Janus structure Ir–Ir3Ce@IrO2 nanocrystals as excellent bifunctionality catalysts for acidic overall water splitting

Yue Feng a, Ying Chang *a, Shaohong Guo a, Yaqiong Su *b, Jingchun Jia *a and Meilin Jia a
aCollege of Chemistry and Environmental Science, Inner Mongolia Key Laboratory of Green Catalysis and Inner Mongolia Collaborative Innovation Center for Water Environment Safety, Inner Mongolia Normal University, Key Laboratory of Infinite-Dimensional Hamiltonian System and Its Algorithm Application (Inner Mongolia Normal University), Ministry of Education Hohhot, 010022, China. E-mail: jjc1983@126.com; changying@imnu.edu.cn
bSchool of Chemistry, Engineering Research Center of Energy Storage Materials and Devices of Ministry of Education, National Innovation Platform (Center) for Industry-Education Integration of Energy Storage Technology, Xi'an Jiaotong University, Xi'an 710049, China. E-mail: yqsu1989@xjtu.edu.cn

Received 17th February 2025 , Accepted 21st August 2025

First published on 28th October 2025


Abstract

Enhancing the catalytic performance of iridium (Ir)-based catalysts for water electrolysis under acidic conditions presents a significant challenge. This difficulty primarily arises from the substantial energy barriers associated with the four-electron–proton coupled oxygen evolution reaction (OER) at the anode and the two-electron transfer hydrogen evolution reaction (HER) at the cathode. This study successfully synthesized a novel type of Ir–Ir3Ce@IrO2 nanoparticles supported on defective carbon materials (DCMs; Ir–Ir3Ce@IrO2/DCMs) using freeze-drying and conversion methods. Notably, the catalyst core features a unique Janus structure comprising metal and alloy components. This catalyst demonstrates exceptional acidic OER activity, overall water-splitting catalytic performance, and high stability. Experimental results indicate that the Ir–Ir3Ce@IrO2/DCMs electrocatalyst delivers an ultralow overpotential of 210 mV at 10 mA cm−2 for OER in 0.5 M H2SO4. Both structural characteristics and theoretical calculations suggest that Ir–Ir3Ce@IrO2/DCMs facilitate charge redistribution owing to various factors, including the alloying of precious metals and rare earth alloys, the Janus structure, and heterogeneous interfaces. The Ir–Ir3Ce@IrO2/DCMs || Ir–Ir3Ce@IrO2/DCMs electrolyzer can operate in acidic electrolytes for >40 h. This study presents a viable strategy to address the issues of instability and low efficiency associated with Ir-based OER electrocatalysts for acidic overall water splitting.


1. Introduction

Because of the environmental crisis caused by fossil fuel combustion, there is growing awareness among the public regarding this issue, and clean energy sources such as hydrogen are increasingly gaining interest.1,2 Hydrogen, recognized as a clean energy alternative, offers benefits such as affordability and a lack of pollution. Hydrogen production through the electrolysis of water is a promising technological solution, given the abundance of water resources on earth.3,4 Additionally, using wind or solar power to generate electricity allows water to be electrolyzed to produce hydrogen and store it, providing another method of storing solar and wind energy.5–7 However, the four-electron–proton coupling of the oxygen evolution reaction (OER) occurring at the anode presents significant energy barriers, severely limiting the kinetics of the electrochemical water-splitting process.8,9 Furthermore, the acidic environment can accelerate the corrosion of the electrode.10–12 There is an urgent need to develop bifunctional catalysts that can effectively improve the efficiency of the OER and hydrogen evolution reaction (HER) and optimize the electrochemical water-splitting process.

The noble metal iridium (Ir) and its oxide forms serve as catalysts in the electrochemical water-splitting process because of their excellent stability for OER.13 However, their performance in the OER still needs improvement compared with ruthenium-based catalysts, and their HER performance also needs enhancement compared with platinum-based catalysts. Moreover, the high cost of precious metals restricts the large-scale industrialization of catalysts.14–16 Alloying other metals with Ir has been commonly used in recent years to reduce catalyst costs and improve performance. The charge transfer between two atoms alters the electronic structure of Ir and the adsorption energy of reaction intermediates.17–19 The most widely used strategy is alloying 3d transition metals with precious metals, but dealloying often results in poor stability.20–24 However, lanthanide series in the rare earth elements are often overlooked.25 Among rare earth elements, cerium (Ce) is commonly used in catalysts for overall water splitting due to its unique electronic structure, and rare earth elements also have lanthanide contraction, which can control the lattice spacing of the alloy.26–28 However, the alloying of rare earth elements with precious metals for overall water splitting has rarely been reported. If Ir and Ce form an alloy, the 4f orbital of Ce will be located near the Fermi level, which reduces the energy barrier for d–d electron transfer and the 5d orbital of Ir, also close to the Fermi level, will rearrange electrons owing to d–d and d–f coupling. The adsorption strength of species on the catalyst surface is influenced by the energy level at the center of the d-band, thereby impacting the catalytic performance.18 The displacement of the d-band center optimizes the adsorption energy of intermediates on the catalyst, improves the stability of the alloy to a certain extent, and can exhibit electron-rich features to support faster electron transfer, thus protecting the Ir site in acidic media.25,29,30 However, because the trivalent Ce is easily oxidized to form oxides in air, there is an urgent need to develop an effective method to synthesize the alloy of Ir and Ce.31,32

Janus nanocrystals have asymmetric structures on both sides and were recently discovered as a new bifunctional material that can optimize functions beyond what traditional structures can achieve.33,34 Since the concept of “Janus” was proposed by Gennes in 1991,35 it has been applied in various electrocatalysis fields and has demonstrated excellent performance owing to its unique structure. The preparation of isotropic nanoparticles has been well developed, and further advancements in structure and properties are limited,36,37 prompting scientists to turn their attention to anisotropic nanoparticles. Janus nanoparticles possess two distinct regions that are independent yet interact with each other.38 The asymmetric structure of Ir–Ir3Ce forms the interface between Ir and Ir3Ce, significantly enhancing the electron transfer capability of the nanoparticles compared to Ir nanoparticles or Ir3Ce alloy nanoparticles. A substantial amount of electron transfer occurs at the interface, optimizing the chemisorption of precursor molecules and reaction intermediates, while the electron cloud density adjusts to reconstruct the active center and increase the active site density.39,40

As previously mentioned, alloying Ir and Ce can significantly reduce the dissolution of precious metals at oxidation potentials; however, the extent of this improvement remains limited. The formation of soluble Irn+ species readily occurs under oxidation potential,30 and the development of IrOx species on metal or alloy surfaces acts as a barrier that hinders the further oxidation and dissolution of these materials.41,42 The incomplete oxidation of Ir is crucial for preserving the remarkable electrical conductivity of the composite throughout the electrocatalytic process.41,43 The robust Ir–O bond considerably increases the theoretical dissolution potential of Ir during electrochemical oxidation, thereby enhancing the stability of the catalyst under acidic conditions.44 Furthermore, as the outer layer of Janus nanocrystals, IrO2 establishes two distinct interfaces with the nanocrystals: the Ir and IrO2 interface and the Ir3Ce and IrO2 interface. These two interfaces can significantly influence interfacial effects and optimize the adsorption energy of the reaction intermediates. Therefore, adding IrO2 can enhance the durability of the catalyst and improve its catalytic performance.45,46

Identifying new carriers with superior electrical conductivity and enhanced corrosion resistance compared with carbon powder presents a significant challenge in the development of catalysts for overall water splitting.47 The combination of a carbon atom with a nearby, more electronegative nitrogen atom alters the charge distribution.48,49 However, it is often overlooked that structural defects in carbon with an electron-conjugated structure can fulfill this requirement and may exhibit better electrocatalytic activity than heteroatom doping. Modifying the electronic structure can significantly enhance the adsorption energy associated with the OER and facilitate electron transfer.50–53 By directly calcining g-C3N4/C to introduce numerous defects as active sites, the loss of N atoms can adjust the electronic structure and enhance electron transfer in electrocatalysis.54–58 Regarding specific surface area, structures with porous features and highly exposed surfaces can reveal more active sites, thereby improving the catalytic performance of the catalysts.59 Consequently, defective carbon-based materials have emerged as the preferred choice for electrocatalyst carriers owing to their environmentally friendly properties and adjustable structures.

