO conversion

Da Ke a, Bingjie Sun a, Yanjun Zhang b, Fan Tian b, Yu Chen c, Qingwen Meng a, Yixuan Zhang a, Zhangyi Hu a, Hongzhou Yang a, Chenyu Yang a, Xuyang Xiong *a and Tengfei Zhou *a
aInstitutes of Physical Science and Information Technology, Key Laboratory of Structure and Functional Regulation of Hybrid Material (Ministry of Education), Anhui University, Hefei 230601, China. E-mail: tengfeiz@ahu.edu.cn; xuyang@ahu.edu.cn
bSchool of Chemistry and Environmental Engineering, Wuhan Institute of Technology, Wuhan, 430205, China
cShanghai Synchrotron Radiation Facility, Shanghai Advanced Research Institute, Chinese Academy of Sciences, Shanghai, 201204, China

Received 2nd April 2024 , Accepted 3rd June 2024

First published on 3rd June 2024


Abstract

Solar-driven conversion of bicarbonate (HCO3) to carbonaceous fuels and/or chemicals provides an alternative route for the development of sustainable carbon economies. However, promoting the HCO3 reduction rate and tuning product selectivity remain significant challenges. This study reports the identification of isolated Ni/Cu atomic pairs dispersed on a BiOCl surface (Ni1/Cu1-BOC) as a promising candidate for efficient HCO3 reduction under UV-vis light irradiation. The optimized photocatalyst exhibits a high CO formation rate of 157.1 μmol g−1 h−1 with nearly 100% selectivity, even in the absence of added proton sources, sacrificial agents, or sensitizers. Experimental and theoretical investigations reveal that the atomically dispersed Ni/Cu pairs facilitate the protonation of HCO3 to CO2, which then undergoes a H+-assisted reduction pathway to produce CO, with *COOH as the intermediate. The synergistic effects of the Ni/Cu atomic pairs simultaneously promote the HCO3-to-CO2 conversion and the subsequent CO2-to-CO reduction, providing valuable insights for the development of efficient diatomic catalysts for photocatalytic HCO3 reduction reactions.


image file: d4ta02199a-p1.tif

Tengfei Zhou

Prof. Tengfei Zhou's research is focused on CO2 conversion, metal–CO2 batteries, low-temperature batteries and interfacial catalysis problems encountered in energy storage systems. Zhou started his professional research career as an Australian Research Council (ARC) DECRA Fellow in the Institute for Superconducting and Electronic Materials at the University of Wollongong from 2018 to 2021, under the supervision of Prof. Zaiping Guo. He is currently a professor at the Institutes of Physical Science and Information Technology at Anhui University.


Introduction

Direct photocatalytic reduction of bicarbonate (HCO3R) using naturally abundant carbonate/bicarbonate minerals as the feedstock, without additional CO2 supplementation in aqueous media, is a challenging yet fascinating approach for the development of economical carbon-resource conversion technologies.1 The dissolved HCO3 inevitably undergoes dissociation (HCO3 → CO2 (aq) + OH) and/or protonation (HCO3 + H+ → CO2 (aq) + H2O) processes, resulting in a near-zero concentration of CO2, which then needs to be further reduced to carbonaceous products.2,3 This CO2-mediated HCO3 conversion rate is kinetically sluggish due to the ultra-low CO2 concentration and the unfavourable alkaline pH that restricts CO2 formation.4 Moreover, the excessive presence of HCO3 anions on the CO2 reduction (CO2R) sites prevents the adsorption and activation of CO2 molecules, further lowering the overall activity.5,6 Therefore, designing catalysts to enhance local CO2(aq) concentration with specific HCO3/CO2 conversion sites is crucial for the high-performance HCO3R reaction.

Recent studies have shown that H+-driven HCO3 speciation through long-range proton transport (∼100 nm) can efficiently promote the local CO2R rate compared to HCO3 self-dissociation.7 The H+ generated from nearby redox sites can directly react with HCO3 and/or OH to generate in situ CO2(aq), which can then undergo H+-assisted CO2R to carbonaceous products, thereby facilitating the entire HCO3R process.8,9 However, the use of external proton sources, such as inorganic/organic acids (HCl, H2SO4, and H2BQ) or organic additives (EtOH and MeOH) and buffer solution (NaH2PO4/Na2HPO4), not only adds extra costs to the HCO3R systems but also introduces undesirable corrosion of the reactors.10–12 Additionally, the presence of H+ can competitively consume photo-induced electrons, further lowering the reaction activity and selectivity.13 Therefore, developing cost-effective proton sources (particularly H2O) that can provide H+ under neutral conditions, and novel catalysts that selectively reduce CO2, are crucial priorities for the practical development of efficient HCO3R systems.

