Multi-interfacial nanosheet-intercalated structure with abundant oxygen vacancies promotes electrocatalytic oxygen evolution

Di Wanga, Zhe Suna, Wenguang Cuib, Chaozhen Hec and Zhongkui Zhao*a
aState Key Laboratory of Fine Chemicals, Department of Catalysis Chemistry and Engineering, School of Chemical Engineering, Dalian University of Technology, 2 Linggong Road, Dalian 116024, P. R. China. E-mail: zkzhao@dlut.edu.cn
bCollege of Chemical Engineering Shijiazhuang University, 288 Zhufeng Street, Shijiazhuang 050035, Hebei, China
cInstitute of Environmental and Energy Catalysis, School of Materials Science and Chemical Engineering, Xi'an Technological University, Xi'an, 710021, China

Received 22nd May 2024 , Accepted 17th July 2024

First published on 18th July 2024


Abstract

The electron coupling effect at the interface and the introduction of oxygen vacancies (Ov) play critical roles in the electrocatalytic activity. The key to lowering the energy barrier of the oxygen evolution reaction (OER) is to build the interface properly and increase oxygen vacancies. In this work, a nickel phosphide on nickel foam-derived catalyst with rich Ov and a multi-interfacial nanosheet intercalated structure, labeled as (Fe,La)Ni2P-r, was created on nickel foam by using a straightforward two-step technique, namely hydrothermal and electrochemical oxidation. The addition of Fe–La creates a heterogeneous interface on the catalyst surface, causes electron transfer and redistribution, and lowers the binding energy of intermediates. At the same time, connected with DFT, it was discovered that the addition of Fe–Ov–La significantly lowered the Gibbs free energy of the reaction process, enhanced the intermediate species adsorption, and hastened the oxygen release. Only 197 mV was required to obtain a current density of 10 mA cm−2 with a Faraday efficiency of around 100%, and the required voltage is 390 mV at a current density of 800 mA cm−2. This study not only presents an excellent Ov-enriched multi-interface OER electrocatalyst, but also paves a path for the development of cost-effective noble metal and polymetallic catalysts.


1. Introduction

As a new type of renewable energy, hydrogen energy is widely regarded as the key to future energy transformation because of its high calorific value, low energy consumption and excellent combustion behavior.1–3 Hydrogen can be produced by electrolysis of water, natural gas transformation, biomass conversion and other ways, among which hydrogen production by electrolysis of water is pollution-free, and is one of the most ideal solutions to replace traditional fossil energy. However, precious metal-based materials are currently used as industrial catalysts for the oxygen evolution reaction (OER) to overcome the high overpotential in the water-splitting process, but their high cost limits their large-scale commercial applications.4–6 Therefore, the development of non-precious metal water-splitting electrocatalysts with low price, excellent performance and low energy consumption has become the top priority of current research.

In recent years, transition metal phosphide has been frequently used in the OER due to its unique hydrogenase-like catalytic mechanism.7,8 With a growing understanding of the OER mechanism, studies have shown that spontaneously reconstructed metal (oxy)hydroxides after the OER are the true active species,9,10 but the optimal conditions for the reconstruction process cannot be ensured. Therefore, controlled electrochemical reconstruction is an effective and feasible method to enhance OER activity, as demonstrated in previous work by our group.11 Furthermore, interface engineering is considered an efficient strategy for improving the intrinsic activity of OER catalysts. Studies have reported that constructing crystalline–amorphous interfaces can more effectively promote OER activity, with the crystalline phase accelerating the electron transfer, and the irregular amorphous phase resulting in a random distribution of internal atoms, generating numerous unsaturated coordination atoms and providing abundant active adsorption sites.12–14 Generally, low-cost transition metals, especially Fe, Mn, Co, and other 3d transition metals, are considered ideal additives for regulating interface construction.15–17 It is now believed that oxygen vacancies (Ov), as the active sites in the OER, can regulate surface-oxygen interactions, thus accelerating the reaction kinetics and enhancing the catalytic activity.18,19 In recent years, rare earth elements have also attracted attention due to their unique 4f sub electronic layer structure. Some studies have confirmed that introducing rare earth elements into transition metal catalysts can create corresponding oxygen vacancies. By utilizing the synergistic effect of the multi-interface catalyst and oxygen vacancies, the electronic structure can be optimized, the kinetic energy barrier can be lowered, and the performance of catalysts can be significantly improved.20,21