Inspired by the aforementioned perspectives, this study synthesized Ir–Ir3Ce Janus nanocrystalline nanoparticles loaded with IrO2 on defective carbon using freeze-drying and conversion methods, which were then applied to overall water splitting under acidic conditions. Compared with traditional techniques, this approach creates multiple interfaces that facilitate significant electron transfer. Furthermore, the alloying effect of Ce, which forms an alloy with Ir, enhances the inherent activity of the catalyst. As anticipated, the Ir–Ir3Ce@IrO2/DCMs demonstrate exceptional catalytic performance for overall water splitting in acidic electrolytes, requiring an overpotential of only 210 mV at 10 mA cm−2 for the OER. The distinctive core–shell architecture exhibits remarkable stability. Additionally, theoretical calculations indicate that the transfer of charge among Ir, Ir3Ce, and IrO2 can influence the adsorption characteristics of crucial intermediates in the reactions, thereby lowering the energy barrier for both the OER and HER, which contributes to the impressive performance of the catalyst in overall water splitting.

2. Results and discussion

2.1 Morphology and structure

The Ir–Ir3Ce@IrO2/DCMs nanocrystal was synthesized using freeze-drying and conversion methods (Fig. 1a). The synthesized Ir–Ir3Ce@IrO2/DCMs exhibited a porous coral morphology, as observed in scanning electron microscope images (Fig. S1), providing many exposed active sites on these coral-like structures for the electrocatalytic reaction. The structure of Ir–Ir3Ce@IrO2/DCMs was further investigated by X-ray diffraction (XRD) and transmission electron microscopy (TEM). The XRD results (Fig. 1b) showed that the prepared Ir–Ir3Ce@IrO2/DCMs corresponded well with the diffraction peaks of Ir and IrO2, while the diffraction peaks of Ir3Ce were not detected, which can be attributed to the low content of this alloy in the catalysts. Nevertheless, the diffraction peak of Ir in Ir–Ir3Ce@IrO2/DCMs is significantly shifted compared to that of Ir@IrO2/DCMs, which may be due to the formation of the alloy. The TEM images of Ir–Ir3Ce@IrO2/DCMs (Fig. 1c) reveal that the nanoparticles were uniformly loaded on the coral-like carriers, with no signs of agglomeration. The sizes of the nanoparticles are highlighted in yellow circles in Fig. 1d for statistical purposes, and the size of the synthesized nanoparticles is clearly shown in Fig. 1e to be 9.5 ± 1.5 nm, indicating a small size. The high-resolution transmission electron microscopy (HRTEM) image (Fig. 1f) of Ir–Ir3Ce@IrO2/DCMs shows continuous and distinct Ir3Ce lattice fringes at 0.115 and 0.111 nm. Their existence is further confirmed by the presence of two distinct Debye rings in the selected electron diffraction (SAED) diagram, corresponding to the Ir3Ce (101) and (012) crystal faces respectively, indicating that the Ir3Ce alloy exists in nanocrystals, and the continuous diffraction rings indicate that it has a polycrystalline structure. Fig. 1g shows the lattice fringes of the (111) crystal face of Ir and the (220) crystal face of IrO2, which clearly demonstrates that the structure of Ir–Ir3Ce@IrO2/DCMs consists of Janus structure nanocrystals formed by IrO2-coated Ir and Ir3Ce alloys loaded on the defective carbon-based materials.
image file: d5qi00479a-f1.tif
Fig. 1 (a) Schematic diagram of the synthesis method, (b) XRD patterns of Ir@IrO2/DCMs, Ir–Ir3Ce/DCMs and Ir–Ir3Ce@IrO2/DCMs, (c) and (d) TEM image of Ir–Ir3Ce@IrO2/DCMs, (e) particle size distribution of Ir–Ir3Ce@IrO2/DCMs, (f) and (g) HRTEM images of Ir–Ir3Ce@IrO2/DCMs, and (h) HRTEM and EDS element mapping of the as-prepared Ir–Ir3Ce@IrO2/DCMs.

Inductively coupled plasma spectrometry/mass spectrometry (ICP-AES/MS) (Table S1) indicates that Ir and Ce comprise only a tiny percentage of the catalyst, yet they outperform commercial Pt/C with a 20% loading. ICP-AES/MS (Table S2) of the centrifuged supernatant revealed that many Ir and Ce ions were present in the supernatant, which were not incorporated into the precursor interlayer during vacuum stirring and can be recovered and reused. Energy dispersive spectroscopy (EDS) mapping (Fig. 1h) demonstrates that Ir, Ce, C, and O elements are uniformly distributed in the clusters, with no accumulation observed. We also conducted TEM on Ir–Ir3Ce/DCMs without secondary calcination and on Ir@IrO2/DCMs without Ce doping (Fig. S2 and S3). The Ir–Ir3Ce Janus structure remains observable in the Ir–Ir3Ce/DCMs without secondary calcination. Similarly, in the Ir@IrO2/DCMs without Ce doping, the apparent core–shell structure of Ir and IrO2 can be seen. This suggests that IrO2 is formed during secondary calcination, which is an effective method for creating metal–metal oxide core–shell structures.

The introduction of glucose into the synthesis of g-C3N4/C increases the carbon source, contributing to the formation of a substantial amount of porous defective carbon-based materials after subsequent calcination. This porous defective carbon-based materials enhance the electron transfer capacity of metal nanoparticles and improves the accessibility of active sites. To investigate the pore size and pore type of Ir–Ir3Ce@IrO2/DCMs, we conducted Barrett–Emmett–Teller (BET) tests on g-C3N4/C and Ir–Ir3Ce@IrO2/DCMs (Fig. S4a). The N2 adsorption isotherm curves of Ir–Ir3Ce@IrO2/DCMs exhibited typical type III isotherm characteristics. We noted that the desorption curves of g-C3N4/C and Ir–Ir3Ce@IrO2/DCMs are nearly identical to the adsorption curves, with no hysteresis loops or inflection points,60,61 indicating predominantly well-connected open pores with a large pore volume. This leads to a significantly larger specific surface area of Ir–Ir3Ce@IrO2/DCMs compared to g-C3N4/C, thereby exposing more active sites to enhance catalytic activity. The pore size distribution diagram (Fig. S4b) indicates that both g-C3N4/C and Ir–Ir3Ce@IrO2/DCMs are primarily mesoporous. Ir–Ir3Ce@IrO2/DCMs exhibit a significantly higher pore count than g-C3N4/C, attributed to multiple high-temperature calcination processes; however, the pore size is considerably smaller than that of g-C3N4/C, owing to metal loading. It can be concluded that Ir–Ir3Ce@IrO2/DCMs are crucial for better exposure and utilization of active sites, thereby improving catalytic activity.