Single atomic catalysts (SACs) with highly exposed reactive sites, unsaturated coordinate geometries, and unique chemical/physical properties have garnered extensive research interest in the fields of CO2 reduction, N2 fixation and H2O splitting.14–16 The isolated, dispersed metal sites in SACs may outperform metal clusters/alloys in selective chemical conversion to target products, as they are less susceptible to competing reactions due to their single-type atomic geometry, particularly in H+-assisted molecular catalysis that prevents undesirable H2 formation.17 For instance, manipulating Ni single atoms on carbon nitride/ZrO2 substrates to form Ni-based SACs has favoured electro/photo-chemical CO2 reduction to CO with >99% CO Faraday efficiency and/or ∼100% CO selectivity, even in aqueous media.18,19 However, while known SACs comprising noble metals (Ru, Ir and Re) and earth-abundant metals (Co, Ni and Fe) can reduce CO2 to CO, relatively few of them can realize HCO3R.20–26 The CO2-to-CO undergoes a 2e/2H+ transfer process, while HCO3-to-CO requires an additional H+ to pre-generate CO2(aq). Incorporating metals capable of accomplishing multi-e/H+ transfer/formation into SACs as catalytic pairs could theoretically drive the HCO3 reduction reaction.27 Previous work has demonstrated a self-adaptive dual-metal-site catalyst (DMSC) with flexible Ni–Cu pairs that can reduce CO2 to CH4,28,29 motivating the design of photocatalysts with Ni/Cu catalytic pairs for HCO3R attempts.

Herein, we present a PVP-mediated solvothermal method for fabricating a diatomic catalyst by immobilizing Ni and Cu single atoms on the photoreactive (001) surface of BiOCl (abbreviated as Ni1/Cu1-BOC). The single-site Ni and Cu atoms in their respective forms enabled photocatalytic CO2R and HCO3R under mild conditions without the assistance of sacrificial agents or photosensitizers. Ni1/Cu1-BOC exhibited an impressive HCO3-to-CO (in aqueous media) reduction rate of around 157.1 μmol g−1 h−1, with nearly 100% CO selectivity, outperforming the individual Ni1-BOC (115.2 μmol g−1 h−1), Cu1-BOC (89.1 μmol g−1 h−1) and pristine BOC (34.6 μmol g−1 h−1) catalysts. Detailed characterization revealed that the CO formation was closely correlated with the local CO2(aq) from the protonation of HCO3R, resulting in a CO2-intermediated HCO3R mechanism. In situ Fourier-transform infrared spectroscopy (in situ FTIR) and density functional theory (DFT) calculations confirmed that the Ni/Cu catalytic pairs controlled the reaction path to CO by significantly promoting *COOH as the intermediate and reducing the energy barriers for CO formation.

Results and discussion

The synthetic procedure for the Ni1/Cu1-BOC photocatalyst is described in Fig. 1a. NiCl2·6H2O, CuCl2·2H2O and Bi(NO3)3·5H2O respectively served as Ni2+, Cu2+ and Bi3+/Cl sources with ethylene glycol (EG) as the solvent. During the solvothermal stage, PVP (K30) was added as a surface-mediating agent to create more sites for the immobilization of singly dispersed Cu and Ni on the BiOCl substrate. PVP can also prevent the agglomeration of BiOCl nanoparticles, favoring the crystallization of BiOCl in well-defined morphologies with large specific surface areas.30 Other control photocatalysts, namely Ni1-BOC, Cu1-BOC and BOC, were prepared using a similar synthetic method. The Ni and Cu loadings in the different photocatalysts were quantified by inductively coupled plasma optical emission spectroscopy (ICP-OES) (Table S1).
image file: d4ta02199a-f1.tif
Fig. 1 (a) Scheme depicting the synthesis of Ni1/Cu1-BOC. (b and c) XRD patterns and Raman spectra of Ni1/Cu1-BOC, Ni1-BOC, Cu1-BOC and BOC. (d–f) TEM, HRTEM and EDS mapping images of Ni1/Cu1-BOC.

The XRD patterns of the Ni1/Cu1-BOC, Ni1-BOC and Cu1-BOC photocatalysts showed only the typical reflection peaks of tetragonal BiOCl, indicating that any Ni and Cu were highly dispersed on the support (Fig. 1b). Raman spectra of these photocatalysts also showed identical vibration peaks A1g2 internal Bi–Cl phonon mode (145 cm−1) and Eg internal Bi–Cl phonon mode (190 cm−1) to that of BiOCl,31 suggesting that the Ni and Cu were not incorporated by interlaminar Cl atoms (Fig. 1c). TEM images revealed that the Ni1/Cu1-BOC photocatalyst had a square-like morphology and was free of any irregular impurities or clusters (Fig. 1d). High-resolution TEM (HRTEM) images for the selected regions showed sets of interlaced lattice fringes, with interplanar spacings of around 0.27 nm which were well-crystallized along the [001] direction (Fig. 1e and S1). Energy dispersive spectroscopy (EDS) elemental mapping images for Bi, O, Cl, Ni and Cu revealed a uniform distribution of these elements in the entire Ni1/Cu1-BOC nanosheets (Fig. 1f).

