Kaixi
Wang
ab,
Yifei
Xu
a,
Vahid
Daneshvariesfahlan
b,
Moniba
Rafique
c,
Qiang
Fu
*d,
Hang
Wei
e,
Yumin
Zhang
a,
Jiheng
Zhang
b,
Bing
Zhang
b and
Bo
Song
*abcdf
aSchool of Astronautics, Harbin Institute of Technology, Harbin, 150001, China
bZhengzhou Research Institute, Harbin Institute of Technology, Zhengzhou 450046, China
cNational Key Laboratory of Science and Technology on Advanced Composites in Special Environments, Harbin Institute of Technology, Harbin, 150001, China. E-mail: songbo@hit.edu.cn; Tel: +86-451-86403753 Web: https://homepage.hit.edu.cn/pages/songbo
dSchool of Physics, Harbin Institute of Technology, Harbin, 150001, China
eCollege of Chemistry and Chemical Engineering, Inner Mongolia Engineering and Technology Research Center for Catalytic Conversion and Utilization of Carbon Resource Molecules, Inner Mongolia University, Hohhot, 010021, China
fLaboratory for Space Environment and Physical Sciences; Frontiers Science Center for Matter Behave in Space Environment, Harbin Institute of Technology, Harbin, 150001, China
First published on 5th February 2025
The oxygen-evolution reaction (OER) is an indispensable component of various energy storage and conversion electrocatalytic systems. However, the slow reaction kinetics have forced the development of advanced, efficient, and inexpensive OER electrocatalysts to break through the bottleneck of its application. Recently, the structural reconstruction of precatalysts has provided a promising avenue to boost the catalytic activity of electrocatalysts. Structural reconstruction implies atomic rearrangement and composition change of the pristine catalytic materials, which is a very complex process. Therefore, it is very crucial to have a deep understanding of the reconstruction chemical process and then modulate the reconstruction by deliberate design of electrochemical conditions and precatalysts. However, a systematic review of the structural reconstruction process, research methods, influencing factors and structure–performance relationship remains elusive, significantly impeding the further developments of efficient electrocatalysts based on structural reconstruction chemistry. This critical review is dedicated to providing a deep insight into the structural reconstruction during alkaline water oxidation, comprehensively summarizing the basic research methods to understand the structural evolution process and various factors affecting the structural reconstruction process, and providing a reference and basis for regulating the dynamic reconstruction. Moreover, the impact of reconstruction on the structure and performance is also covered. Finally, challenges and perspectives for the future study on structural reconstruction are discussed. This review will offer future guidelines for the rational development of state-of-the-art OER electrocatalysts.
Over the past decade, many advanced electrocatalytic systems have been developed for efficient energy storage and conversion, including producing high-purity green hydrogen through water electrolysis,12–14 obtaining high-value chemicals or fuels via the electrochemical CO2 reduction reaction (CO2RR),15,16 ammonia production from the nitrogen reduction reaction (NRR),17,18 and rechargeable metal–air batteries.19,20 These fundamental electrocatalytic reactions and devices further advance the chemical industry and the field of new energy vehicles (Fig. 1). Typically, electrocatalytic transformations usually consist of two independent half-reactions, namely the small molecule reduction reaction at the cathode and the oxygen evolution reaction (OER) at the anode.21,22 Therefore, the OER is an indispensable component in these electrocatalytic systems. The OER involves a four-electron transfer, which is a more complex process, leading to slow reaction kinetics and making the anode overpotential much higher than that of the cathode.23,24 Thus, developing advanced, efficient, and inexpensive OER electrocatalysts is of great theoretical and practical significance for breaking through the bottleneck of large-scale application of electrocatalytic energy storage and conversion systems.25
To date, many transition metal alloys26–29 and compounds have been explored for the OER, such as oxides/hydroxides,24,30,31 phosphides,32–34 chalcogenides,35,36 nitrides,37,38 perovskites,2,39 metal–organic frameworks,40,41etc. Numerous experimental and theoretical studies have shown that most of these materials act merely as precatalysts and that they will undergo structural reconstruction at high oxidation potentials during the OER process.42–44 Structural reconstruction implies atomic rearrangement, which usually leads to specific changes in the morphology, structure, and composition of the pristine catalytic materials.2,11 Previous studies have indicated that the precatalyst-derived metal oxides/(oxy)hydroxides serve as the actual active species and stable phases towards the OER.45,46 It should be noted that the reconstruction of catalysts is a complex process due to the complicated reaction conditions, and the reconstructed structure is not always conducive to the improvement of catalytic performance. It depends on the structure and properties of the new active components, including corrosion resistance, binding strength with OER active intermediates, electronic configuration, conductivity, etc.47,48 In this regard, precise control of precatalyst reconstruction is essential for optimizing the OER performance, and it requires a deep understanding of the reconstruction chemical processes, and then regulates the reconstruction by deliberate design of electrochemical conditions and precatalysts. Although some reviews about the structural reconstruction during the OER have been reported, most of them focused on summarizing the reconstruction phenomenon, typical precatalysts, advanced characterization and the regulation strategies of reconstruction.2,36,41,44,48–54 Currently, a systematic review of the structural reconstruction process, research methods, influencing factors and structure–performance relationship is unavailable.
This review aims to provide a deep insight into structural reconstruction, which is critical for developing superior OER electrocatalysts. First, basic research methods to understand the structural evolution process are introduced. Subsequently, various factors affecting the structural reconstruction process are summarized, including external reaction conditions as well as the structure and component of the precatalyst itself, which provides a reference and basis for the regulation of the reconstruction process. In addition, this review also covers the impact of reconstruction on the structure and performance. Finally, challenges and perspectives for the future study on structural reconstruction are discussed. We believe this review can provide future guidelines for the rational development of advanced OER electrocatalysts.