Taking inspiration from the above, we envision a strategy based on electrochemical oxidation to obtain OER catalysts with higher activity by constructing interfaces and introducing oxygen vacancies. Therefore, we used the urea hydrothermal method to grow Ni2P nanosheets on nickel foam (NF), and then introduced Fe and La in the process of Ni2P reconstruction to obtain a multi-interfacial catalyst with a unique nanosheet-intercalated structure and rich oxygen vacancies, named (Fe,La)Ni2P-r. Interestingly, in a 1.0 M KOH solution, the catalyst exhibited excellent OER catalytic activity, comparable to most reported catalysts. The unique structure exposes more active sites in the catalyst. Combined with DFT results, it was found that the strong electronic coupling at the interface and the synergistic effect of oxygen vacancies promote the breaking of the HO–O bond, leading to an increase in the content of intermediate species in *OOH and accelerating the release of oxygen. In addition, this study also provides a promising approach for designing multi-interfacial catalysts rich in oxygen vacancies.

2. Results and discussion

Based on the previous work of this group, the optimal conditions were first obtained by screening different metals and optimizing metal loading amounts and various parameters. Then, a unique nanosheet stacking structural (Fe,La)Ni2P-r catalyst supported on a Ni foam (NF) substrate was prepared using a simple two-step method (Scheme 1). The atomic ratios were determined by ICP-AES and XPS tests, and the differences in the data were observed due to variation in the characterization methods (Table S1). From the scanning electron microscopy images (Fig. 1A and S1), it can be seen that the (Fe)Ni2P-r and (Fe,Mn)Ni2P-r catalysts exhibit a stacked nanosheet-intercalated morphology. However, after the introduction of Fe and La, a novel sheet stack structure with local bulges was unexpectedly discovered, as shown in Fig. 1B–D. The nanosheet stack structure of the catalyst is assembled with nanosheet units in a spiral layer-by-layer stacking manner, forming a unique structure of sheet assembly, which subsequently influences the electrochemically active surface area of the catalyst (Fig. S6 and Table S5). The TEM image of the (Fe,La)Ni2P-r catalyst (Fig. 1E) provides strong evidence that the catalyst is assembled from nanosheet units, consistent with the SEM results. The unique structure facilitates mass transfer in the oxygen evolution reaction.22 Further, the HAADF-STEM image (Fig. 1F) and EDX analysis (Fig. 1G) demonstrate the uniform distribution of Fe, La, Ni, and O elements on the (Fe,La)Ni2P-r catalyst.
image file: d4cy00653d-s1.tif
Scheme 1 Schematic illustration of the synthesis process of the (Fe,La)Ni2P-r catalyst.

image file: d4cy00653d-f1.tif
Fig. 1 A) SEM image of the Ni2P-r catalyst. B–D) SEM images of the (Fe,La)Ni2P-r catalyst with different magnifications. E) TEM image of the (Fe,La)Ni2P-r catalyst. F) HAADF-STEM image of the (Fe,La)Ni2P-r catalyst. G) The corresponding EDX elemental mappings of Fe, La, Ni and O for (Fe,La)Ni2P-r. H) HRTEM image of (Fe,La)Ni2P-r. I) XRD patterns of different catalyst samples. Pink is (Fe,La)Ni2P-r, light blue is (Fe,Mn)Ni2P-r and orange is (Fe)Ni2P-r.

The XRD patterns in Fig. 1I reveal that the (Fe,La)Ni2P-r catalyst differs from the reference samples by the appearance of new diffraction peaks at 31.59°, 40.93°, 42.99°, 44.30°, 44.65°, 47.78°, 49.15°, 49.76°, 50.28°, and 51.59°, matching perfectly with the (−110), (−201), (−202), (200), (003), (−113),(−212), (−203), (−210), and (120) crystal planes of LaOOH (PDF#19-0656), indicating the formation of the new species LaOOH. Clear lattice fringes of Ni2P and LaOOH can be observed in the HRTEM image in Fig. 1H, confirming their high crystallinity, consistent with the XRD results. Meanwhile, the (Fe,La)Ni2P-r catalyst exhibits distinct interfaces, with lattice spaces of 2.460 Å and 2.700 Å corresponding to the (201) plane of Ni2P and the (−212) plane of LaOOH (Fig. S9B). In addition, some regions have been observed without an obvious lattice fringe, confirming the formation of surface heterojunction interfaces.23 Notably, the introduction of Fe and La induces a strong interface effect, effectively promoting the oxidation of Ni2P, as evidenced by the disappearance of the diffraction peaks near 33°.22,23 Overall, the formation of interfaces facilitates the generation of more active *OOH species, accelerating oxygen release. Additionally, the reconstruction layer thickness of 5.4 nm, significantly thicker than those of the (Fe)Ni2P-r and (Fe,Mn)Ni2P-r catalysts (as shown in Fig. S2), further confirms this point.