To characterize the chemical state and valence state of Ir–Ir3Ce@IrO2/DCMs, an X-ray photoelectron spectroscopy (XPS) was carried out. The full spectrum of Ir–Ir3Ce@IrO2/DCMs (Fig. S5), compared with Ir–Ir3Ce/DCMs without secondary calcination and Ir@IrO2/DCMs without Ce incorporation, can clearly show the coexistence of Ir, Ce, C and O elements, in which the peak of N element is found in Ir–Ir3Ce/DCMs. It may be the residual N element in g-C3N4/C. Ir 4f spectra at Ir–Ir3Ce@IrO2/DCMs (Fig. S6a) shows that Ir mainly has obvious peaks at 66.6 eV and 64.7 eV, and 63.8 eV and 61.7 eV. They correspond to Ir4+ 4f5/2, Ir0 4f5/2, Ir4+ 4f7/2 and Ir0 4f7/2 respectively, and the peak area of Ir0 is larger.20,62 The predominant form of Ir in Ir–Ir3Ce@IrO2/DCMs is still Ir0. Compared with Ir@IrO2/DCMs without Ce (Fig. 2a), we find that the 4f peak of Ir has a slight positive shift, indicating electron transfer between Ir and Ce. Compared with Ir–Ir3Ce/DCMs without secondary calcination (Fig. 2b), the 4f peak of Ir has a significant negative shift, indicating that electrons have been transferred from Ir–Ir3Ce Janus nanocrystals to IrO2. Janus nanocrystals and IrO2 also form electron transfer at the interface. In the XPS spectra of Ce 3d for Ir–Ir3Ce@IrO2/DCMs, one can identify eight prominent peaks corresponding to Ce 3d3/2 and Ce 3d5/2(Fig. S6b). Cerium exists in the oxidized state as Ce3+ and Ce4+, but more often as Ce3+ (Fig. S7a).63,64 Furthermore, O 1s spectra for Ir–Ir3Ce@IrO2/DCMs reveal three significant peaks: the metal–oxygen bond at 530.3 eV, oxygen vacancies at 532.5 eV, and the hydroxyl group at 533.8 eV (Fig. S6c). Compared with Ir–Ir3Ce/DCMs without secondary calcination (Fig. S7b), we found that the M–O bonds of the catalyst after secondary calcination increased, which may be due to the formation of the IrO2 shell (Fig. 2c). The C 1s spectra of Ir–Ir3Ce@IrO2/DCMs exhibit three peaks at 283, 284.8, and 285.8 eV, which belong to C[double bond, length as m-dash]C, C–C and C–O, respectively (Fig. S6d).65 XPS further confirmed the incorporation of Ce and the formation of IrO2 shells. Electron transfer existed at various interfaces formed, optimizing intermediates’ adsorption energy.66


image file: d5qi00479a-f2.tif
Fig. 2 (a) and (b) Fitted Ir 4f peaks of Ir@IrO2/DCMs, Ir–Ir3Ce@IrO2/DCMs and Ir–Ir3Ce/DCMs; (c) fitted O 1s peaks of Ir–Ir3Ce/DCMs and Ir–Ir3Ce@IrO2/DCMs, (d) Ir L3-edge XANES and (e) Ir L3-edge EXAFS spectra of Ir–Ir3Ce@IrO2/DCMs, IrO2 and Ir foil; (f) structural illustration of Ir–Ir3Ce@IrO2/DCMs, and (g) wavelet transform of Ir L3-edge EXAFS data of Ir foil, IrO2 and Ir–Ir3Ce@IrO2/DCMs.

The local electronic structure of Ir–Ir3Ce@IrO2/DCMs electrocatalysts was probed by X-ray absorption near-edge structure spectroscopy (XANES) and extended X-ray absorption fine structure spectroscopy (EXAFS) with Ir foil and IrO2 as reference samples. The line intensities at the Ir L3 edge were lower for Ir–Ir3Ce@IrO2/DCMs than for IrO2 and higher than for Ir foils (Fig. 2d), indicating that the valence of Ir in Ir–Ir3Ce@IrO2/DCMs is intermediate between that of Ir foils and IrO2. Subsequently, the coordination environments of Ir atoms were analyzed by EXAFS (Fig. S8), and the Ir–Ir3Ce@IrO2/DCMs showed three prominent peaks at ≈1.62, 2.64, and 3.50 Å from Ir–O bond, Ir–Ir bond, and Ir–Ce bond, respectively (Fig. 2e), with coordination numbers of 2.3, 7.2, and 4.3 for the three bonds, respectively (Table S3). Wavelet transform analysis (Fig. 2g) was next performed to gain more insight into the coordination information of the iridium species. Contour plots of Ir–Ir3Ce@IrO2/DCMs clearly show the Ir–O (k ≈ 5.5 Å) and Ir–Ir (k ≈ 11 Å) coordination of both intensities, similar to IrO2 (k ≈ 5.3 Å) as well as Ir foil (k ≈ 11.7 Å), with Ir–Ce (k ≈ 6.7 Å) coordination being relatively less pronounced, and further combined with the EXAFS data and fitting results, the presence of Ir–Ce coordination in the samples can be determined. These results suggest that Ir atoms in Ir–Ir3Ce@IrO2/DCMs coexist in IrO2, metallic Ir, and Ir3Ce alloy crystal structures (Fig. 2f).

To verify that the nanocrystals synthesized with a core–shell structure are loaded onto defective carbon-based materials, the XRD pattern of the g-C3N4/C precursor synthesized via freeze-drying is shown in Fig. S9a. The prepared g-C3N4/C exhibits a g-C3N4 structure with typical characteristic diffraction peaks at 13.1° and 27.1°.67–69 Compared with g-C3N4 synthesized via the direct calcination of dicyandiamide, the rapid dissolution and recrystallization of dicyandiamide result in more oligomers and smaller particle size, thus forming dicyandiamide nanocrystals favorable for the production of reticulate 3D g-C3N4.69–71 The Raman spectrum (Fig. S9b) shows characteristic D and G bands at 1350 and 1588 cm−1, respectively, where the D band corresponds to the disordered nature of the carbon material and the G band corresponds to its ordered nature. In addition to these two characteristic peaks, the Raman spectrum includes the D′′ band at 1496 cm−1 and the T band at 1200 cm−1, corresponding to the five-membered ring and the free state of carbon in the material, respectively. From the peak area fitting in Fig. S9b and S9c, it is evident that the proportion of the D′′ band and T band is smaller in the g-C3N4/C precursor after only 550 °C calcination compared with that of the carbon material after 550 °C calcination followed by 750 °C calcination and after 750 °C calcination followed by 400 °C air calcination. This indicates that the five-membered rings and free carbon states in the carbon material increase after several calcination cycles. The relative intensity ratio between the D band and G band (ID/IG) is often used to determine the degree of disorder in carbon materials. We found that carbon materials calcined at 750 °C and then air-calcined at 400 °C exhibited a significantly increased ID/IG ratio, suggesting that more carbon defects were created during the high-temperature calcination and subsequent air-calcination processes.72–74 To determine whether these carbon defects persist after metal loading, we conducted electron paramagnetic resonance spectroscopy (EPR) tests on g-C3N4/C and Ir–Ir3Ce@IrO2/DCMs (Fig. S9d). We observed that compared with g-C3N4/C, Ir–Ir3Ce@IrO2/DCMs clearly show a stronger peak at 2.002 G, indicating that metal loading introduces additional carbon defects into the material.75