![]() | ||
Fig. 2 Schematic illustration of various research methods for the structural reconstruction process. Reproduced with permission from ref. 55. Copyright 2020, Royal Society of Chemistry. |
Types | Precatalysts | Electrolyte | Reconstructed (active) species | Reconstruction type | Research methods | Ref. |
---|---|---|---|---|---|---|
Oxides/hydroxides | β-Ni(OH)2 | 1 M KOH | β-NiOOH/Ni1−xO | Surface reconstruction | In situ TEM, Raman | 57 |
NiMoFeO@NC | 1 M KOH | NiFeOOH/NiFe-LDH | Deep reconstruction | In situ Raman | 62 | |
Mn2O3 | 1 M KOH | MnOx | Surface reconstruction | In situ TEM | 68 | |
β-Co(OH)2 | 0.1 M KOH | CoOOH | Surface reconstruction | Operando EC-AFM, STXM | 69 | |
Co3O4 | 0.1 M KPi | CoOx(OH)y | Surface reconstruction | In situ grazing-incident XRD and XAS | 86 | |
VO-Co3O4 | 1 M KOH | Co-OOH˙ species | Surface reconstruction | Operando EIS, XAS, and quasi-operando XPS | 67 | |
NiCeOxHy | 1 M KOH | γ-NiOOH | Complete reconstruction | In situ Raman, XRD | 87 | |
Co3O4/Co(OH)2 | 1 M KOH | CoO(OH) | Surface reconstruction | Operando APXPS | 90 | |
NiFeOxHy | 0.1 M KOH | Fe-NiOxHy | Surface reconstruction | In situ ICP-MS, XAS | 102 | |
Chalcogenides | CoSx | 1 M KOH | CoOOH | Complete reconstruction | In situ TEM, FTIR | 65 |
Co9S8@Fe3O4 | 1 M KOH | CoOOH@Fe3O4 | Surface reconstruction | In situ Raman, FTIR, XAS | 92 | |
Phosphides | V25%-Ni2P/NF-AC | 1 M KOH | β-NiOOH | Surface reconstruction | Operando EIS | 66 |
F–Fe-CoP | 1 M KOH | F–Fe–CoOOH | Complete reconstruction | In situ Raman | 144 | |
Carbide | Co3C | 1 M NaOH | CoOx | Complete reconstruction | — | 118 |
Nitrides | NiMoN@NiFeN | 1 M KOH + seawater | NiFe oxides/(oxy)hydroxides, Ni(OH)2 | Surface reconstruction | In situ Raman | 58 |
Co3−xFexMo3N | 1 M KOH | CoFeMoOOH | Surface reconstruction | — | 59 | |
Borides | NixB | 1 M KOH | Nickel oxyhydroxide | Surface reconstruction | Operando XAS | 60 |
Ir/CoNiB | 1 M KOH | IrOx-oxides/(oxy)hydroxides | Surface reconstruction | In situ Raman, XAS, DEMS | 61 | |
MOFs | ZIF-67 | 1 M KOH | CoOOH/Co(OH)2 | Complete reconstruction | In situ UV-vis, Raman | 114 |
CoFe2O4@CoBDC | 1 M KOH | CoOxHy | Complete reconstruction | In situ Raman | 76 | |
Alloys | FeNi3 and NiCu | 1 M KOH | Fe doped NiOOH and Cu doped NiOOH | Complete reconstruction | Operando ATR FT-IR | 93 |
CrMnFeCoNi | 0.05 M KOH | NiFe-rich oxyhydroxide/Mn-rich oxide | Surface reconstruction | SFC-ICP-MS | 101 | |
Others | Sr2CoO3−xF | 1 M KOH | CoOOHx | Deep reconstruction | In situ Raman, XAS, DEMS | 77 |
Co2(OH)3Cl@NiMoO4 | 1 M KOH | CoOOH@NiOOH | Complete reconstruction | In situ Raman, FTIR | 91 |
![]() | ||
Fig. 3 (a) CV curves at a scan rate of 5 mV s−1 after 80% iR-correction in 1 M KOH. (b) Cdl and ECSA values after different numbers of CV cycles under an OER potential window of 1.1–1.6 V. Reproduced with permission from ref. 57. Copyright 2024, American Chemical Society. (c) LSV curves of OER catalysis on NiMoFeO@NC after different LSV scans in 1 M KOH. Reproduced with permission from ref. 62. Copyright 2020, Elsevier. (d) Chronopotentiometry curve of CoSx at a low anodic current density of 0.5 mA cm−2. Reproduced with permission from ref. 65. Copyright 2018, American Chemical Society. (e) Operando EIS measurement of V25%-Ni2P/NF-AC at different applied potentials versus RHE in 1 M KOH. Reproduced with permission from ref. 66. Copyright 2021, Wiley-VCH. (f and g) BSE-STEM images of β-Ni(OH)2 before and after structural reconstruction. (h and i) TEM images of β-Ni(OH)2 before and after structural reconstruction. Reproduced with permission from ref. 57. Copyright 2024, American Chemical Society. (j) In situ TEM sequential images showing the evolution of both surface layer and oxygen nanobubble associated with the OER on the Mn2O3 nanocatalyst surface. Reproduced with permission from ref. 68. Copyright 2021, American Chemical Society. (k) Operando EC-AFM images of a β-Co(OH)2 particle in 0.1 M KOH at different applied voltages. Scale bars, 500 nm. (l) Differential height compared to the particle morphology at the open-circuit voltage (0.96 V). Scale bar: 500 nm. Reproduced with permission from ref. 69. Copyright 2021, Springer Nature. |
Another differentiation criterion is the stability test at a constant potential for a long time using the CP technique.64,65 Taking the CoSx precatalyst as an example, Fan et al. confirmed the transformation process of Co species from CoSx to Co(OH)2 and then to CoOOH through analyzing the chronopotentiometry curve of CoSx at an anodic current density of 0.5 mA cm−2.65 As shown in Fig. 3d, the CP curve can be divided into three parts: the electrochemical transformation of CoSx into the Co(OH)2 intermediate takes place in the first part (∼0–0.6 h). This is followed by the second part (∼0.6–1.1 h), where Co(OH)2 experiences ion intercalation and converts to the CoOOH phase prior to reaching OER conditions. Finally, the derived CoOOH serves as the true catalytic species for a stable OER in the third part (above ∼1.1 h). Besides, EIS is also used to track the structural reconstruction process, as reconstructed structures typically exhibit different electron transfer rates.66,67 For instance, Zhao et al. conducted operando EIS measurements of V25%-Ni2P/NF-AC to study the optimized adsorption of *OH reaction intermediates during the OER (Fig. 3e).66 The smaller charge transfer resistance (Rct) with increasing applied potential suggests the evolution of *OH intermediates on the catalyst surface.