To probe the coordination structures of Ni and Cu atoms in the Ni1/Cu1-BOC photocatalyst, Ni and Cu K-edge X-ray absorption near-edge structure (XANES), extended X-ray absorption fine structure (EXAFS) and wavelet transform (WT) data were collected for Ni1/Cu1-BOC and related reference materials. As shown in Fig. 2a, the Ni K-edge of Ni1/Cu1-BOC and Ni1-BOC exhibited higher pre-edge energies than that of Ni foil and NiO, indicating that the Ni center in Ni1/Cu1-BOC and Ni1-BOC had a higher valence state than NiO. The Ni K-edge EXAFS spectra of Ni1/Cu1-BOC and Ni1-BOC showed one intense peak at approximately 1.3 Å, closely related to the first shell of Ni–O coordination possibly due to the embedded Ni in the [Bi2–O2]2+ matrix (Fig. 2b). No Ni–Ni bond (2.17 Å for Ni foil and 2.34 Å for NiO) was observable in Ni1/Cu1-BOC and Ni1-BOC, indicating that Ni atoms were singly dispersed. The atomic dispersion of Ni species in Ni1/Cu1-BOC and Ni1-BOC was further confirmed by WT analysis of Ni K-edge EXAFS oscillations. The WT contour plots of Ni1/Cu1-BOC and Ni1-BOC showed maximum intensities attributable to Ni–O bonding (at 4.9 and 4.8 Å−1, respectively), whereas the Ni–Ni bonding was absent (Fig. 2c).32Fig. 2d shows the Cu K-edge XANES spectra of Ni1/Cu1-BOC, Cu1-BOC, Cu foil and CuO. The Cu K-edge of Ni1/Cu1-BOC and Cu1-BOC was close to that of CuO, suggesting that Cu atoms in Ni1/Cu1-BOC and Cu1-BOC were positively charged with valence states near +2. The Cu K-edge EXAFS spectra of Ni1/Cu1-BOC and Cu1-BOC showed peaks only assigned to the Cu–O bond (1.5 Å) while Cu–Cu interactions (2.24 Å for Cu foil and 2.46 Å for CuO) were undetectable (Fig. 2e). This indicated that Cu atoms in Ni1/Cu1-BOC and Cu1-BOC were also isolated and dispersed, for which the WT contour plots of Ni1/Cu1-BOC demonstrated different Cu features than that of Cu foil and CuO (Fig. 2f). To better understand the electronic and spatial structures of Ni and Cu atoms in the Ni1/Cu1-BOC photocatalyst, Ar ion sputtering and high-resolution X-ray photoelectron spectroscopy (XPS) were performed over the Ni 2p, Cu 2p, Bi 4f, O 1s, and Cl 2p regions of Ni1/Cu1-BOC along with selected reference samples including Ni1-BOC, Cu1-BOC and BOC. As shown in Fig. 2g, the Ni 2p spectra of Ni1/Cu1-BOC was deconvoluted into five peaks including Ni2+ (2p3/2 855.4 eV and 2p1/2 873 eV), Niδ+,δ>2 (2p3/2 857.3 eV) and Ni2+ shake up satellites (880.2 eV and 861.4 eV).33 This evidenced the presence of over-oxidized Ni, consistent with the afore-analyzed Ni K-edge XANES results. The Ni 2p signals almost vanished after Ar ion sputtering, implying that Ni single atoms were mainly dispersed on the outermost surfaces of Ni1/Cu1-BOC. Fig. 2h shows the Cu 2p spectra of Ni1/Cu1-BOC. The intense peak centered at 940.5 eV was ascribed to Bi 4s orbitals while peaks at 932.1 eV and 951.0 eV were assigned to Cu2+ cations.34 The Cu atoms were also localized on the outermost layer of Ni1/Cu1-BOC, and its oxidative states were rarely influenced by Ni loading (Fig. S2). The Bi 4f spectra of Ni1/Cu1-BOC are depicted in Fig. 2i. Two peaks at 158.4 eV and 164.4 eV were clearly resolved, related to partially reduced Bi3−x states. The surface Bi atoms were relatively negatively charged in contrast to the inner form, indicating that Bi atoms were unconventionally coordinated. This might be attributed to the additives Ni and Cu in the BiOCl substrate, which significantly modified the delocalized electrons near Bi, O and Cl atoms, thus altering the redox capacity (Fig. S3).