2.2 Acidic OER and HER activity

The test used a standard three-electrode system in 0.5 M H2SO4.76 Analysis of polarization curves of the OER (Fig. 3a) shows that Ir–Ir3Ce@IrO2/DCMs exhibit better OER performance than commercial IrO2 and Pt/C. Additionally, the catalyst outperforms the same structure loaded on carbon black (Ir–Ir3Ce@IrO2/C; Fig. S10a). Specifically, the Ir–Ir3Ce@IrO2/DCMs electrocatalyst demonstrates a lower overpotential of 210 mV at 10 mA cm−2 compared with Ir@IrO2/DCMs and Ir–Ir3Ce/DCMs, which do not form the Ir3Ce alloy or a heterojunction with IrO2 (Fig. 3b). We observed that almost no OER performance was detected in Ce/DCMs, indicating that Ir and its oxide are the primary catalytic sites in the OER. The alloying of Ce and Ir plays a role in optimizing performance, but Ce itself possesses no OER catalytic properties. Ir–Ir3Ce@IrO2/DCMs also exhibit a lower Tafel slope (25.73 mV dec−1; Fig. 3c) and demonstrate a significant advantage in reaction kinetics.77 Moreover, Cdl values (Fig. 3d) were obtained from the cyclic voltammetry (CV) curves (Fig. S11). Ir–Ir3Ce@IrO2/DCMs have an electrochemically active surface area (EASA), indicating that the alloying of Ce and the formation of a heterojunction with IrO2 can enhance the activity of this system and increase the number of active sites.13,78 Electrochemical impedance spectroscopy (EIS) (Fig. 3e) reveals that Ir–Ir3Ce@IrO2/DCMs have a minimum polarization resistance of only 2.5 ohms, a rapid electron transfer rate, and high activity.79,80 When comparing Ir–Ir3Ce@IrO2/DCMs with Ir- and Ru-based catalysts reported in recent years (Fig. 3f and Table S4), the catalytic performance for OER is found to be exceptional. The OER performance of catalysts synthesized with different calcination temperatures and feed ratios was also tested (Fig. S12), and none of these catalysts exhibited inferior catalytic performance compared with Ir–Ir3Ce@IrO2/DCMs.
image file: d5qi00479a-f3.tif
Fig. 3 OER catalytic performances. (a) Polarization curves of Pt/C, Ir/C, Ir–Ir3Ce/DCMs, Ir–Ir3Ce@IrO2/DCMs, Ir@IrO2/DCMs, Ce/DCMs, IrO2/DCMs, and IrO2, (b) comparison of overpotential (10 mA cm−2), (c) corresponding Tafel plots; (d) ECSA plots, (e) EIS, (f) comparison of OER properties with other electrocatalysts. HER catalytic performances, (g) polarization curves, (h) ECSA plots, and (i) EIS.

We further evaluated long-term operation stability of Ir-Ir3Ce@IrO2/DCMs electrocatalyst in acidic mediium for OER. At a current density of 10 mA cm−2, after 72 h of continuous operation, there was only a minimal increase in the potential of Ir–Ir3Ce@IrO2/DCMs electrocatalysts compared to the initial measurements (Fig. S13a). On the other hand, Ir–Ir3Ce/DCMs were rapidly deactivated in a short period, and TEM and HRTEM (Fig. S13b and c) showed that the structure of Ir–Ir3Ce@IrO2/DCMs did not collapse before and after catalysis. The characteristic peaks can also be clearly observed in the XRD pattern and XPS spectra after the durability test (Fig. S14a). However, in order to ensure the accuracy of the test, the working electrode is directly tested in the XRD and XPS tests at this time, and the electrocatalyst is loaded on the carbon paper, so the carbon peaks in the XRD pattern are more obvious. These results indicate that IrO2 as a shell plays a significant role in preventing the dissolution of Ir and Ce during the catalytic process, ensuring the high structural stability of the acidic OER reaction.

For HER reaction in acidic media, Ir–Ir3Ce@IrO2/DCMs also show excellent catalytic performance, as shown in Fig. 3g. Ir–Ir3Ce@IrO2/DCMs and Ir–Ir3Ce/DCMs exhibit slightly better performance than Pt/C. At an overpotential of 74 mV, a current density of 100 mA cm−2 can be achieved. This performance is also superior to catalysts of the same structure supported on carbon black (Fig. S10b). Notably, after secondary air calcination to form precious metal oxides, a prominent hydrogen adsorption peak was observed at a potential of −0.07 to 0.1 V, with the current density fluctuating significantly due to the strong adsorption of *H on the catalyst surface. Therefore, we conclude that the oxide formed by secondary calcination, or the interface created between the oxide and the alloy or metal, contributes to the strong adsorption of *H on the catalyst surface. After overcoming the kinetic energy barrier to desorb *H, we observe that the polarization curve gradually stabilizes.81–83 However, we found that Ir–Ir3Ce@IrO2/DCMs and Ir–Ir3Ce/DCMs exhibited relatively similar catalytic activities, indicating that the active center for HER is Ir–Ir3Ce. As a shell oxide, OER optimization does not compromise the HER performance. As shown in Fig. 3h, the Cdl value obtained by Ir3Ce@IrO2/DCMs from the CV curves (Fig. S15) is also relatively higher than that of other catalysts, demonstrating the largest EASA. Simultaneously, the lower impedance (EIS) value (Fig. 3i) indicates its strongest electron transfer rate.78 The low Tafel slope (Fig. S16a) indicates that it has good reaction kinetics. Compared with some other Ir-based and Ru-based HER catalysts reported, the performance of Ir–Ir3Ce@IrO2/DCMs and Ir–Ir3Ce/DCMs is also more outstanding (Fig. S16b).

At a current density of 10 mA cm−2, in the 40 h stability test (Fig. S17a), high stability was nearly maintained, and the catalyst structure before and after the experiment (Fig. S17b and c) showed no significant differences. The HER performance of catalysts synthesized at various calcination temperatures and feed ratios was also tested in the experiment (Fig. S18), and the catalytic performance of these catalysts was lower than that of Ir–Ir3Ce@IrO2/DCMs and Ir–Ir3Ce/DCMs. This evidence indicates that Ir–Ir3Ce@IrO2/DCMs is a bifunctional catalyst with excellent OER performance and high HER activity.

2.3 Acidic overall water splitting application

As the Ir–Ir3Ce@IrO2/DCMs catalyst exhibits strong catalytic performance for both OER and HER reactions in an acidic environment, it is used as the catalyst for both the cathode and anode of the electrolyzer.84,85 The cell performance was observed to further evaluate the catalytic effect of the catalyst on overall water splitting (Fig. 4a). The Ir–Ir3Ce@IrO2/DCMs || Ir–Ir3Ce@IrO2/DCMs electrolyzer demonstrates better overall hydrolysis performance compared with the IrO2 || Pt/C electrolyzer, requiring only a low voltage of 1.42 V to achieve a current density of 10 mA cm−2 compared with 1.54 V for the IrO2 || Pt/C electrolyzer. Only 1.68 V is needed to reach a current density of 100 mA cm−2 (Fig. 4b). The Ir–Ir3Ce@IrO2/DCMs || Ir–Ir3Ce@IrO2/DCMs electrolyzer can operate continuously for >40 h, while the IrO2 || Pt/C electrolyzer shows significant performance degradation even in the initial phase. In terms of durability as a measure of catalytic performance, Ir–Ir3Ce@IrO2/DCMs also hold a clear advantage (Fig. 4c). These results undoubtedly reaffirm that the Ir–Ir3Ce@IrO2/DCM catalyst is an excellent bifunctional electrocatalyst for overall water splitting in a 0.5 M H2SO4 electrolyte.
image file: d5qi00479a-f4.tif
Fig. 4 Overall acidic water splitting of Ir–Ir3Ce@IrO2/DCMs || Ir–Ir3Ce@IrO2/DCMs and IrO2 || Pt/C electrolyzer in 0.5 M H2SO4. (a) Schematic diagram, (b) polarization plots, and (c) chronopotentiometric curves.