X-ray photoelectron spectroscopy (XPS) was employed to analyse the surface chemical composition and electronic structure of the samples. Fig. S3A shows the presence of Fe and La, confirming the successful introduction of Fe and La elements, consistent with the EDX results. All high-resolution spectra were calibrated based on the C 1s peak at 284.6 eV. As depicted in Fig. 2A, the Ni 2p spectrum of the (Fe,La)Ni2P-r sample exhibits peaks at 853.8 eV attributed to metallic nickel and at 855.8 eV and 857.0 eV attributed to Ni2+ and Ni3+, indicating the coexistence of Ni2+ and Ni3+ in the (Fe,La)Ni2P-r sample.24–27 Comparison of the binding energies of Ni2+ in the different catalysts reveals a positive shift relative to the (Fe)Ni2P-r sample for the (Fe,Mn)Ni2P-r sample and the (Fe,La)Ni2P-r sample by 0.1 eV and 0.3 eV, suggesting a stronger Fe–La interaction than Fe–Mn, facilitating more Ni oxidation to higher oxidation states, potentially beneficial for *OOH generation.24,28 Furthermore, the ratio of Ni3+/Ni2+ in the (Fe,La)Ni2P-r, (Fe,Mn)Ni2P-r, and (Fe)Ni2P-r samples confirms this, with values of 0.83, 0.75, and 0.72, respectively. In the Fe 2p XPS spectra (Fig. 2B), the (Fe,Mn)Ni2P-r catalyst exhibits a slight negative shift (0.3 eV) compared to the (Fe)Ni2P-r catalyst. Upon the transition of the metal from Mn to La, the (Fe,La)Ni2P-r catalyst experiences an additional negative shift (0.2 eV) relative to the (Fe,Mn)Ni2P-r catalyst (855.6 eV), attributed to the low electronegativity of metal La, facilitating the electron density transfer from La to Fe, resulting in more unsaturated Fe states, thereby altering the adsorption energy on the active sites and affecting its inherent OER activity.29–31 The O 1s spectrum is deconvoluted into four characteristic peaks at 530.0 eV, 530.8 eV, 531.6 eV, and 532.8 eV (Fig. 2C), attributed to M–O (metal–oxygen in lattice oxygen), O–H (hydroxyl bond), Ov, and H2O (adsorbed water molecules), respectively. Integration calculations of O under different states reveal that the Ov content of the (Fe,La)Ni2P-r catalyst (23.1%) is higher than that of the (Fe,Mn)Ni2P-r catalyst (17.6%). Electron paramagnetic resonance spectra further confirm the introduction of oxygen vacancies (Ov) at g = 2.000 for the (Fe,La)Ni2P-r and (Fe,Mn)Ni2P-r catalysts (Fig. S4), with the signal intensity of the (Fe,La)Ni2P-r catalyst being greater than that of the (Fe,Mn)Ni2P-r catalyst, indicating that La introduction promotes the generation of oxygen vacancies (Ov), allowing *OOH adsorption and acquiring more active species, as confirmed.32,33


image file: d4cy00653d-f2.tif
Fig. 2 High-resolution XPS spectra of Ni 2p (A), Fe 2p (B), O 1s (C) and La 3d (D) for different catalysts.

More interestingly, under the influence of La3+, the adsorption energy of water on the surface of (Fe,La)Ni2P-r sample is lower than that of (Fe,Mn)Ni2P-r, demonstrating the excellent surface water adsorption ability of (Fe,La)Ni2P-r. Relevant reports have shown that electron induced charge transfer can promote the adsorption ability of water, supporting the generation of M–OOH.34,35 Similarly, in the XPS spectrum of La 3d, two peaks are observed corresponding to the satellite peaks (sat) of La 2d5/2 and La3+ (Fig. 2D). For the (Fe)Ni2P-r catalyst, the introduction of La is further verified, in accordance with the findings of EDX and ICP. For the P 2p spectrum (Fig. S3B), an enhancement of the P–M peak is observed in the (Fe,La)Ni2P-r catalyst, attributed to the stronger P–La bonding energy caused by La's lower electronegativity (1.1) compared to Fe (1.8) and Mn (1.5).29 Therefore, the XPS results not only confirm the successful preparation of the (Fe,La)Ni2P-r sheet–sphere heterostructure but also demonstrate that the introduction of Fe and La leads to a strong interface effect and enhancement of oxygen vacancies. This enhancement of oxygen vacancies and the shift in element binding energies may be attributed to the surface electron transfer induced by the interface charge redistribution among LaOOH, MOOH (M = Fe, Ni), and Ni2P to maintain system charge neutrality.36–40