Ex situ characterization techniques usually only provide indirect information about OER process studies, and are unable to perform real-time monitoring and capture reaction intermediates during the catalytic process, thus limiting the in-depth understanding of the reaction process. Therefore, in situ research methods are indispensable for an accurate survey of the reconstruction and can offer guidance for the rational design of high-performance electrocatalysts.71In situ TEM allows direct observation of the catalytic process and catalyst evolution.65,68,72,73 Zhu's group utilized an in situ liquid holder in a TEM to unravel the real-time formation of a surface layer on Mn2O3 and surrounding oxygen nanobubbles.68 As shown in Fig. 3j, overall, the volume of the nanobubble increases with time. However, the bubble does not grow continuously but displays volume oscillation, which is due to the competition between oxygen evolution and dissolution. In this way, the volume change of O2 bubbles can be used as an indicator to evaluate the OER rate. Meanwhile, it can be easily observed that nucleation occurs on the surface layer, which then extends across the entire surface of the nanoparticle. Throughout this process, the full size of the nanoparticle remains nearly unchanged, suggesting that the surface layer is formed via the reaction of surface Mn2O3, rather than through the deposition of an overlayer. In addition, the surface layer also exhibits oscillatory growth, indicating a partially reversible surface restructuring during the OER.
As a widely accepted non-contact surface analysis technique, AFM employs an atomically sharp tip to scan the sample surface. During this process, the tip detects subtle changes in the forces between the tip and the surface atoms, and by tracking the position changes of the tip attached to a microcantilever, it achieves atomic resolution imaging of the sample surface. This technique can detail the 3D topography of the sample surface. Mefford et al. used electrochemical atomic force microscopy (EC-AFM) effectively to investigate the in situ 3D morphology transformation of CoOxHy with voltage during the OER process.69 As depicted in Fig. 3k and l, the particle morphology varies non-monotonically with voltage during oxidation. It is disclosed that the catalysts’ lateral expansion reached a maximum at an intermediate potential of 1.39 V, and subsequently, the dimensions reverted to almost their original state as the applied potential was raised to 1.58 V. Finally, with a further increase in the voltage, the particle starts to shrink from the outer edges and moves towards the center of the particle.
![]() | ||
Fig. 5 Schematic illustration of the in situ (a) Raman spectroscopy setting and (b) NEXAFS setup. Reproduced with permission from ref. 100. Copyright 2022, Springer Nature. (c) In situ Raman spectra acquired under various applied potentials during the OER on pristine β-Ni(OH)2 after 1000 CV cycles. Reproduced with permission from ref. 57. Copyright 2024, American Chemical Society. Operando XAFS for Co K-edge of (d) pure Co3O4 and (e) VO-Co3O4. The insets show a detailed view of the dotted boxes, respectively. (f) Structural coherence change in the EXAFS coordination number of Co ions under an applied potential relative to the OCP state. Reproduced with permission from ref. 67. Copyright 2020, American Chemical Society. |
Apart from tracking the transformation of surface phases, the detection of oxygenated intermediates, such as *OOH or *O–O, provides a deeper understanding of the fundamental mechanisms in the OER. The detailed mechanistic understanding is closely related to the surface reconstruction process, as the formation of active species often accompanies structural changes on the catalyst surface. However, these intermediates typically exhibit weak Raman signals, making it difficult to collect useful information. In light of this, surface-enhanced Raman spectroscopy (SERS), known for its heightened sensitivity and spatial resolution, has been utilized for the in situ characterization of OER catalysts.95–97 For example, Hu et al. constructed bifunctional Au@Ni3FeOx core-satellite superstructures to study the interfacial OER process on the Ni3FeOx catalyst.97 The Au core plays the role of a SERS enhancing substrate in this structure. The SERS data indicate that the Fe atoms serve as the active centers for the initial oxidation of OH− to O–O−. The O–O− species, adsorbed between adjacent Fe and Ni sites, undergo further oxidation due to electron transfer to NiIII, leading to the formation of the final O2 product.
Xiao and coworkers performed in situ XAS to identify of the dynamic reconstruction behavior of oxygen vacancy-rich Co3O4 (VO-Co3O4) for the OER.67 The Co K-edge data were recorded from the open circuit potential (OCP) to 1.75 V vs. RHE (Fig. 5d and e). As can be seen, the edge peaks of pure Co3O4 and VO-Co3O4 around 7719 eV show a similar positive shift trend with increasing applied potential, indicating the oxidation of Co. However, the oxidation rate of Co ions in VO-Co3O4 is faster than that in pure Co3O4. According to the previous literature,99 it means that VO will facilitate the adsorption of OH ions on the Co sites and subsequent deprotonation process to form reactive oxygen species (Co–OOH˙). Then, the analysis of relative changes in coordination number also reveals that VO can promote cobalt pre-oxidation and structural reconstruction during the OER. In addition, the oxygen species involved in the OER process for the pure Co3O4 and VO-Co3O4 are mostly related to the octahedral cobalt coordinated oxygen and the tetrahedral cobalt coordinated oxygen, respectively (Fig. 5f). Very recently, Shao's group reported the establishment of the relationship between oxygen-evolving performance and operational structural properties on model oxides through advanced operando characterization.77 They found that the pyramidal structure is more vulnerable to OH− attack than tetrahedral and octahedral structures due to its inherent unsaturation and asymmetry as well as its constant single-electron occupancy on the active z2 orbital during reaction, which facilitates the transformation from the surface to the bulk, resulting in the formation of active, amorphous, high-valence CoOOHx with edge-sharing structures. Operando soft XAS demonstrates that the non-uniform dehydrogenation process becomes more difficult with time (Co3+OOH → Co3+/4+OOHx → Co4+OO) due to the increased covalency of Co–O with a higher energy barrier. Lattice oxygen participates in the formation of active CoOOHx at the expense of stability.