image file: d4ta02199a-f2.tif
Fig. 2 (a) Ni K-edge XANES spectra, (b) k3-weighted Fourier transforms of the Ni K-edge EXAFS spectra and (c) relative WT plots of Ni foil, Ni1/Cu1-BOC, Ni1-BOC and a NiO standard. (d) Cu K-edge XANES spectra, (e) k3-weighted Fourier transforms of the Cu K-edge EXAFS spectra and (f) relative WT plots of Cu foil, Ni1/Cu1-BOC, Cu1-BOC and a CuO standard. (g–i) Ar ion sputtering X-ray photoelectron spectra of Ni 2p, Cu 2p and Bi 4f orbitals in Ni1/Cu1-BOC.

The photocatalytic HCO3R experiments were performed under UV-vis light irradiation without any assistance of photosensitizers and sacrificial agents. Prior to each test, the custom-built photoreactor was purged using high-purity Ar to eliminate gaseous impurities. Fig. 3a shows that Ni1-BOC delivered a HCO3-to-CO production rate of 115.2 μmol g−1 h−1, much higher than that of BOC (34.6 μmol g−1 h−1). After the introduction of Cu single atoms, the photocatalytic activity was further improved. Ni1/Cu1-BOC offered the highest CO formation rate of 157.1 μmol g−1 h−1, which was 1.4, 1.8 and 4.5 times higher than the activities of Ni1-BOC, Cu1-BOC and the BOC reference photocatalyst, respectively. The H2 formation was negligible during CO production, probably due to the alkaline HCO3R conditions that prevent H+-coupling. This superb HCO3R rate demonstrated by Ni1/Cu1-BOC could also be extended to the CO2R reaction, where humid CO2 gas replaced aqueous HCO3 solution as the catalytic feed. As shown in Fig. 3b, Ni1/Cu1-BOC afforded the optimal CO2R performance amongst Ni1-BOC, Cu1-BOC and BOC catalysts. The CO produced from CO2 was less than that from HCO3, which might be because of the lower solubility of CO2 in aqueous media. Fig. 3c shows the HCO3R results over Ni1/Cu1-BOC at different pH values. The CO formation rate was dramatically increased by H+ addition, and suppressed under excessively alkaline conditions. Based on the Bjerrum plot, the fraction of water-soluble CO2 was nearly 0 at pH > 8.5.7 This could be utilized to explain the almost 0 μmol g−1 h−1 CO formation rate at pH ≈ 9, because HCO3 might not evolve into CO2 thus suspending the overall HCO3-to-CO conversion. Lowering the pH value to a neutral value and/or increasing the local CO2 contents could promote CO production, further evidencing that CO2 was the key intermediate during the HCO3R reaction. Cycling experiments of Ni1/Cu1-BOC showed no decrease in the CO formation after 5 runs, indicating that the photocatalyst possessed good photostability (Fig. S4). XRD patterns and TEM images of Ni1/Cu1-BOC before and after the reaction were almost identical, suggesting that Ni/Cu catalytic pairs were stable against aggregating during the photocatalytic tests (Fig. S5). To gain deeper insights into the mechanism underpinning the fast HCO3R conversion on Ni1/Cu1-BOC as compared with BOC, in situ FTIR spectra were deliberately employed to monitor the reaction details. As shown in Fig. 3d and e, IR spectra of Ni1/Cu1-BOC and BOC showed peak positions ranging from 1200 to 1700 cm−1, respectively assigned to chemisorption CO2 (˙CO2, 1248 cm−1, 1700 cm−1), bicarbonates (HCO3, 1400–1435 cm−1), monodentate carbonates (m-CO32−, 1381 cm−1 and 1451 cm−1) and bidentate carbonates (b-CO32−, 1357 cm−1 and 1556 cm−1).35–37 The intensity of these peaks intensified on prolonging the irradiation time, indicating that carbonaceous species accumulated on the catalyst surfaces. The obvious differences between Ni1/Cu1-BOC and BOC were the symmetric IR peaks centered at 1650 and 1540 cm−1, which can be readily indexed as the *OH (–OH in H2O) and *COOH groups.38,39 The *OH signal gradually increased on BOC and almost vanished on Ni1/Cu1-BOC, implying that Ni1/Cu1-BOC was unfavourable for *OH or H2O adsorption in contrast to BOC. The *COOH peak of Ni1/Cu1-BOC was strengthened compared to that of BOC, evidencing that Ni1/Cu1-BOC facilitated protonation during the HCO3R process with *COOH as the intermediate. Fig. 3f directly provides IR peak intensities of *OH and *CO collected over Ni1/Cu1-BOC and BOC. The *CO formation slope showed negative correlation with the surface *OH uptake, clearly emphasizing that excess *OH impeded HCO3-to-CO conversion. This phenomenon can be explained as *OH consumed additional H+ involved in HCO3-to-CO2 and CO2-to-CO paths, thus weakening CO2/*COOH evolution as well as lowering the CO production. The afore-mentioned analysis depicted a typical HCO3R route through HCO3-CO2-*COOH–CO, during which Ni1/Cu1-BOC outperformed BOC in CO formation due to its efficiency in H+ utilization (Fig. 3g).