The corresponding electrochemical tests on the prepared catalyst samples were performed in a 1.0 M KOH solution. The (Fe)Ni2P-r catalyst and (Fe,Mn)Ni2P-r catalyst are used as control samples. Fig. 3A shows the LSV curves of the different catalysts, and compared with the control catalyst samples and the pristine Ni2P,11 the (Fe,La)Ni2P-r heterostructure exhibits a significantly higher current density at the same potential, demonstrating the strongest OER activity (Table S2). The oxidation of metallic Ni before water oxidation causes the bulge in the curve, a phenomenon already documented.41–43 To obtain the heterogeneous catalyst with the highest performance, the OER activity of different metals (Ce, Mn, In, Bi, Co, Zn, Cu, and Ca) is screened by an orthogonal test (Fig. S5A, Tables S7 and S8), revealing that the (Fe,La)Ni2P-r catalyst exhibits the highest current density at a constant potential. When the parameters of the (Fe,La)Ni2P-r heterogeneous structure are modified, it is discovered that after 40 minutes of reconstruction at 1.60 V (vs. RHE), the best performance is obtained with a molar ratio of 10[thin space (1/6-em)]:[thin space (1/6-em)]1 (Fig. S5B–D).


image file: d4cy00653d-f3.tif
Fig. 3 A) LSV curves of RuO2, (Fe)Ni2P-r, (Fe,Mn)Ni2P-r and (Fe,La)Ni2P-r catalysts in 1.0 M KOH. B) Time–potential plots of the samples at different current densities. C) Tafel plots and D) Nyquist plots (inset: the equivalent circuit diagram) of the catalysts.

More importantly, the performance of (Fe,La)Ni2P-r surpasses that of the commercial catalyst RuO2 (overpotential of 246 mV at 10 mA cm−2 current density), with (Fe,La)Ni2P-r exhibiting an overpotential of 197 mV at 10 mA cm−2 current density (Fig. 3B), comparable to most other reported catalysts (Table S9). The Tafel slope reflects the intrinsic kinetics of the reaction. In Fig. 3C, the Tafel slope of the (Fe,La)Ni2P-r catalyst introduced with La is 30 dec−1, superior to those of the control samples (Table S3). The electrochemical impedance spectroscopy results of the catalysts are displayed in Fig. 3D, where Rct stands for the charge transfer resistance and Rs stands for the solution interface resistance. The Rct of the (Fe,La)Ni2P-r catalyst is the smallest (Table S4), indicating that the introduction of Fe and La reduces the energy barrier encountered by the charge transfer at the electrode/electrolyte interface. Hence, the enhancement of the OER activity in the (Fe,La)Ni2P-r catalyst may be attributed to the synergistic promotion of charge transfer by oxygen vacancies and interfaces, accelerating their intrinsic kinetics.34 Exploring the specific activity and number of active sites of the catalysts are crucial for evaluating the OER activity. The CV curves of all catalysts at scanning rates of 20 mV s−1 to 160 mV s−1 were collected (Fig. 4A, S6C and D), obtaining proportional curves of the scan rate and current density (Fig. 4B), and then calculating the ECSA of different catalysts based on the double-layer capacitance Cdl. The electrochemically active area of the (Fe,La)Ni2P-r catalyst is 78.1 cm−2, which is less than those of other catalysts, as indicated in Fig. S6A and Table S5. Normalizing the ECSA to LSV polarization curves (Fig. 4C), it is evident that the specific activity of (Fe,La)Ni2P-r is the highest, indicating the optimal intrinsic activity of the (Fe,La)Ni2P-r sample. Fig. S6B visually reflects the number of active sites of the catalysts. At 1.55 V (vs. RHE) potential, the TOF values of the (Fe,La)Ni2P-r, (Fe,Mn)Ni2P-r, and (Fe)Ni2P-r catalysts are 0.52 s−1, 0.34 s−1, and 0.29 s−1, respectively. Therefore, the higher OER activity of the (Fe,La)Ni2P-r catalyst is partially attributed to the increase in the number of active sites of the catalyst. In addition, the catalyst has a Faraday efficiency of about 100% (Fig. S11 and Table S6).


image file: d4cy00653d-f4.tif
Fig. 4 A) CV curves at different scanning rates for (Fe,La)Ni2P-r. B) Graphs of the proportional relationship between the sweep rate and current density. C) Normalized LSV curves based on ECSA. D) The chronoamperometry measurement of the catalyst up to 100 h at 20 mA cm−2 current density.