In situ XPS can provide real-time information about the chemical state and electronic structure changes on and near the catalyst surface under reaction conditions. Favaro et al. reported the utilization of operando ambient-pressure XPS (APXPS) to study the OER mechanism and structural reconstruction on CoOx (Fig. 6a–c).90 Through spectral simulation and multiplet fitting, it has been discovered that the catalyst experiences chemical and structural changes in response to the applied anodic potential. The Co(OH)2 surface layer in the biphasic catalyst facilitates structural transformation, including complete oxidative conversion to CoO(OH) and partial conversion of the underlying spinel Co3O4 to CoO(OH), making the active phase thick enough to provide high concentrations of catalytic sites. In contrast, for the monophasic catalyst, only a small portion of the surface undergoes partial conversion of Co3O4 to Co(OH)2 and only a thin layer of CoO(OH) is formed at the OER potential. In addition, Xiao et al. used quasi-operando XPS to obtain the variation of the Co2+/Co3+ ratio in cobalt oxide at different potentials, revealing the reconstruction process of cobalt oxide during the OER.67
![]() | ||
Fig. 6 (a) Three-electrode electrochemical setup used for the operando electrochemical APXPS characterization of the OER electrocatalyst. (b) Co 2p3/2 APXPS core levels acquired at the OCP and under OER conditions for two different CoOx catalysts. The green-shaded component at low BE, whose nature is explained in the text and assigned to Co4+, is observable only at OER potentials. (c) Chemical composition and subsurface structure for the monophasic and biphasic catalysts, passing from pristine to OER conditions. Reproduced with permission from ref. 90. Copyright 2017, American Chemical Society. (d) Schematic illustration of the in situ ATR-FTIR electrochemical cell. (e) Referred line plot of in situ FT-IR spectroscopy of Co2(OH)3Cl@NiMoO4 in 1 M KOH during the OER. Reproduced with permission from ref. 91. Copyright 2024, Springer Nature. (f) Schematic diagram of the operando O18 isotope experiment (top panel) and measured results for model oxides (bottom panel) in 1 M KOH. Reproduced with permission from ref. 77. Copyright 2024, Wiley-VCH. |
In addition to the above-mentioned spectroscopic analysis techniques, FTIR is also valid for probing the changes of chemical bonds during reconstruction. Most solid electrodes are impervious to infrared light, so in situ FTIR tests are in most cases in reflectance mode, such as attenuated total reflectance (ATR)-FTIR (Fig. 6d).56 For example, Yu's group employed in situ FTIR and other ex situ characterization techniques to observe the structural evolution of the amorphous electrocatalyst CoSx into crystallized CoOOH in the OER directly.65 Ji and coworkers deployed in situ synchrotron radiation-based FTIR (SR-FTIR) analysis to investigate the key active intermediates during OER.92 Recently, Ma et al. performed in situ FTIR to capture the signals of CoO, Co–O and Mo
O bonds in the OER (Fig. 6e). The significant enhancement of the Co
O signal indicates the conversion from Co2(OH)3Cl to CoOOH caused by Cl− leaching. Furthermore, the enhancement of Mo
O bond strength should come from the leaching of MoO42− derived from NiMoO4 reconstruction.91
The mass spectrometry technique can identify and quantify compounds in a sample by measuring the mass and charge of the sample molecules or atoms. By integrating mass spectrometry with an electrochemical cell, electrochemical MS enables the real-time and in situ monitoring of electrochemical species, including reactants, intermediates, and products in the electrocatalytic process. Luan et al. elucidated the structure–activity–stability relationship of CrMnFeCoNi toward the OER using an online scanning flow cell inductively coupled plasma mass spectrometer (SFC-ICP-MS), providing insights into the activation and degradation mechanism of high-entropy alloy (HEA) electrocatalysts.101 Markovic's group used a method of coupling in situ ICP-MS with isotopic labeling, and confirmed the concomitant dissolution and redeposition of Fe on the NiFeOxHy electrode during the OER, which is critical for the formation of a stable electrocatalyst–electrolyte interface.102 Differential electrochemical mass spectrometry (DEMS) is commonly used to detect volatile or gaseous species to study the OER mechanism. Shao and collaborators carried out operando O18 isotope labeling experiments to directly evidence the lattice oxygen participation mechanism (LOM) on model coordinated oxides (Fig. 6f).77 They labeled the partial lattice oxygen (O16) with O18 isotope through electrochemical treatment, and then detected a higher intensity of the O16O18 signal on pyramidal Sr2CoO3−xF, indicating its strongest LOM process in the OER.