image file: d4ta02199a-f3.tif
Fig. 3 (a) CO production rates from HCO3R collected over Ni1/Cu1-BOC, Ni1-BOC, Cu1-BOC and BOC. Reaction conditions: 1 mol per L HCO3 aqueous solution, full spectrum irradiation and 10 mg of the catalyst. (b) CO production rates from CO2R collected over Ni1/Cu1-BOC, Ni1-BOC, Cu1-BOC and BOC. (c) CO formation rates of Ni1/Cu1-BOC under different pH conditions. (d and e) In situ FTIR spectra of BOC and Ni1/Cu1-BOC during photocatalytic HCO3R reactions. (f) *OH and *CO signals of BOC and Ni1/Cu1-BOC collected from the in situ FTIR spectra. (g) Schematic diagrams showing the HCO3R mechanism over Ni1/Cu1-BOC.

Note that the multi-protons/electron transfer process was involved in HCO3-to-CO conversion, and the photoelectric properties and electron-migration capacities of Ni1/Cu1-BOC and BOC were then respectively investigated. Fig. 4a displays the light absorption characteristics of Ni1/Cu1-BOC and BOC. The BOC sample had an adsorption edge near 380 nm, arising from the O 2p to Bi 6p transitions. Ni1/Cu1-BOC showed additional peaks ranging from 400–700 nm, involving O 2p to Ni/Cu 3d transitions and d–d transitions in Ni/Cu 3d orbitals, respectively. The band gap values of Ni1/Cu1-BOC and BOC were calculated to be 3.31 eV and 3.46 eV based on the Kubelka–Munk function, clearly emphasizing that Ni/Cu incorporation changed the intrinsic band alignment of BiOCl (Fig. 4b). The conduction band (CB) levels of Ni1/Cu1-BOC and BOC were determined to be −0.85 V and −0.7 V (vs. NHE) from the Mott–Schottky plots, all satisfying the thermodynamic requirements for CO production from CO2 (Fig. S6).40 Additionally, the valence band (VB) positions of Ni1/Cu1-BOC and BOC were estimated to be about 2.46 eV and 2.76 eV, all above the redox potential for H2O oxidation to O2 (Fig. 4c).41 This observation suggested that CO and O2 possibly served as the products from reduction/oxidation-half reactions, with pre-analyzed surface *OH probably being the oxygen evolution intermediate.42 The dynamics of photo-induced charge carriers were then employed by photoluminescence (PL) and electrochemical impedance spectroscopy (EIS). As shown in Fig. 4d and e, Ni1/Cu1-BOC displayed weakened PL peaks with the smallest EIS semicircles as compared with BOC. Note that the quenched PL signals reflected the suppressed electron–hole recombination and the EIS radius empirically correlated with the inner-resistance; the afore-mentioned results thus collectively indicated that Ni1/Cu1-BOC possessed the best charge separation efficiency due to its high conductivity.43–48Fig. 4f shows the photocurrent signals collected over Ni1/Cu1-BOC and BOC during the same time interval. The photovoltaic peaks demonstrated by Ni1/Cu1-BOC surpassed that of BOC, further evidencing that Ni1/Cu1-BOC facilitated photoelectric conversion and charge carrier migration.49–51


image file: d4ta02199a-f4.tif
Fig. 4 (a and b) Diffuse reflectance spectra and Tauc plots of Ni1/Cu1-BOC and BOC. (c) Band alignments of Ni1/Cu1-BOC and BOC. (d–f) Steady-state PL spectra, EIS and photocurrent patterns of Ni1/Cu1-BOC and BOC.