One of the most crucial criteria for evaluating the practical performance of a catalyst in OER applications is its long-term durability under strong alkaline operating conditions. The microsphere heterostructure material (Fe,La)Ni2P-r maintains its activity for 100 hours at 20 mA cm−2 without significant degradation, as evidenced by the nearly overlapping polarization curves before and after 100 hours, indicating the excellent stability of the catalyst (Fig. 4D and S7). Characterization of the morphology and surface electronic structure of the catalyst after electrochemical testing and after 100 hours reveals, in Fig. S8, that the catalyst largely retains its original microsphere structure, which is conducive to the exposure of active sites and facilitates the activation of O2 due to its layered sheet–ball structure. Transmission electron microscopy images (Fig. S9) further confirm that the morphology of the catalyst remains largely unchanged after 100 hours, with clear lattice stripes of LaOOH and Ni2P and the presence of heterointerfaces. In the Fe 2p and Ni 2p spectra before and after the reaction (Fig. S10A and B), a significant positive shift can be observed, attributed to the conversion of Fe–P and La–P bonds to Fe–O and La–O bonds, indicating surface restructuring of the catalyst during the OER.28 The disappearance of the P–M bond in Fig. S10C provides further evidence of surface restructuring to form FeOOH and LaOOH.44 In Fig. S10D, the presence of La3+ ions is still clearly observable after the reaction, confirming the precise introduction of La. Overall, all results demonstrate that the introduction of La generates a unique nanosheet stack structure with local bulges. The presence of interfaces between LaOOH, M–OOH (M = Fe and Ni), and Ni2P, along with the enhancement of oxygen vacancies (Ov), induces charge redistribution, resulting in changes in the electronic structure, altering the adsorption energies of H and OH on the active sites, and promoting an increase in the *OOH content.

To investigate the influence of oxygen vacancies and heterointerfaces on the intrinsic mechanism of the catalyst reaction, density functional theory (DFT) calculations were employed. Firstly, models for all catalysts were established (Fig. 5A and S12). In addition, a catalyst model without Ov (denoted as (Fe,La)Ni2P-r(s)) was established on the surface of the (Fe,La)Ni2P-r catalyst to visualize the effect of Ov (Fig. 5A). In order to better understand the influence of oxygen vacancies, the formation energies of oxygen vacancies in the different catalysts were calculated, and it was found that the energy barrier required for (Fe,La)Ni2P-r was the lowest (Fig. 5B). At the same time, in order to make the calculation more accurate, as shown in Fig. S14, we calculated the oxygen vacancies between the three different atoms of Fe–Ov–La, Fe–Ov–Ni, and La–Ov–Ni for (Fe,La)Ni2P-r at different positions, and found that the formation energy between Fe–Ov2–La was the lowest (Fig. S14). Fig. 5C demonstrates that the presence of oxygen vacancies reduces the Gibbs free energy from 1.88 eV to 1.49 eV, indicating that the oxygen vacancies between Fe and La significantly decrease the Gibbs free energy of the reaction process, facilitating the adsorption of the intermediate species and accelerating the oxygen release.45


image file: d4cy00653d-f5.tif
Fig. 5 A) Structural models of (left) (Fe,La)Ni2P-r (s) and (right) (Fe,La)Ni2P-r (the orange area is the oxygen vacancy area). Color codes: green (Fe), light yellow (La), blue (Ni), pink (P), red (O) and white (H). B) Oxygen vacancy formation energy of different catalysts at different positions. C) Gibbs free energies and overpotentials of all catalysts. D) Differential charge diagrams of different catalysts (the top left is the (Fe,La)Ni2P-r catalyst, the top right is the (Fe,Mn)Ni2P-r catalyst, the lower left is the (Fe)Ni2P-r catalyst, and the lower right is the (Fe,La)Ni2P-r(s) catalyst).

Furthermore, from Fig. 5C, it can be observed that the *OH to *O step in the (Fe,La)Ni2P-r catalyst is the rate-limiting step for the whole reaction. The Gibbs free energy and overpotential (1.49 eV and 0.26 V) of the (Fe,La)Ni2P-r catalyst are lower than those of (Fe,Mn)Ni2P-r (1.76 eV and 0.53 V) and (Fe)Ni2P-r (2.26 eV and 1.03 V), reflecting the superior OER performance of the (Fe,La)Ni2P-r catalyst. Here, (Fe,Ni)OOH, (Fe,Mn,Ni)OOH, (Fe,La,Ni)OOH, and (Fe,La,Ni)OOH(s) represent the Ov-containing (Fe)Ni2P-r, (Fe,Mn)Ni2P-r, (Fe,La)Ni2P-r, and Ov-free (Fe,La)Ni2P-r catalysts, respectively. Combined with SEM, TEM, and XRD, it is evident that all catalysts retain a significant amount of Ni2P after reconstruction, indicating that the doping of heteroatoms does not completely oxidize Ni2P, while the generated MOOH (M = Fe, Mn, Ni, and La) exists on the surface of the catalyst. Unexpectedly, it was found that the doping of La generates a new species LaOOH, which exists in a crystalline form, indicating that amorphous FeNiOOH and crystalline LaOOH also formed a non-homogeneous interface at the catalyst surface.