For the metal-based precatalysts, given that their dynamic structural transformation typically results in (oxy)hydroxides, which are widely regarded as the true active species in the OER, it is essential to construct reasonable theoretical models to analyze the catalytic process. The structural models are usually determined based on the results of the in situ analyses described above. In addition to static DFT calculations, certain studies have implemented ab initio molecular dynamics (AIMD) to investigate the catalysts’ dynamic behavior. For example, Zagalskaya et al. conducted AIMD-based simulations to examine the dissolution of Ir at the IrO2(110)/water interface.105 They disclosed that IrO2OH species are produced on the surface, which are thermodynamically stable under varying electrode potentials and can be transformed into IrO3 at high anodic potentials. Conversely, at low anodic potentials, IrIII is generated on the surface that can be reoxidized back to IrO2, and continue to be adsorbed as IrIII, or dissolved into the solution as Ir(OH)3. Zhou et al. studied the role of Fe species on NiOOH in the OER using AIMD and found that adsorption and intercalation of the Fe ion on NiOOH can introduce proton-coupled electron transfer, which significantly reduces the overpotential and promotes the OER.106
![]() | ||
Fig. 7 (a and b) Schemes of the reconstructed CoBP-Pi and CoBP-Bi in KPi and KBi electrolytes. The spheres with blue, red, purple–grey, and light-pink colors represent cobalt, oxygen, phosphorus, and hydrogen atoms, respectively. Reproduced with permission from ref. 107. Copyright 2022, Wiley-VCH. (c) CVs of the Ni45Fe55 catalyst in electrolytes with different pH values. Reproduced with permission from ref. 109. Copyright 2017, American Chemical Society. (d) OER activity comparison in 1 M KOH with different SeO32− concentrations. Reproduced with permission from ref. 112. Copyright 2020, Wiley-VCH. (e) The mechanism of surface reconstruction after being immersed in electrolytes with oxyanions. (f) HRTEM images of pristine LaNiO3−δ powders, after being immersed in 1 M KOH electrolyte for 24 hours and after being immersed in 1 M KOH with the addition of 0.1 M SO42− for 24 hours. Reproduced with permission from ref. 113. Copyright 2023, Wiley-VCH. (g) Illustration of the structural evolution of ZIF-67 during amperometry at the different potential values. Reproduced with permission from ref. 114. Copyright 2019, American Chemical Society. (h) Schematic of the thermal-induced complete reconstruction on the NiMoO4 nanowire. Reproduced with permission from ref. 115. Copyright 2020, Wiley-VCH. |
The pH value of the electrolyte is also known to have much influence on the reconstruction of various electrocatalysts.44,74 It can directly affect the redox potential of metal species, thereby altering the electrochemical performance.108 Görlin et al. revealed the effect of the electrolyte pH on the redox behavior and OER activity of the Ni–Fe(OOH) catalyst.109 As shown in Fig. 7c, the precatalytic voltammetric charge of the redox peak couple increased significantly with increasing pH. Meanwhile, the redox peak potential showed a pH-sensitive shift towards the cathodic region, thereby enhancing OER activity by 2–3 times. In addition, for certain metal–organic frameworks (MOFs), once the pH of the electrolyte exceeds the acceptable range of these materials, it will disrupt their structure and stimulate the metal oxidation steps, thus promoting the structural reconstruction.41,110 Furthermore, the pH could also affect the reconstruction degree. For example, unlike the surface reconstruction of anhydrous NiMoO4 in low concentration alkaline solution (0.1–1 M KOH), high concentration industrial alkaline solution (20–30 wt% KOH) causes it to undergo complete reconstruction.111
The electrolyte additives are another factor to facilitate the structural reconstruction of precatalysts. To verify the hypothesis that the surface selenate plays a crucial role in OER performance, Shi et al. tested the OER activity of Ni(OH)2 in 1 M KOH with the extra addition of SeO32−.112Fig. 7d shows that the OER activity improved with the addition of SeO32−, and the activity reached a maximum when the SeO32− concentration was increased to 0.1 M. Compared with Ni(OH)2 in pure KOH, the current density at an overpotential of 500 mV increased from 47 to 221 mA cm−2 in the presence of 0.1 M SeO32−, confirming the important contribution of surface-adsorbed chalcogenates to the OER activity. Theoretical calculations showed that the Gibbs free energy of the OER intermediates decreased after selenate adsorption, strongly substantiating the positive effect of the adsorbed selenate by promoting the adsorption of the OER intermediates. Moreover, it has been found that other additives, including SO42−, CO32−, NO3−, etc., can also affect the OER process. Tang et al. reported that the added oxyanions would be more prone to adsorb at the solid–liquid interface, which could break the dynamic balance between the adsorbed OH− ions and the release of OH− ions during the surface reconstruction process in the inner Helmholtz plane (IHP) layer. Consequently, the easier release of OH− ions into the electrolyte could expedite the surface reconstruction process (Fig. 7e). The HRTEM images show that a thicker amorphous layer of about 5 nm formed on the LaNiO3−δ surface after being immersed in electrolytes with 0.1 M SO42− (Fig. 7f), illustrating that the surface reconstruction process could be significantly promoted in the presence of oxyanions in 1 M KOH.113
Additionally, the electro-oxidation time is another key factor that affects the reconstruction process and degree. In the previous example, the authors also studied the correlation between amperometric time and OER activity.114 It was found that either too long or too short a treatment time for ZIF-67 was not conducive to the optimal activity of the reconstructed catalyst, and that only a suitable duration would provide acceptable sites for optimal activity. This is because even though prolonged durations can produce sufficient sites, these sites are predominantly present in a low activity form (β-Co(OH)2), resulting in a declining turnover frequency (TOF) and reduced activity. In contrast, some studies suggest that reconstruction after sufficient treatment time is important to achieve better OER activity.118 For example, Kim et al. reported that the amorphous CoOx reconstructed from Co3C after a sufficiently long oxidation time has a relatively low overpotential compared to the Co3C@CoOx core–shell structure obtained in a short treatment time, which is attributed to the increased ECSA during the complete oxidation of crystalline Co3C.118 Therefore, the need for sufficient treatment time may depend on whether the reconstructed structure and composition are highly active for the OER.