To further elucidate the photocatalytic HCO3R dynamics at the atomic level, DFT calculations of elementary HCO3R steps over Ni1/Cu1-BOC and BOC were separately simulated. Based on the XAFS studies, we constructed surface models as BOC (001) with/without Ni/Cu catalytic pairs (Ni/Cu-BOC and BOC) to represent the likely coordination geometries of Ni1/Cu1-BOC and BOC samples (Fig. 5a). Fig. 5b shows Gibbs energy barriers involved in HCO3-to-CO2 conversion. The solvated HCO3 anions were first chemically adsorbed on the catalytic surfaces, followed by C–OH cleavage (by dehydration) to form *CO2. Ni/Cu-BOC offered a lower energy barrier for *HCO3 protonation/dehydration to *CO2, evidencing that the Ni1/Cu1-BOC catalyst favoured H+ utilization. The subsequent transformation of CO2 into CO was predicted to occur endothermically over the Ni/Cu-BOC model, and the formation of adsorbed *COOH was the rate limiting step (Fig. 5c). The summed-up energy barriers required for CO production within the Ni/Cu-BOC model (ΔG = 0.57 eV) also lower than that of the BOC model (ΔG = 0.63 eV), indicating the advantages of Ni/Cu catalytic pairs in HCO3R/CO2R reactions. The dissociative H+ tended to be captured by BOC as adsorbed *H, while Ni/Cu-BOC unfavoured *H formation thus preferred H+-driven HCO3 speciation/H+-assisted CO2 reduction (Fig. 5d).


image file: d4ta02199a-f5.tif
Fig. 5 (a) DFT models representing BOC and Ni1/Cu1-BOC samples. (b–d) Energy profiles for the elementary steps of HCO3-to-CO2, CO2-to-CO and H+-to-H2 pathways over BOC and the Ni/Cu-BOC model.

Conclusions

In conclusion, a Ni1/Cu1-BOC photocatalyst comprising singly dispersed Ni/Cu catalytic pairs over a BiOCl substrate afforded excellent performance for the photocatalytic reduction of bicarbonate (HCO3) to carbon monoxide (CO) under UV-vis irradiation. Benefiting from the surface-immobilized Ni/Cu catalytic pairs, a remarkable CO formation rate of 157.1 μmol g−1 h−1 was realized in the HCO3 aqueous solution, surpassing the activities of the individual Ni1-BOC, Cu1-BOC and BOC catalysts. In situ FTIR results and photoelectric tests showed that the protons and electrons required for the HCO3-to-CO2 conversion were dramatically increased and promoted on the Ni/Cu-involved catalyst, facilitating the subsequent CO2 reduction to CO with *COOH as the reactive intermediate. DFT calculations also demonstrated that the Ni/Cu-BOC model lowers the energy barriers involved in the HCO3 reduction process by significantly promoting the utilization of H+ through a series of protonation and dehydration pathways leading to CO production. These findings shed light on the rational design of diatomic photocatalysts for sustainable HCO3/CO2 reduction reactions. The Ni1/Cu1-BOC photocatalyst developed in this study delivers an unusually high CO formation rate during the photocatalytic HCO3 reduction tests, showcasing its potential for practical applications in the development of economical carbon-resource conversion technologies.

Data availability

All the data supporting this article have been included in the main text and the ESI.

Author contributions

D. Ke, B. Sun and Y. Zhang: investigation, visualization and writing – original draft. F. Tian and H. Yang: supervision and conceptualization. Y. Chen: resources and software. Q. Meng, Y. Zhang, Z. Hu and C. Yang: data curation, formal analysis and visualization. X. Xiong and T. Zhou: supervision, writing – review & editing, resources and funding acquisition.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The authors are grateful for the financial support of the Anhui Provincial Natural Science Foundation for Outstanding Young Scholar (2208085Y05), Anhui Provincial Scientific Reuter Foundation for Returned Scholars (2022LCX030), and the National Natural Science Foundation of China (52202198 and 22272207). The authors also thank Dr Yu Chen from the Shanghai Synchrotron Radiation Facility for his help with the XAS data analysis.