2.4 Study of the electrocatalytic mechanism

To gain a more intuitive understanding of the OER and HER electrocatalytic mechanisms of Ir–Ir3Ce@IrO2/DCMs, density functional theory (DFT) calculations and in situ characterization were employed. First, charge density difference analysis was performed on Ir–Ir3Ce@IrO2/DCMs (Fig. 5a) to reveal charge redistribution among Ir3Ce, Ir, and IrO2.86,87 The calculations indicated that 0.396 eV of electrons were transferred from the catalyst's nucleus (Janus nanocrystals) to the shell (IrO2), suggesting that charge redistribution between this metal oxide and the metal and alloy is likely to regulate the adsorption behavior of critical intermediates. Many investigations have reported that surface electron-deficient precious metal sites can serve as active centers for various electrocatalytic reactions.39,40,46,88–91 Therefore, charge redistribution can generate more active sites. In the process of water splitting, a standard measure of whether the electrocatalyst is suitable is its ability to reduce the energy barrier of the reaction, thus accelerating the reaction rate. For the OER mechanism, there are three standard models: the adsorption evolution mechanism (AEM), the lattice oxygen evolution mechanism (LOM), and the oxide path mechanism (OPM).75,92,93 Attenuated total reflection in situ surface-enhanced infrared absorption spectroscopy (ATR-SEIRAS) detected a gradually increasing signal peak at approximately 1090 cm−1 as the voltage increased (Fig. 5b). The AEM path was chosen for DFT calculations of the Ir–Ir3Ce@IrO2/DCMs electrocatalysts. As shown in Fig. 5c, the d-band center of Ir 5d in Ir–Ir3Ce@IrO2/DCMs shows a significant positive shift, reaching −1.098 eV. Compared with Ir–Ir3Ce/DCMs (−1.804 eV) and Ir@IrO2/DCMs (−2.015 eV) (the catalyst structure is shown in Fig. 5d), this shift indicates an increased adsorption strength for various OER intermediates, which may enhance OER kinetics.88,92,94 Consequently, we calculated the Gibbs free energy profiles of Ir–Ir3Ce@IrO2/DCMs, Ir–Ir3Ce/DCMs, Ir@IrO2/DCMs, Ir/C and IrO2 for OER (Fig. 5e) and constructed the adsorption processes of relevant OER intermediates on each catalyst (Fig. S19). The Ir–Ir3Ce@IrO2/DCMs exhibit an OER energy barrier of 2.38 eV, which is superior to 2.71 eV for 2.71 eV for Ir–Ir3Ce/DCMs, 2.57 eV for Ir@IrO2/DCMs, 2.89 eV for Ir/C, and 3.17 eV for IrO2, demonstrating the more favorable OER kinetics on the Ir–Ir3Ce@IrO2/DCMs, with an AEM path of the OER is shown in Fig. 5g. The conversion from *O to *OOH is the rate-limiting step catalyzed by Ir–Ir3Ce@IrO2/DCMs. The energy barrier of only 1.59 eV. This is lower than the 1.70 eV of Ir–Ir3Ce/DCMs, 1.88 eV of Ir@IrO2/DCMs, 2.09 eV of Ir/C, and 2.23 eV of IrO2(Fig. 5f), further confirming that Ir–Ir3Ce@IrO2/DCMs exhibit more favorable OER dynamics. In terms of HER, the process of HER intermediate H* adsorption on each catalyst is shown in Fig. S20, Ir–Ir3Ce@IrO2/DCMs has the lowest HER energy barrier (0.17 eV; Fig. 5h and i). These findings indicate that Ir–Ir3Ce@IrO2/DCMs have high theoretical activity for both OER and HER and can be applied to acidic overall water splitting.
image file: d5qi00479a-f5.tif
Fig. 5 (a) Differential charge density distribution and the electron transfer process, (b) ATR-SEIRAS measurements under various potentials for the Ir–Ir3Ce@IrO2/DCMs electrocatalyst during the OER process, (c) calculated PDOS plots of the Ir d-band for Ir–Ir3Ce/DCMs, Ir@IrO2/DCMs, and Ir–Ir3Ce@IrO2/DCMs with aligned Fermi levels, (d) structural illustration of Ir–Ir3Ce/DCMs and Ir@IrO2/DCMs, (e) free energy profiles of the OER process, (f) energy barrier of the rate-limiting step (conversion from *O to *OOH; U = 0 V), (g) schematic illustration of the adsorption evolution mechanism (AEM), (h) free energy profiles of the HER process, and (i) corresponding theoretical HER overpotentials.

3. Conclusion

In conclusion, we have successfully synthesized Ir–Ir3Ce@IrO2/DCM electrocatalysts loaded on defective carbon with Janus nanocrystals as cores, and they exhibit highly efficient and robust acidic water splitting performance. XRD, electron microscopy and XAFS have demonstrated the formation of alloys and a more complex structure that forms multiple interfaces to facilitate electron transfer. The electrocatalyst provides a low overpotential of 210 mV for OER at 10 mA cm−2 and outstanding stability under the protection of metal oxides. More importantly, its application in an acidic electrolyzer shows that only 1.42 V is required to reach a current density of 10 mA cm−2 as well as having a long-term stability of more than 40 hours, which outperforms the benchmark IrO2 || Pt/C. DFT calculations confirm a significant reduction in the conversion energy barrier from *O to *OOH, with lower OER and HER energy barrier compared to other catalysts, which improves the OER and HER catalytic activity of the Ir–Ir3Ce@IrO2/DCMs electrocatalysts. This work explores the effects of interfacial effects, alloying effects, and metal oxide shell layer on the overall water splitting performance, providing ideas for future design of more novel overall water splitting catalysts.

4. Experimental section

4.1. Electrocatalyst synthesis

Synthesis of g-C3N4/C composites. The g-C3N4/C composite was obtained by mixing 1 g of dicyandiamide, 10 g of glucose, and 300 mL of deionized water in a beaker at room temperature (≈20 °C) for 12 h, followed by freeze-drying for 24 h and calcining under Ar gas at a heating rate of 3 °C min−1 at 550 °C for 4 h.
Synthesis of Ir–Ir3Ce@IrO2/DCMs. 139 mg of IrCl3·H2O, 8 mg of Ce(NO)3·3H2O, 30 mL of glycol, and 50 mL of ethanol were placed into a beaker and stirred under vacuum for 8 h. Once dissolved, 400 mg of g-C3N4/C was added and subjected to vacuum for 30 min. The solids were collected via centrifugation to remove the Ir and Ce ions that were not incorporated into the precursor interlayer. The precipitate was then redissolved in a mixture of 70 ml of ethanol and 30 mL of glycol. At 180 °C under 300 rpm oil bath for 1 h and after cooling to room temperature, the mixture was washed with ethanol three times, vacuum-dried at 60 °C for 8 h, calcined at 750 °C at 3 °C min−1 heating rate under Ar gas for 4 h, resulting in Ir–Ir3Ce/DCMs. Further calcination at 400 °C at a heating rate of 3 °C min−1 under air for 4 h produced Ir–Ir3Ce@IrO2/DCMs. The comparison samples Ir@IrO2/DCMs and Ce/DCMs were prepared using the same method, except that Ce(NO)3·3H2O and IrCl3·H2O were not included. IrO2/DCMs were prepared in the same manner, but they were not calcined at 750 °C. Ir–Ir3Ce@IrO2/C is obtained by replacing the added g-C3N4/C with Vulcan XC-72R carbon black.

Conflicts of interest

The authors report no declarations of interest.

Data availability

Data for this article are available at mendeley data at https://data.mendeley.com/datasets/nyjbjdkx38/1.

Supplementary information: original data from SEM tests, TEM tests, XPS tests, XRD tests, BET tests, EPR tests, Raman tests, ATR-SEIRAS tests, XANES and EXAFS tests, electrochemical tests, as well as DFT theoretical calculations. See DOI: https://doi.org/10.1039/d5qi00479a.

Acknowledgements

We would like to thank the National Natural Science Foundation of China (Project No. 22062019, 22362028), the Key Laboratory of Infinite-dimensional Hamiltonian System and Its Algorithm Application (Inner Mongolia Normal University), the Ministry of Education (2023KFYB01), and the Natural Science Foundation of Inner Mongolia of China (2022QN02002). Supercomputing facilities were provided by Hefei Advanced Computing Center and Computing Center in Xi'an.