Operating at relatively high temperatures has been proven to be an effective method for accelerating the reconstruction process.49 High temperature can enhance mass transfer and diffusion kinetics. Mai et al. investigated the reconstruction process of the NiMoO4 nanowire at an industrial temperature and proposed a thermal-induced complete reconstruction strategy (Fig. 7h).115 A high temperature of 51.9 °C can promote the leaching of Mo species and deep penetration of the electrolyte into the bulk. Consequently, the NiMoO4 that undergoes surface reconstruction at room temperature can achieve complete reconstruction at high temperatures. The phenomenon of high temperature deepening the reconstruction degree has also been reported by other groups.119,120 For instance, Zhang et al. found that a higher testing temperature of NiCo2O4 can regulate the reversible transformation of spinel-to-oxyhydroxide active species for the OER.120
![]() | ||
Fig. 8 (a) The HRTEM image of Co2.75Fe0.25O4 after reconstruction. (b) The schematic illustration of the spin pinning effect at the interface between FM magnetic domains and the thin PM oxyhydroxide layer, and the spins in the PM oxyhydroxide layer can be aligned under magnetization. (c) LSV curves of Co3−xFexO4 (s) after reconstruction in 1 M KOH following different conditions. Reproduced with permission from ref. 122. Copyright 2021, Springer Nature. (d) Schematic diagram of the Co site configuration evolution model with CV activation. The yellow, blue, and red spheres represent Co2+, Co3+, and Co site with OV, respectively. (e) Schematic diagram of more Co2+ formation in CFO@CoOxHy. Reproduced with permission from ref. 76. Copyright 2024, Wiley-VCH. (f) CV curves with and without laser irradiation for Ni(OH)2 nanosheets and Ni(OH)2-Au hybrids. The inset is the normalized transformation of NiII/NiIII/IV on the basis of Ni(OH)2 nanosheets (dark). (g) CV curves with and without laser irradiation for Au nanoparticles supported by the GC electrode. The inset displays the enlarged oxidation peaks of Au. (h) Schematic electron transfer paths in the Ni(OH)2-Au electrode under laser irradiation. The dashed line corresponds to the Fermi level of the Au nanoparticle. Yellow, cyan, and orange balls represent Ni, O, and Au atoms, respectively. Reproduced with permission from ref. 126. Copyright 2016, American Chemical Society. |
Studies of the effect of magnetic fields on the OER have focused on stable or reconstructed catalysts; however, the effect of magnetic fields on the reconstruction process of catalysts has rarely been investigated. Recently, our group rationally designed a ferromagnetic/paramagnetic CoFe2O4@CoBDC (CFO@CoBDC) core–shell structure and systematically investigated its structural reconstruction process under magnetic fields.76 It should be pointed out that a local gradient magnetic field will be induced around the ferromagnetic CFO under an external magnetic field.125Fig. 8d depicts the possible structural evolution of the Co site configuration during CV activation for CFO@CoBDC. As can be seen, the paramagnetic Co2+ obtained by spontaneous hydrolysis of CoBDC in alkali would be subjected to both Coulomb force (FC) and Kelvin force (FK) during activation under a magnetic field, leading to directional aggregation and stacking relatively regularly on the CFO surface. By contrast, CFO@CoBDC/CV behaves more disordered as it is only affected by FC. Subsequently, a portion of Co2+ is irreversibly oxidized to Co3+ during the initial CV cycle. Finally, the Co2+ and Co3+ species undergo a sequence of redox reactions, forming CoOOH. In addition, even though CoOOH is the main species after activation, there is still a higher concentration of divalent Co in CFO@CoBDC/MCV under a gradient magnetic field, suggesting that OV is to be accompanied by the generation of the CoOOH phase to balance the charge (Fig. 8e). Thus, we demonstrated that the FK induced by a gradient magnetic field directionally modulated the surface reconstruction of CoBDC, resulting in more active Co2+ in derived CoOxHy. The Co sites with optimized electronic configuration exhibit moderate adsorption energy for oxygen-containing intermediates and lower the energy barrier of the overall catalytic reaction, thereby significantly enhancing the OER performance. This work showcases the directional structural reconstruction strategy for an improved OER.
Surface plasmon resonance (SPR) is another useful tool to regulate the reconstruction process during the OER.126,127 Liu et al. observed the plasmon-enhanced OER phenomenon on Au-nanoparticle-decorated Ni(OH)2 hybrids (Ni(OH)2-Au) after irradiation under a 532 nm laser.126 The increased integrated oxidation peak area from CV curves increases with laser irradiation (Fig. 8f), which suggests that the plasmonic excitation of Au nanoparticles enhances the oxidation of NiII to NiIII/IV active sites in Ni(OH)2–Au, thereby greatly improving the OER catalysis. Furthermore, the photoelectrochemical voltammetry (Fig. 8g) analysis directly indicates that SPR-excitation-induced hot-electron injection occurs on Au nanoparticles under laser irradiation. The authors put forward that during the plasmon-enhanced OER, the hot electrons produced on Au nanoparticles act as effective electron trappers to capture electrons from Ni(OH)2 and facilitate the reconstruction of inactive NiII to active NiIII/IV, enabling the OER. Concurrently, the hot electrons generated by plasmon effects are transferred to the glassy carbon (GC) electrode under the external potential (Fig. 8h). The SPR-enhanced OER catalysis has also been observed with CoO-Au and FeOOH-Au catalysts, highlighting the generality of this method.
For instance, Wu et al. synthesized a nickel foam (NF)-supported Ni3S2 nanosheet array with a thickness of 20 nm decorated with ultrasmall NixCo3−xS4 nanoparticles (3–5 nm in diameter) through a partial cation exchange process.132 It was observed that the tiny nanoparticles completely converted into hydroxides after the OER, while the underlying crystalline structure of the nanosheets remained unchanged. In another typical example, Liu et al. introduced a lithiation-induced deep reconstruction strategy to obtain NiO with a size of less than 10 nm, and then these ultrasmall nanoparticles were deeply reconstructed to NiOOH through electro-oxidation.129 As a comparison, only about a 5 nm-thick reconstructed NiOOH layer was generated on the surface of the bulk Ni precatalyst without lithiation, and the limited reconstruction degree is due to the impaired electrolyte permeation. Compared to the incompletely reconstructed sample, the as-fabricated completely reconstructed NiOOH achieved significantly enhanced mass activity and better stability owing to a significant increase in the number of catalytic species and the polycrystalline features with abundant defects.