Fig. 5D shows the differential charge distribution maps for all catalysts, where the red and blue regions represent electron accumulation and depletion, respectively. An intense electron density is observed around Ni in the (Fe,La)Ni2P-r catalyst, indicating a strong coupling effect of electrons at the interface, which plays a crucial role in modulating the electronic structure, consistent with XPS results.20 To determine the primary active species in the catalyst, adsorption energy calculations were performed for all possible sites. Benefiting from the interface effect and the weakening effect of oxygen vacancies on the limiting step barrier, Ni sites exhibit a lower Gibbs free energy and overpotential, with optimized intermediate models shown in Fig. S15. It is noteworthy that different active sites have different rate-limiting steps for adsorption. For example, the overpotential of the Fe site in the *O to *OOH step is the lowest, with the lowest energy barrier (Fig. S16). Thus, the experimental results are well confirmed by DFT calculations. The addition of Fe and La forms heterointerfaces and promotes the enhancement of oxygen vacancies, leading to an increase in the number of Ni primary active sites. The synergistic effect of the sharply increased Ni active sites and complementary multiple active sites is also one of the main reasons for the enhancement of OER activity.

3. Conclusions

In summary, based on electrochemical oxidation strategies, we introduced bimetallic Fe and La into the reconstruction process of a Ni2P substrate and developed a nanosheet assembly structure with local protrusions. The electrocatalyst exhibits a low overpotential of 197 mV at a current density of 10 mA cm−2, with a Faraday efficiency of about 100%. This excellent performance can be attributed to three aspects: (i) due to the unique structure, the synthesized catalyst has more active sites and excellent electrical conductivity, which promotes the mass transfer process, accelerates the intrinsic kinetics of the catalyst and promotes the release of oxygen. (ii) The redistribution of electrons at the interface and the synergistic effect between multiple active sites cause changes in the electronic structure, optimizing the adsorption and desorption strength of interface sites and optimizing the binding energy of different intermediates on the surface. (iii) The introduction of La induces the enhancement of Ov, thus changing the adsorption energy of H and OH at the active sites and increasing the content of *OOH, and the generated intermediate species *OOH was easily adsorbed on the surface of the (Fe,La)Ni2P-r catalyst, improving the OER activity of the catalyst. In addition, the synergistic action of the primary active sites and complementary multiple active sites also promoted the OER. This study provides an effective strategy for constructing highly efficient multi-interface non-precious metal OER electrocatalysts rich in oxygen vacancies.

Data availability

The data supporting this article have been included as part of the ESI.

Author contributions

D. Wang conceived and performed the experiments, collected and analyzed the data, and wrote the paper. Z. Sun and C. He performed the DFT calculation and participated in the discussion. W. Cui participated in the discussion. Z. Zhao conceived the idea, supervised the project work, led the data analysis and discussion, and prepared and finalized the paper. All the authors commented on the manuscript and have given approval to the final version of the manuscript.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This study was supported by the National Natural Science Foundation of China (21978030) and the Ministry of Education's New Century Excellent Talents Program for Universities (NCET-12-0079).