References

  1. D. Li, Y. Yu, W. Zhao, F. Wang, Y. Su and L. Cao, Rechargeable zinc-water battery for sustainable hydrogen generation, Nano Energy, 2024, 128, 109806 CrossRef CAS.
  2. J. Guo, J. D. Butson, Y. Zhang, G. Hu, X. Fan, G. K. Li, Y. Zhang, G. Hu, X. Fan and G. K. Li, Hydrogen generation from atmospheric water, J. Mater. Chem. A, 2024, 12, 12381–12396 RSC.
  3. H. Xie, Z. Zhao, T. Liu, Y. Wu, C. Lan, W. Jiang, L. Zhu, Y. Wang, D. Yang and Z. Shao, A membrane-based seawater electrolyser for hydrogen generation, Nature, 2022, 612, 673–678 CrossRef CAS PubMed.
  4. L. Miao, W. Jia, X. Cao and L. Jiao, Computational chemistry for water-splitting electrocatalysis, Chem. Soc. Rev., 2024, 53, 2771–2807 RSC.
  5. G. Singh, K. Ramadass, V. D. B. C. DasiReddy, X. Yuan, Y. Sik Ok, N. Bolan, X. Xiao, T. Ma, A. Karakoti, J. Yi and A. Vinu, Material-based generation, storage, and utilisation of hydrogen, Prog. Mater. Sci., 2023, 135, 101104 CrossRef CAS.
  6. J. J. T. Shibuya, D. E. Macphee and A. Cuesta, Putting stored hydrogen to work without consuming it: A flexible system for energy conversion and water desalination, J. Power Sources, 2024, 613, 234906 CrossRef CAS.
  7. S. Ghotia, T. Rimza, S. Singh, N. Dwivedi, A. K. Srivastava and P. Kumar, Hetero-atom doped graphene for marvellous hydrogen storage: unveiling recent advances and future pathways, J. Mater. Chem. A, 2024, 12, 12325–12357 RSC.
  8. W. Wang, C. Li, C. Zhou, X. Xiao, F. Li, N. Y. Huang, L. Li, M. Gu and Q. Xu, Enrooted–type metal–support interaction boosting oxygen evolution reaction in acidic media, Angew. Chem., Int. Ed., 2024, 63, e202406947 CrossRef CAS PubMed.
  9. W. Jia, X. Cao, X. Chen, H. Qin, L. Miao, Q. Wang and L. Jiao, γ–MnO2 as an electron reservoir for RuO2 oxygen evolution catalyst in acidic media, Small, 2024, 2310464 CrossRef CAS.
  10. J. Chen, Y. Ma, T. Huang, T. Jiang, S. Park, J. Xu, X. Wang, Q. Peng, S. Liu, G. Wang and W. Chen, Ruthenium–based binary alloy with oxide nanosheath for highly efficient and stable oxygen evolution reaction in acidic media, Adv. Mater., 2024, 36, 2312369 CrossRef CAS PubMed.
  11. Z. Zhang, C. Jia, P. Ma, C. Feng, J. Yang, J. Huang, J. Zheng, M. Zuo, M. Liu, S. Zhou and J. Zeng, Distance effect of single atoms on stability of cobalt oxide catalysts for acidic oxygen evolution, Nat. Commun., 2024, 15, 1767 CrossRef CAS.
  12. H. Liu, Z. Zhang, J. Fang, M. Li, M. G. Sendeku, X. Wang, H. Wu, Y. Li, J. Ge, Z. Zhuang, D. Zhou, Y. Kuang and X. Sun, Eliminating over-oxidation of ruthenium oxides by niobium for highly stable electrocatalytic oxygen evolution in acidic media, Joule, 2023, 7, 558–573 CrossRef CAS.
  13. L. Quan, H. Jiang, G. Mei, Y. Sun and B. You, Bifunctional electrocatalysts for overall and hybrid water splitting, Chem. Rev., 2024, 124, 3694–3812 CrossRef CAS PubMed.
  14. H. Sun, X. Xu, H. Kim, Z. Shao and W. Jung, Advanced electrocatalysts with unusual active sites for electrochemical water splitting, InfoMat, 2023, 6, e12494 CrossRef.
  15. L. Tian, Z. Li, X. Xu and C. Zhang, Advances in noble metal (Ru, Rh, and Ir) doping for boosting water splitting electrocatalysis, Mater. Chem. A, 2021, 9, 13459–13470 RSC.
  16. B. Lu, C. Wahl, R. dos Reis, J. Edgington, X. K. Lu, R. Li, M. E. Sweers, B. Ruggiero, G. T. K. K. Gunasooriya, V. Dravid and L. C. Seitz, Key role of paracrystalline motifs on iridium oxide surfaces for acidic water oxidation, Nat. Catal., 2024, 7, 868–877 CrossRef CAS.
  17. F. Lv, W. Zhang, W. Yang, J. Feng, K. Wang, J. Zhou, P. Zhou and S. Guo, Ir–based alloy nanoflowers with optimized hydrogen binding energy as bifunctional electrocatalysts for overall water splitting, Small Methods, 2019, 4, 1900129 CrossRef.
  18. J. Song, C. He, C. Ma, J. Xia, F. Feng, X. Ma, S. Han, H. Zhang, Y. Yang, B. Li, Q. Lu, W. Cao and L. Zhu, Atomically ordered Ir3Ti intermetallics for pH-universal overall water splitting, Mater. Chem. A, 2024, 12, 18167–18174 RSC.
  19. Y. Xie, Y. Feng, S. Pan, H. Bao, Y. Yu, F. Luo and Z. Yang, Electrochemical leaching of Ni dopants in IrRu alloy electrocatalyst boosts overall water splitting, Adv. Funct. Mater., 2024, 2406351 CrossRef CAS.
  20. Z. Zhang, S. Liu, Y. Zhou, J. Li, L. Xu, J. Yang, H. Pang, M. Zhang and Y. Tang, Engineering ultrafine PtIr alloy nanoparticles into porous nanobowls via a reactive template-engaged assembly strategy for high-performance electrocatalytic hydrogen production, Mater. Chem. A, 2024, 12, 10148–10156 RSC.
  21. X. Wang, Z. Qin, J. Qian, L. Chen and K. Shen, IrCo nanoparticles encapsulated with carbon nanotubes for efficient and stable acidic water splitting, ACS Catal., 2023, 13, 10672–10682 CrossRef CAS.
  22. D. Cao, J. Wang, H. Zhang, H. Xu and D. Cheng, Growth of IrCu nanoislands with rich IrCu/Ir interfaces enables highly efficient overall water splitting in non-acidic electrolytes, Chem. Eng. J., 2021, 416, 129128 CrossRef CAS.
  23. W. H. Lee, J. Yi, H. N. Nong, P. Strasser, K. H. Chae, B. K. Min, Y. J. Hwang and H.-S. Oh, Electroactivation-induced IrNi nanoparticles under different pH conditions for neutral water oxidation, Nanoscale, 2020, 12, 14903–14910 RSC.
  24. H. Mou, Q. Chang, Z. Xie, S. Hwang, S. Kattel and J. G. Chen, Enhancing glycerol electrooxidation from synergistic interactions of platinum and transition metal carbides, Appl. Catal., B, 2022, 316, 121648 CrossRef CAS.
  25. Y. Jiang, Z. Liang, J.-C. Liu, H. Fu, C.-H. Yan and Y. Du, Stimulating electron delocalization of lanthanide elements through high-entropy confinement to promote electrocatalytic water splitting, ACS Nano, 2024, 18, 19137–19149 CrossRef CAS PubMed.
  26. Y. Cheng, L. Zhu and Y. Gong, Ce doped Ni(OH)2/Ni-MOF nanosheets as an efficient oxygen evolution and urea oxidation reactions electrocatalyst, Int. J. Hydrogen Energy, 2024, 58, 416–425 CrossRef CAS.