Among various metal elements, Fe is the most commonly used. In a representative work, Wu et al. reported an Fe substitution approach to boost the surface reconstruction and OER activity of the inactive spinel CoAl2O4 (Fig. 9a).137 They found that the incorporation of Fe triggers the preliminary pre-oxidation of Co at a low potential, which not only promotes surface reconstruction but also facilitates the subsequent deprotonation process on the newly formed oxyhydroxide to induce negatively charged oxygen as an active site, thereby significantly increasing the OER activity of CoAl2O4 (Fig. 9b). Further research indicated that Fe substitution elevates the O 2p band centre, which contributes to VO formation in CoFe0.25Al1.75O4 with lattice oxygen oxidation. The VO accumulates on the oxide surface with great structural instability, inducing surface reconstruction into oxyhydroxides (Fig. 9c). They also investigated the reconstruction-terminating mechanism. It was revealed that the leaching of Al lowered the O 2p energy level of the oxide, leading to the termination of lattice oxygen oxidation, thus preventing further reconstruction as VO was no longer produced (Fig. 9d). Similarly, using LaNi1−xFexO3 perovskite oxides as model catalysts, An et al. explored the influence of Fe substitution on the surface reconstruction process.133 It was demonstrated that a low Fe content in LaNi1−xFexO3 significantly accelerates the reconstruction rate and improves the electrocatalytic activity. Nevertheless, the reconstruction degree of LaNi1−xFexO3 does not align with its OER activity. The volcano-shaped activity trend and the thinner reconstructed layer with increasing Fe substitution reveal that the key factor to determine the activity of reconstructed surfaces is the composition of the LaNi1−xFexO3 perovskite, instead of the surface reconstruction degree (Fig. 9e and f). Recently, Kim and co-workers reported that the incorporation of Ru dopants into the NiFe2O4/NiMoO4 heterointerface can modulate the electronic configuration and induce the high-valence state of Ni3.6+δ, which promotes the surface reconstruction to a highly active phase of Ru-doped NiFeOOH/NiOOH. DFT calculations reveal that Ru doping can enrich electron density and optimize intermediate adsorption on the active Ni species, and thus enhance the OER activity.79 In addition, a strategy of electronic-ferry in metal element migration has been proposed to promote deep reconstruction of NiFe-based phosphide for the highly efficient and stable OER.142
![]() | ||
Fig. 9 (a) CV curves of CoFexAl2−xO4 (x = 0, 0.25 and 2) in 1 M KOH. The inset is the corresponding Tafel plots after oxide surface area normalization. (b) The schematic of reconstruction from spinel CoFe0.25Al1.75O4 into oxyhydroxide with an activated negatively charged oxygen ligand. (c) Schematic diagram of a surface reconstruction mechanism for CoFe0.25Al1.75O4. Green, blue, cyan, and red balls represent Al, Co, Fe, and O atoms, respectively. (d) Schematic of Al3+ leaching along with surface reconstruction of the spinel oxide. Reproduced with permission from ref. 137. Copyright 2019, Springer Nature. (e) Schematic of changes in the structural reconstruction degree and OER activity with increasing Fe substitution amount. Green, blue, brown, and red balls represent La, Ni, Fe, and O atoms, respectively. (f) Current densities at 1.6 V and the thickness of the newly formed reconstructed surface layer for LaNi1−xFexO3. Reproduced with permission from ref. 133. Copyright 2025, Springer Nature. (g) TEM images. (h) HAADF-STEM image and the corresponding EDS mapping of the surface region. Inset is the EDS line cut. (i) Schematic diagram of the in situ surface restructuring process of LiCoO1.8Cl0.2 and LiCoO2 during the OER. Yellow, blue, green, and red balls represent Li, Co, Cl, and O atoms, respectively. Reproduced with permission from ref. 147. Copyright 2021, Springer Nature. |
The incorporation of nonmetallic elements to optimize the restructuring and OER kinetics of precatalysts was also widely investigated. The reported anions to date predominantly involve those with large electronegativity, such as F,143–145 Cl,146–148 P,149,150 S,151–153 and so on. For example, Zhang et al. introduced the F anion into the LaNi0.75Fe0.25O3 (LNFO) perovskite by fluorination annealing and found that the incorporation of F can trigger a dynamic surface reconstruction to form an electrochemically active oxyhydroxide layer on the perovskite oxide, which reduces the energy barrier of the OER.143 Fan et al. revealed that S doping can improve the reconfiguration degree of NiFe LDH nanosheets and promote the phase transformation into highly active S-doped oxyhydroxides, optimizing the adsorption of reaction intermediates and enhancing the OER kinetics.151 Likewise, Su et al. prepared sulfur-doped NiCr LDH and studied the role of S doping in NiCr LDH for the OER.152 The results indicate that S incorporation not only facilitates the reconstruction of NiCr LDH by modulating Cr leaching, but also increases the covalency of the Ni–O bond and shifts the O 2p band center to a higher position, which in turn facilitates the oxidation of lattice oxygen.
Rationally manipulating the dynamic surface reconstruction and manipulating the in situ formed surface active species are highly desired for an efficient OER. Wang et al. selected layered LiCoO2 as a model material and prepared Cl-doped LiCoO2 (LiCoO2−xClx, x = 0, 0.1 or 0.2) by a solid-state reaction method. They demonstrated a cationic redox-tuning approach to engineer catalyst leaching and redirect the dynamic surface restructuring of LiCoO2−xClx under the OER.147 The TEM images display an amorphous layer on the surface of cycled LiCoO1.8Cl0.2, which is quite different from that of the cycled LiCoO2 (Fig. 9g). Further observation using STEM-EDS confirmed the presence of Co, O, and Cl in the restructured surface with no significant Cl depletion (Fig. 9h). These results together with in situ and ex situ XAS and ICP-MS analyses revealed that Cl doping triggered the in situ Co oxidation and Li extraction of LiCoO1.8Cl0.2 during the OER and the surface was transformed into self-terminated Cl-doped cobalt (oxy)hydroxide with a layered structure at a lower electrochemical potential (Fig. 9i). In contrast, Cl-free LiCoO2 required a higher potential and longer cycles to finish the surface restructuring into spinel-type Li1±xCo2O4, leading to an inferior OER performance.