Notes and references

  1. W. Xiang, N. Yang, X. Li, J. Linnemann, U. Hagemann, O. Ruediger, M. Heidelmann, T. Falk, M. Aramini, S. DeBeer, M. Muhler, K. Tschulik and T. Li, Nat. Commun., 2022, 13, 179 CrossRef CAS PubMed .
  2. L. Gao, X. Cui, C. D. Sewell, J. Li and Z. Lin, Chem. Soc. Rev., 2021, 50, 8428–8469 RSC .
  3. Z.-L. Wang, D. Xu, J.-J. Xu and X.-B. Zhang, Chem. Soc. Rev., 2014, 43, 7746–7786 RSC .
  4. Y. Li, H. Wang, C. Priest, S. Li, P. Xu and G. Wu, Adv. Mater., 2020, 33, 2000381 CrossRef PubMed .
  5. F. Zeng, C. Mebrahtu, L. Liao, A. K. Beine and R. Palkovits, J. Energy Chem., 2022, 69, 301–329 CrossRef CAS .
  6. S. Li, S. Zhao, F. Hu, L. Li, J. Ren, L. Jiao, S. Ramakrishna and S. Peng, Prog. Mater. Sci., 2024, 145, 101294 CrossRef CAS .
  7. P. Liu and J. Rodriguez, J. Am. Chem. Soc., 2005, 127, 14871–14878 CrossRef CAS PubMed .
  8. S.-H. Li, M.-Y. Qi, Z.-R. Tang and Y.-J. Xu, Chem. Soc. Rev., 2021, 50, 7539–7586 RSC .
  9. H. Liao, T. Luo, P. Tan, K. Chen, L. Lu, Y. Liu, M. Liu and J. Pan, Adv. Funct. Mater., 2021, 31, 2102772 CrossRef CAS .
  10. K. Kawashima, R. A. Márquez, L. A. Smith, R. R. Vaidyula, O. A. Carrasco-Jaim, Z. Wang, Y. J. Son, C. L. Cao and C. B. Mullins, Chem. Rev., 2023, 123, 12795–13208 CrossRef CAS PubMed .
  11. D. Wang, M. Liu, Z. Sun, C. Zhang, W. Cui, C. He and Z. Zhao, Catal. Sci. Technol., 2023, 13, 6625–6630 RSC .
  12. G. Wu, X. Zheng, P. Cui, H. Jiang, X. Wang, Y. Qu, W. Chen, Y. Lin, H. Li, X. Han, Y. Hu, P. Liu, Q. Zhang, J. Ge, Y. Yao, R. Sun, Y. Wu, L. Gu, X. Hong and Y. Li, Nat. Commun., 2019, 10, 4855 CrossRef PubMed .
  13. Y. Zhao, J. Zhang, Y. Xie, B. Sun, J. Jiang, W.-J. Jiang, S. Xi, H. Y. Yang, K. Yan, S. Wang, X. Guo, P. Li, Z. Han, X. Lu, H. Liu and G. Wang, Nano Lett., 2021, 21, 823–832 CrossRef CAS PubMed .
  14. L. Zhao, M. Wen, Y. Tian, Q. Wu and Y. Fu, J. Energy Chem., 2022, 74, 203–211 CrossRef CAS .
  15. M. Liu, Z. Sun, S. Li, X. Nie, Y. Liu, E. Wang and Z. Zhao, J. Mater. Chem. A, 2021, 9, 22129–22139 RSC .
  16. C.-F. Li, H.-B. Tang, J.-W. Zhao and G.-R. Li, J. Mater. Chem. A, 2023, 11, 5841–5850 RSC .
  17. X. Li, Z. Kou, S. Xi, W. Zang, T. Yang, L. Zhang and J. Wang, Nano Energy, 2020, 78, 105230 CrossRef CAS .
  18. L. Xu, Q. Jiang, Z. Xiao, X. Li, J. Huo, S. Wang and L. Dai, Angew. Chem., Int. Ed., 2016, 55, 5277–5281 CrossRef CAS PubMed .
  19. J. Bao, X. Zhang, B. Fan, J. Zhang, M. Zhou, W. Yang, X. Hu, H. Wang, B. Pan and Y. Xie, Angew. Chem., Int. Ed., 2015, 54, 7399–7404 CrossRef CAS PubMed .
  20. H. Liu, H. Duan, J. Yu, C. Qiu, R. Yu, J. Gao, S. Li, X. Du, Z. Si and S. Yang, ACS Mater. Lett., 2022, 4, 2572–2578 CrossRef CAS .
  21. T. Zhou, Y. Yang, Y. Jing, Y. Hu, F. Yang, W. Sun and L. He, Chem. Sci., 2023, 14, 13453–13462 RSC .
  22. T. Dai, X. Zhang, M. Sun, B. Huang, N. Zhang, P. Da, R. Yang, Z. He, W. Wang, P. Xi and C. H. Yan, Adv. Mater., 2021, 33, 2102593 CrossRef CAS PubMed .
  23. W. Q. Han, Z. H. Liu, Y. B. Pan, G. N. GuoJ, X. Zou, Y. Xia, Z. M. Peng, W. Li and A. G. Dong, Adv. Mater., 2020, 32, 2002584 CrossRef CAS PubMed .