Notes and references

  1. C. Hepburn, E. Adlen, J. Beddington, E. Carter, S. Fuss, N. Mac Dowell, J. Minx, P. Smith and C. Williams, Nature, 2019, 575, 87–97 CrossRef CAS PubMed.
  2. S. Zhang, C. Chen, K. Li, H. Yu and F. Li, J. Mater. Chem. A, 2021, 9, 18785–18792 RSC.
  3. D. Pimlott, A. Jewlal, Y. Kim and C. Berlinguette, J. Am. Chem. Soc., 2023, 145, 25933–25937 CrossRef CAS.
  4. M. Dunwell, Q. Lu, J. Heyes, J. Rosen, J. Chen, Y. Yan, F. Jiao and B. Xu, J. Am. Chem. Soc., 2017, 139, 3774–3783 CrossRef CAS PubMed.
  5. C. Stalder, S. Chao and M. Wrighton, J. Am. Chem. Soc., 1984, 106, 3673–3675 CrossRef CAS.
  6. N. Queyriaux, ACS Catal., 2021, 11, 4024–4035 CrossRef CAS.
  7. R. Yanagi, T. Zhao, M. Cheng, B. Liu, H. Su, C. He, J. Heinlein, S. Mukhopadhyay, H. Tan and D. Solanki, J. Am. Chem. Soc., 2023, 145, 15381–15392 CrossRef CAS PubMed.
  8. J. Schneider, H. Jia, J. Muckerman and E. Fujita, Chem. Soc. Rev., 2012, 41, 2036–2051 RSC.
  9. Q. Dong, X. Zhang, D. He, C. Lang and D. Wang, ACS Cent. Sci., 2019, 5, 1461–1467 CrossRef CAS PubMed.
  10. M. Nielsen, E. Alberico, W. Baumann, H. Drexler, H. Junge, S. Gladiali and M. Beller, Nature, 2013, 495, 85–89 CrossRef CAS PubMed.
  11. M. Symes and L. Cronin, Nat. Chem., 2013, 5, 403–409 CrossRef CAS PubMed.
  12. K. Joya, Y. Joya and H. De Groot, Adv. Energy Mater., 2014, 4, 1301929 CrossRef.
  13. N. Elgrishi, M. Chambers and M. Fontecave, Chem. Sci., 2015, 6, 2522–2531 RSC.
  14. A. Wang, J. Li and T. Zhang, Nat. Rev. Chem, 2018, 2, 65–81 CrossRef CAS.
  15. S. Ding, M. Hülsey, J. Pérez-Ramírez and N. Yan, Joule, 2019, 3, 2897–2929 CrossRef CAS.
  16. Q. Wu and C. Wu, J. Mater. Chem. A, 2023, 11, 4876–4906 RSC.
  17. X. Yang, A. Wang, B. Qiao, J. Li, J. Liu and T. Zhang, Acc. Chem. Res., 2013, 46, 1740–1748 CrossRef CAS PubMed.
  18. H. Yang, S. Hung, S. Liu, K. Yuan, S. Miao, L. Zhang, X. Huang, H. Wang, W. Cai and R. Chen, Nat. Energy, 2018, 3, 140–147 CrossRef CAS.
  19. X. Xiong, C. Mao, Z. Yang, Q. Zhang, G. Waterhouse, L. Gu and T. Zhang, Adv. Energy Mater., 2020, 10, 2002928 CrossRef CAS.
  20. S. Takizawa, T. Okuyama, S. Yamazaki, K. Sato, H. Masai, T. Iwai, S. Murata and J. Terao, J. Am. Chem. Soc., 2023, 145, 15049–15053 CrossRef CAS PubMed.
  21. S. Sato, T. Morikawa, T. Kajino and O. Ishitani, Angew. Chem., Int. Ed., 2013, 52, 988–992 CrossRef CAS PubMed.
  22. B. Su, Y. Kong, S. Wang, S. Zuo, W. Lin, Y. Fang, Y. Hou, G. Zhang, H. Zhang and X. Wang, J. Am. Chem. Soc., 2023, 145, 27415–27423 CrossRef CAS PubMed.
  23. L. Chen, Z. Guo, X. Wei, C. Gallenkamp, J. Bonin, E. Anxolabéhère-Mallart, K. Lau, T. Lau and M. Robert, J. Am. Chem. Soc., 2015, 137, 10918–10921 CrossRef CAS PubMed.
  24. H. Rao, L. Schmidt, J. Bonin and M. Robert, Nature, 2017, 548, 74–77 CrossRef CAS PubMed.
  25. K. Sun, Y. Huang, Q. Wang, W. Zhao, X. Zheng, J. Jiang and H.-L. Jiang, J. Am. Chem. Soc., 2024, 146, 3241–3249 CrossRef CAS PubMed.
  26. X. Yu, M. Sun, T. Yan, L. Jia, M. Chu, L. Zhang, W. Huang, B. Huang and Y. Li, Energy Environ. Sci., 2024, 17, 2260–2268 RSC.
  27. B. Mondal, A. Rana, P. Sen and A. Dey, J. Am. Chem. Soc., 2015, 137, 11214–11217 CrossRef CAS PubMed.
  28. J. Li, H. Huang, W. Xue, K. Sun, X. Song, C. Wu, L. Nie, Y. Li, C. Liu and Y. Pan, Nat. Catal., 2021, 4, 719–729 CrossRef CAS.