  27. W. Zhao, F. Xu, J. Yang, X. Hu and B. Weng, Ce single-atom incorporation enhances the oxygen evolution reaction of Co3O4 in acid, Inorg. Chem., 2024, 63, 1947–1953 CrossRef CAS.
  28. Z. Wan, K. Yang, P. Li, S. Yang, X. Wang, R. Gao, X. Xie, G. Deng, M. Yang and Z. Wang, Atomically dispersed Au single sites and nanoengineered structural defects enable a high electrocatalytic activity and durability for hydrogen evolution reaction and overall urea electrolysis, Electrochim. Acta, 2024, 499, 144685 CrossRef CAS.
  29. S. Zhu, L. Yang, J. Bai, Y. Chu, J. Liu, Z. Jin, C. Liu, J. Ge and W. Xing, Ultra-stable Pt5La intermetallic compound towards highly efficient oxygen reduction reaction, Nano Res., 2022, 16, 2035–2040 CrossRef.
  30. S. Zhang, M. Sun, L. Yin, S. Wang, B. Huang, Y. Du and C.-H. Yan, Tailoring the electronic structure of Ir alloy electrocatalysts through lanthanide (La, Ce, Pr, and Nd) for acidic oxygen evolution enhancement, Adv. Energy Sustainability Res., 2023, 4, 2300023 CrossRef CAS.
  31. J. Li, Z. Li, Z. Zheng, X. Zhang, H. Zhang, H. Wei and H. Chu, Tuning the product selectivity toward the high yield of glyceric acid in Pt−CeO2/CNT electrocatalyzed oxidation of glycerol, ChemCatChem, 2022, 14, e202200509 CrossRef CAS.
  32. X. Zhang, D. Zhou, X. Wang, J. Zhou, J. Li, M. Zhang, Y. Shen, H. Chu and Y. Qu, Overcoming the deactivation of Pt/CNT by introducing CeO2 for selective base-free glycerol-to-glyceric acid oxidation, ACS Catal., 2020, 10, 3832–3837 CrossRef CAS.
  33. X. Chen, J. Pu, X. Hu, Y. Yao, Y. Dou, J. Jiang and W. Zhang, Janus hollow nanofiber with bifunctional oxygen electrocatalyst for rechargeable Zn–Air Battery, Small, 2022, 18, 2200578 CrossRef CAS.
  34. M. Lattuada and T. A. Hatton, properties and applications of Janus nanoparticles, Nano Today, 2011, 6, 286–308 CrossRef CAS.
  35. P. G. Gennes, Soft matter, Rev. Mod. Phys., 1992, 64, 645–648 CrossRef.
  36. S. K. T. Aziz, M. Ummekar, I. Karajagi, S. K. Riyajuddin, K. V. R. Siddhartha, A. Saini, A. Potbhare, R. G. Chaudhary, V. Vishal, P. C. Ghosh and A. Dutta, A Janus cerium-doped bismuth oxide electrocatalyst for complete water splitting, Cell Rep. Phys. Sci., 2022, 3, 101106 CrossRef CAS.
  37. H. Lou, K. Qiu and G. Yang, Janus Mo2P3 Monolayer as an electrocatalyst for hydrogen evolution, ACS Appl. Mater. Interfaces, 2021, 13, 57422–57429 CrossRef CAS PubMed.
  38. B. Tang, Y. Zhou, Q. Ji, Z. Zhuang, L. Zhang, C. Wang, H. Hu, H. Wang, B. Mei, F. Song, S. Yang, B. M. Weckhuysen, H. Tan, D. Wang and W. Yan, A Janus dual-atom catalyst for electrocatalytic oxygen reduction and evolution, Nat. Synth., 2024, 3, 878–890 CrossRef CAS.
  39. S. W. Yu, S. Kwon, Y. Chen, Z. Xie, X. Lu, K. He, S. Hwang and J. G. Chen, Construction of a Pt–CeOx interface for the electrocatalytic hydrogen evolution reaction, Adv. Funct. Mater., 2024, 34, 2402966 CrossRef CAS.
  40. X. Mu, Y. Zhu, X. Gu, S. Dai, Q. Mao, L. Bao, W. Li, S. Liu, J. Bao and S. Mu, Awakening the oxygen evolution activity of MoS2 by oxophilic-metal induced surface reorganization engineering, J. Energy Chem., 2021, 62, 546–551 CrossRef CAS.
  41. M. M. Hosseini, M. G. Hosseini, I. Ahadzadeh and R. Najjar, IrO2-ZrO2-SiO2 ternary oxide composites- based DSAs: Activity toward oxygen evolution reaction with long-term stability, J. Taiwan Inst. Chem. Eng., 2024, 161, 105548 CrossRef CAS.
  42. M. Li, J. Qi, H. Zeng, J. Chen, Z. Liu, L. Gu, J. Wang, Y. Zhang, M. Wang, Y. Zhang, X. Lu and C. Yang, Structural impacts on the degradation behaviors of Ir-based electrocatalysts during water oxidation in acid, J. Colloid Interface Sci., 2024, 674, 108–117 CrossRef CAS PubMed.
  43. Y. Shan, J. Pan, Y. Song, Y. Zhu and T. Li, Engineering the orbital splitting and electronic reoccupation at alkali-metal-doped perovskites with face-shared IrO6 octahedral dimers, Chem. Phys., 2024, 582, 112292 CrossRef CAS.
  44. S. Zhao, S.-F. Hung, L. Deng, W.-J. Zeng, T. Xiao, S. Li, C.-H. Kuo, H.-Y. Chen, F. Hu and S. Peng, Constructing regulable supports via non-stoichiometric engineering to stabilize ruthenium nanoparticles for enhanced pH-universal water splitting, Nat. Commun., 2024, 15, 2728 CrossRef CAS.
  45. J. Zhang, Q. Zhang and X. Feng, Support and interface effects in water–splitting electrocatalysts, Adv. Mater., 2019, 31, 1808167 CrossRef PubMed.
  46. Z. Lu, J. Wang, P. Zhang, W. Guo, Y. Shen, P. Liu, J. Ji, H. Du, M. Zhao, H. Liang and J. Guo, Directed electron transport induced surface reconstruction of 2D NiFe-LDH/Stanene heterojunction catalysts for efficient oxygen evolution, Appl. Catal. B Environ. Energy, 2024, 353, 124073 CrossRef CAS.
  47. C. Z. Yuan, H. Zhao, S. Huang, J. Li, L. Zhang, W. Zhao, Y. Weng, X. Zhang, S. Ye and Y. Chen, Designing and regulating catalysts for enhanced oxygen evolution in acid electrolytes, Inorg. Chem., 2023, 2, 467–493 CAS.
  48. Y. Chen, T. Jiang, C. Tian, Y. Zhan, A. Kempf, L. Molina-Luna, J. P. Hofmann, R. Riedel and Z. Yu, Single–source–precursor–derived binary FeNi phosphide nanoparticles encapsulated in N, P Co–doped carbon as electrocatalyst for hydrogen evolution reaction and oxygen evolution reaction, Energy Technol., 2023, 11, 2300233 CrossRef CAS.
  49. Y. Xu, J. Wang, S. Wang, Z. Zhai, B. Ren, Z. Tian, L. Zhang and Z. Liu, Lattice defective and N, S co-doped carbon nanotubes for trifunctional electrocatalyst application, Int. J. Hydrogen Energy, 2023, 48, 34340–34354 CrossRef CAS.
  50. L. Tao, Q. Wang, S. Dou, Z. Ma, J. Huo, S. Wang and L. Dai, Edge-rich and dopant-free graphene as a highly efficient metal-free electrocatalyst for the oxygen reduction reaction, Chem. Comm., 2016, 52, 2764–2767 RSC.