  24. J. Kwon, H. Han, S. Jo, S. Choi, K. Y. Chung, G. Ali, K. Park, U. Paik and T. Song, Adv. Energy Mater., 2021, 11, 2100624 CrossRef CAS .
  25. Y. Zhao, X. Jia, G. Chen, L. Shang, G. I. N. Waterhouse, L.-Z. Wu, C.-H. Tung, D. O'Hare and T. Zhang, J. Am. Chem. Soc., 2016, 138, 6517–6524 CrossRef CAS PubMed .
  26. X. Li, H. Wu, A. M. Elshahawy, L. Wang, S. J. Pennycook, C. Guan and J. Wang, Adv. Funct. Mater., 2018, 28, 1800036 CrossRef .
  27. D. Yu, Y. Hao, S. Han, S. Zhao, Q. Zhou, C. H. Kuo, F. Hu, L. Li, H. Y. Chen, J. Ren and S. Peng, ACS Nano, 2023, 17, 1701–1712 CrossRef CAS PubMed .
  28. M. Liu, Z. Sun, C. Zhang, S. Li, C. He, Y. Liu and Z. Zhao, J. Mater. Chem. A, 2022, 10, 13410–13417 RSC .
  29. G. Zhang, B. Wang, J. Bi, D. Fang and S. Yang, J. Mater. Chem. A, 2019, 7, 5769–5778 RSC .
  30. L. Li, K. Chao, X. Liu and S. Zhou, J. Alloys Compd., 2023, 941, 168952 CrossRef CAS .
  31. S. Zhao, F. Hu, L. Yin, L. Li and S. Peng, Sci. Bull., 2023, 68, 1389–1398 CrossRef CAS PubMed .
  32. R. Li, B. Hu, T. Yu, H. Chen, Y. Wang and S. Song, Adv. Sci., 2020, 7, 1902830 CrossRef CAS PubMed .
  33. C. Dang Van, S. Kim, M. Kim and M. H. Lee, ACS Sustainable Chem. Eng., 2023, 11, 1333–1343 CrossRef CAS .
  34. L. Gao, Y. Li, G. Li, J. Huo and L. Jia, Chem. Eng. J., 2023, 451, 138743 CrossRef CAS .
  35. S. H. Ye, Z. X. Shi, J. X. Feng, Y. X. Tong and G. R. Li, Angew. Chem., Int. Ed., 2018, 57, 2672–2676 CrossRef CAS PubMed .
  36. Y. Wang, L. Yan, K. Dastafkan, C. Zhao, X. Zhao, Y. Xue, J. Huo, S. Li and Q. Zhai, Adv. Mater., 2021, 33, 2006351 CrossRef CAS PubMed .
  37. X. Xiao, D. Huang, Y. Fu, M. Wen, X. Jiang, X. Lv, M. Li, L. Gao, S. Liu, M. Wang, C. Zhao and Y. Shen, ACS Appl. Mater. Interfaces, 2018, 10, 4689–4696 CrossRef CAS PubMed .
  38. K. Yu, H. Yang, H. Zhang, H. Huang, Z. Wang, Z. Kang, Y. Liu, P. W. Menezes and Z. Chen, Nano-Micro Lett., 2023, 15, 186 CrossRef CAS PubMed .
  39. L. Wang, L. Song, Z. Yang, Y. M. Chang, F. Hu, L. Li, L. Li, H. Y. Chen and S. Peng, Adv. Funct. Mater., 2023, 33, 2210322 CrossRef CAS .
  40. F. Hu, D. Yu, M. Ye, H. Wang, Y. Hao, L. Wang, L. Li, X. Han and S. Peng, Adv. Energy Mater., 2022, 12, 2200067 CrossRef CAS .
  41. J. Huang, Y. Li, Y. Zhang, G. Rao, C. Wu, Y. Hu, X. Wang, R. Lu, Y. Li and J. Xiong, Angew. Chem., Int. Ed., 2019, 58, 17458–17464 CrossRef CAS PubMed .
  42. P. W. Menezes, S. Yao, R. Beltrán-Suito, J. N. Hausmann, P. V. Menezes and M. Driess, Angew. Chem., Int. Ed., 2021, 60, 4640–4647 CrossRef CAS PubMed .
  43. D. T. Tran, H. T. Le, V. H. Hoa, N. H. Kim and J. H. Lee, Nano Energy, 2021, 84, 105861 CrossRef CAS .
  44. P. Wen, F. Su, H. Li, Y. Sun, Z. Liang, W. Liang, J. Zhang, W. Qin, S. M. Geyer, Y. Qiu and L. Jiang, Chem. Eng. J., 2020, 385, 123878 CrossRef CAS .
  45. Y. Duan, Z. Y. Yu, S. J. Hu, X. S. Zheng, C. T. Zhang, H. H. Ding, B. C. Hu, Q. Q. Fu, Z. L. Yu, X. Zheng, J. F. Zhu, M. R. Gao and S. H. Yu, Angew. Chem., Int. Ed., 2019, 58, 15772–15777 CrossRef CAS PubMed .

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4cy00653d

This journal is © The Royal Society of Chemistry 2024