  29. Y. Cai, J. Fu, Y. Zhou, Y. Chang, Q. Min, J. Zhu, Y. Lin and W. Zhu, Nat. Commun., 2021, 12, 586 CrossRef CAS.
  30. Y. Lei, G. Wang, S. Song, W. Fan and H. Zhang, CrystEngComm, 2009, 11, 1857–1862 RSC.
  31. A. Phuruangrat, S. Thongtem and T. Thongtem, Appl. Phys. A, 2020, 126, 245 CrossRef CAS.
  32. T. Zhang, X. Han, H. Yang, A. Han, E. Hu, Y. Li, X. Yang, L. Wang, J. Liu and B. Liu, Angew. Chem., Int. Ed., 2020, 59, 12055–12061 CrossRef CAS PubMed.
  33. S. Uhlenbrock, C. Scharfschwerdt, M. Neumann, G. Illing and H. Freund, J. Phys.: Condens. Matter, 1992, 4, 7973–7978 CrossRef CAS.
  34. S. Poulston, P. Parlett, P. Stone and M. Bowker, Surf. Interface Anal., 1996, 24, 811–820 CrossRef CAS.
  35. Y. Zhang, Z. Xu, Q. Wang, W. Hao, X. Zhai, X. Fei, X. Huang and Y. Bi, Appl. Catal., B, 2021, 299, 120679 CrossRef CAS.
  36. E. Köck, M. Kogler, T. Bielz, B. Klötzer and S. Penner, J. Phys. Chem. C, 2013, 117, 17666–17673 CrossRef PubMed.
  37. S. Collins, M. Baltanás and A. Bonivardi, J. Phys. Chem. B, 2006, 110, 5498–5507 CrossRef CAS PubMed.
  38. J. Wu, X. Li, W. Shi, P. Ling, Y. Sun, X. Jiao, S. Gao, L. Liang, J. Xu, W. Yan, C. Wang and Y. Xie, Angew. Chem., Int. Ed., 2018, 57, 8719–8723 CrossRef CAS PubMed.
  39. X. Li, Y. Sun, J. Xu, Y. Shao, J. Wu, X. Xu, Y. Pan, H. Ju, J. Zhu and Y. Xie, Nat. Energy, 2019, 4, 690–699 CrossRef CAS.
  40. S. Nitopi, E. Bertheussen, S. Scott, X. Liu, A. Engstfeld, S. Horch, B. Seger, I. Stephens, K. Chan and C. Hahn, Chem. Rev., 2019, 119, 7610–7672 CrossRef CAS PubMed.
  41. M. Yagi and M. Kaneko, Chem. Rev., 2001, 101, 21–36 CrossRef CAS PubMed.
  42. X. Shi, S. Back, T. Gill, S. Siahrostami and X. Zheng, Chem, 2021, 7, 38–63 CAS.
  43. H. Zhang, C. Wang, H. Luo, J. Chen, M. Kuang and J. Yang, Angew. Chem., Int. Ed., 2023, 62, e202217071 CrossRef CAS PubMed.
  44. C. Ban, Y. Wang, Y. Feng, Z. Zhu, Y. Duan, J. Ma, X. Zhang, X. Liu, K. Zhou and H. Zou, Energy Environ. Sci., 2024, 17, 518–530 RSC.
  45. J. Ma, D. Wu, Y. Feng, C. Ban, L. Xia, L. Ruan, J. Guan, Y. Wang, J. Meng, J. Dai, L. Gan and X. Zhou, Nano Energy, 2023, 115, 108719 CrossRef CAS.
  46. J. Chen, F. Zhang, M. Kuang, L. Wang, H. Wang, W. Li and J. Yang, Proc. Natl. Acad. Sci. U. S. A., 2024, 121, e2318853121 CrossRef CAS PubMed.
  47. Y. Shi, G. Zhan, H. Li, X. Wang, X. Liu, L. Shi, K. Wei, C. Ling, Z. Li and H. Wang, Adv. Mater., 2021, 33, 2100143 CrossRef CAS PubMed.
  48. Y. Shi, J. Li, C. Mao, S. Liu, X. Wang, X. Liu, S. Zhao, X. Liu, Y. Huang and L. Zhang, Nat. Commun., 2021, 12, 5923 CrossRef CAS PubMed.
  49. J. Zhao, Y. Mu, L. Wu, Z. Luo, L. Velasco, M. Sauvan, D. Moonshiram, J. Wang, M. Zhang and T. Lu, Angew. Chem., Int. Ed., 2024, 136, e202401344 CrossRef.
  50. R. Fang, Z. Yang, J. Sun, C. Zhu, Y. Chen, Z. Wang and C. Xue, J. Mater. Chem. A, 2024, 12, 3398–3410 RSC.
  51. C. Ban, Y. Wang, J. Ma, Y. Feng, X. Wang, S. Qin, S. Jing, Y. Duan, M. Zhang, X. Tao, L. Gan and X. Zhou, Chem. Eng. J., 2024, 488, 150845 CrossRef CAS.

Footnotes

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ta02199a
These authors contributed equally.

This journal is © The Royal Society of Chemistry 2024
Click here to see how this site uses Cookies. View our privacy policy here.