In addition to anion vacancies, cation vacancies in precatalysts also play an important role in the reconstruction processes and the related OER activities. For instance, Wu et al. created cationic vacancy (Ni/Fe) defects in NiFe-LDH nanosheets through aprotic-solvent-solvation-induced leaking of metal cations. Observations from in situ Raman spectroscopy during the OER indicate that as the voltage increases, cation defects in NiFe-LDH promote a more facile local transformation of crystalline Ni(OH)x into a defective state, which finally forms the local NiOOH species.116 Recently, Li et al. have demonstrated that nickel vacancies in spinel NiFe2O4 lead to a more pronounced level of electrochemical surface reconstruction. Additionally, DFT calculations have shown that the cation-vacancy-induced effect can facilitate surface reconstruction by enhancing the covalency of the octahedral nickel–oxygen bonds in nickel ferrite. As a result, the in situ formation of amorphous metal oxyhydroxides on the surface provided more active sites, which accelerated the OER kinetics.158 Moreover, due to the unclear impact of structural reconstruction induced by different vacancy defects on the OER performance, Zhang et al. revealed the influence mechanism of defect types on the reconstruction process and the final active structure in the OER based on oxygen-deficient and metal-deficient Co3O4.157 It is found that cobalt oxides underwent a transformation into an amorphous [Co(OH)6] intermediate state, and then the mismatch rates of *OH adsorption and deprotonation resulted in irreversible catalyst reconstruction. The stronger *OH adsorption but weaker deprotonation induced by VO supplied the driving force for reconstruction, whereas the presence of VCo favored dehydrogenation and reduced the reconstruction rate. Although both oxygen and cobalt vacancies triggered highly active bridge Co sites in reconstructed catalysts, cobalt vacancies led to a shortening of the Co–Co distance to 3.38 Å under compressive lattice stress, which exhibited the best OER performance.
Recently, Chen et al. reported the influence of planar defects such as twinning and stacking faults on the structural reconstruction and intrinsic activity of electrocatalytic materials.161 They prepared an FeCoNi/FeAl2O4 hybrid coating on commercially pure titanium via the double cathode sputtering deposition technique, and then induced many nanotwins and stacking faults into the coating through ultrasonic cavitation. The presence of planar defects causes the adsorption energy of metal atoms with oxygen to shift towards a more negative value, enhancing oxygenophilicity and providing conditions for rapid surface reconstruction. As a result, the synergistic effect of different types of defects introduced by ultrasonic cavitation activates the surface reconstruction process, allowing FeCoNi/FeAl2O4-UC-6 min to undergo surface reconstruction at an applied potential of approximately 1.33 V (Fig. 10a) and form Al–(FeCoNi)OOH species with high OER activity. The Fourier transformed alternating current voltammetry (FTACV) curves indicate that the ultrasonically cavitated coating not only possesses a higher current density, but also initiates the surface reconstruction at a lower applied voltage (Fig. 10b), which is consistent with the in situ Raman data. The ICP-MS result in Fig. 10c shows that the leaching amounts of Fe, Co, and Ni are nearly identical throughout the OER process, which may be due to the involvement of all three elements in the surface reconstruction process. The HRTEM image of FeCoNi/FeAl2O4-UC-6 min after the test displays low crystallinity of the metal (oxy) hydroxide layer with a thickness exceeding 10 nm (Fig. 10d). In contrast to a typical reconstruction layer thickness of a few nanometers, FeCoNi/FeAl2O4-UC-6 min with various defect types undergoes a deeper surface reconstruction, and the thicker reconstruction layer enables the stability of the catalyst.
![]() | ||
Fig. 10 (a) In situ Raman spectra of FeCoNi/FeAl2O4-UC-6 min. (b) 5th-harmonic FTACV curves. (c) ICP-MS results for FeCoNi/FeAl2O4-UC-6 min after the chronopotentiometry test. (d) HRTEM image of FeCoNi/FeAl2O4-UC-6 min. Reproduced with permission from ref. 161. Copyright 2024, Wiley-VCH. (e) Schematic illustration of the self-reconstruction of the MOF heterojunction. Reproduced with permission from ref. 162. Copyright 2022, Wiley-VCH. |
Even though some advancements have been made in gaining insight into the structural reconstruction chemistry and designing superior electrocatalysts based on the structural reconstruction process, considerable challenges still remain: (1) The emerging external fields, such as magnetic fields, have been demonstrated to be effective methods to regulate the reconstruction process to enhance OER performance. However, the development of an external field and precatalyst types, as well as the mechanism of the effect of external fields on the reconstruction process, are still very limited, and need to be further extended and investigated. (2) Restoring the activity of degraded electrocatalysts under catalytic operating conditions is of great interest for increasing the lifetime of electrochemical devices, which can be achieved through a dynamic structural reconstruction process. There is a great need to develop new methods and recovery mechanisms to obtain ultra-long-life catalysts. (3) In most studies, the structural reconstruction occurs in low-concentration (0.1–1 M) KOH at room temperature, which is different from the industrial conditions (20–30 wt% KOH, at 50–80 °C). This will result in the unsuitability of the catalyst due to the different reconstruction processes. Hence, elucidating the fundamental mechanisms that govern structural reconstruction under practical conditions is both significant and imperative. (4) For some incompletely reconstructed precatalysts, the reconstructed species and the substrate may interact with each other, thus affecting the catalytic performance. This interaction is often overlooked but can play a crucial role in determining the overall catalytic behavior. Future research should focus on understanding and controlling these interactions to optimize the performance of electrocatalysts. (5) The tremendous progress of artificial intelligence (AI) and machine learning will undoubtedly drive the rapid development of materials science. These technologies in combination with theoretical calculations can be employed in precatalyst design and screening as well as the reconstruction process prediction to accelerate the discovery of new efficient electrocatalysts.
This journal is © The Royal Society of Chemistry 2025 |