  51. D. Bi, N. Shen, Z. Tang, Z. Yang, S. A. Grigoriev, P. He, Q. Lai and Y. Liang, Defective carbon derived using a dissolution–recrystallization strategy for oxygen reduction electrocatalysis, ACS Appl. Mater. Interfaces, 2023, 15, 30179–30186 CrossRef CAS.
  52. X. Yan, Y. Jia, J. Chen, Z. Zhu and X. Yao, Defective–activated–carbon–supported Mn–Co nanoparticles as a highly efficient electrocatalyst for oxygen reduction, Adv. Mater., 2016, 28, 8771–8778 CrossRef CAS PubMed.
  53. A. Kostuch, E. Negro, G. Pagot, S. Zoladek, I. A. Rutkowska, O. Siamuk, A. Chmielnicka, V. Di Noto and P. J. Kulesza, Importance of defective structure of ceria additive on selectivity and stability of carbon-supported low-Pt-content-catalysts during oxygen reduction, Catal. Today, 2024, 441, 114889 CrossRef CAS.
  54. W. Wang, L. Shang, G. Chang, C. Yan, R. Shi, Y. Zhao, G. I. N. Waterhouse, D. Yang and T. Zhang, Intrinsic carbon–defect–driven electrocatalytic reduction of carbon dioxide, Adv. Mater., 2019, 31, 1808276 CrossRef.
  55. Y. Wang, S. Zhao, Y. Zhang, J. Fang, Y. Zhou, S. Yuan, C. Zhang and W. Chen, One-pot synthesis of K-doped g-C3N4 nanosheets with enhanced photocatalytic hydrogen production under visible-light irradiation, Appl. Surf. Sci., 2018, 440, 258–265 CrossRef CAS.
  56. L. Yang, X. Liu, Z. Liu, C. Wang, G. Liu, Q. Li and X. Feng, Enhanced photocatalytic activity of g-C3N4 2D nanosheets through thermal exfoliation using dicyandiamide as precursor, Ceram. Int., 2018, 44, 20613–20619 CrossRef CAS.
  57. Q. Wu, J. Gao, J. Feng, Q. Liu, Y. Zhou, S. Zhang, M. Nie, Y. Liu, J. Zhao, F. Liu, J. Zhong and Z. Kang, A CO2 adsorption dominated carbon defect-based electrocatalyst for efficient carbon dioxide reduction, J. Mater. Chem. A, 2020, 8, 1205–1211 RSC.
  58. R. Kumar, S. Kumar, S. G. Chandrappa, N. Goyal, A. Yadav, N. Ravishankar, A. S. Prakash and B. Sahoo, Nitrogen-doped carbon nanostructures embedded with Fe-Co-Cr alloy based nanoparticles as robust electrocatalysts for Zn-air batteries, J. Alloys Compd., 2024, 984, 173862 CrossRef CAS.
  59. R. Daiyan, X. Tan, R. Chen, W. H. Saputera, H. A. Tahini, E. Lovell, Y. H. Ng, S. C. Smith, L. Dai, X. Lu and R. Amal, Electroreduction of CO2 to CO on a mesoporous carbon catalyst with progressively removed nitrogen moieties, ACS Energy Lett., 2018, 3, 2292–2298 CrossRef CAS.
  60. Z. Li, D. Liu, Y. Cai, Y. Wang and J. Tengi, Tunable subradiant mode in free-standing metallic nanohole arrays for high-performance plasmofluidic sensing, Fuel, 2019, 257, 116031 CrossRef CAS.
  61. Y. Zhu, S. Zhao, K. Deng, P. Wu, K. Feng and L. Li, Integrated mRNA and small RNA sequencing reveals a microRNA regulatory network associated with starch biosynthesis in lotus (nelumbo nucifera gaertn.) rhizomes, Int. J. Mol. Sci., 2022, 23, 7605 CrossRef CAS.
  62. H. Hu, F. M. D. Kazim, Z. Ye, Y. Xie, Q. Zhang, K. Qu, J. Xu, W. Cai, S. Xiao and Z. Yang, Electronically delocalized Ir enables efficient and stable acidic water splitting, J. Mater. Chem. A, 2020, 8, 20168–20174 RSC.
  63. D. Rathore, S. Ghosh, A. Gupta, J. Chowdhury and S. Pande, Ce-doped NiSe nanosheets on carbon cloth for electrochemical water-splitting, ACS Appl. Nano Mater., 2024, 7, 9730–9744 CrossRef CAS.
  64. Y. Liao, X. Lu, X. Jin, H. Chen, X. Huang, Y. Li and J. Li, Doping CeO2/Ce(OH)CO3/CF nanohybrids with Gd for structural tuning and oxygen evolution reaction performance enhancing, Int. J. Hydrogen Energy, 2024, 72, 288–296 CrossRef CAS.
  65. J. Chen, P. Cui, G. Zhao, K. Rui, M. Lao, Y. Chen, X. Zheng, Y. Jiang, H. Pan, S. X. Dou and W. Sun, Low–coordinate iridium oxide confined on graphitic carbon nitride for highly efficient oxygen evolution, Angew. Chem., Int. Ed., 2019, 58, 12540–12544 CrossRef CAS.
  66. R. Qin, G. Chen, X. Feng, J. Weng and Y. Han, Ru/Ir–based electrocatalysts for oxygen evolution reaction in acidic conditions: from mechanisms, optimizations to challenges, Adv. Sci., 2024, 11, 2309364 CrossRef CAS PubMed.
  67. S. Guo, Z. Deng, M. Li, B. Jiang, C. Tian, Q. Pan and H. Fu, Phosphorus-doped carbon nitride tubes with a layered micro-nanostructure for enhanced visible-light photocatalytic hydrogen evolution, Angew. Chem., Int. Ed., 2016, 55, 1830–1834 CrossRef CAS.
  68. J. Chen, Y. Xiao, N. Wang, X. Kang, D. Wang, C. Wang, J. Liu, Y. Jiang and H. Fu, Facile synthesis of a Z-scheme CeO2/C3N4 heterojunction with enhanced charge transfer for CO2 photoreduction, Sci. China Mater., 2023, 66, 3165–3175 CrossRef CAS.
  69. C. Zhang, J. Wang, Y. Liu, W. Li, Y. Wang, G. Qin and Z. Lv, Electrocatalytic HER enhancement of C3N4 in NiCo2O4, Chem. - Asian J., 2022, 17, e202200377 CrossRef CAS PubMed.
  70. J. Tian, Z. Chen, J. Jing, C. Feng, M. Sun and W. Li, Enhanced photocatalytic performance of the MoS2/g-C3N4 heterojunction composite prepared by vacuum freeze drying method, J. Photochem. Photobiol., A, 2020, 390, 112260 CrossRef CAS.
  71. C. Hu, W.-L. Chiu, C.-Y. Wang and V.-H. Nguyen, Freeze-dried dicyandiamide-derived g-C3N4 as an effective photocatalyst for H2 generation, J. Taiwan Inst. Chem. Eng., 2021, 129, 128–134 CrossRef CAS.
  72. J. E. Weis, J. Vejpravova, T. Verhagen, Z. Melnikova, S. Costa and M. Kalbac, Surface–enhanced Raman spectra on graphene, J. Raman Spectrosc., 2017, 49, 168–173 CrossRef.
  73. L. Li, J. Liu, X. Liu and W. Wu, Preparation of mesocarbon microbeads by thermal polymerization of the blended coal tar and biomass tar pitch as anode material for sodium-ion batteries, Int. J. Electrochem. Sci., 2022, 17, 221150 CrossRef CAS.
  74. Q. Lai, J. Zheng, Z. Tang, D. Bi, J. Zhao and Y. Liang, Optimal configuration of N-Doped carbon defects in 2D turbostratic carbon nanomesh for advanced oxygen reduction electrocatalysis, Angew. Chem., Int. Ed., 2020, 59, 11999–12006 CrossRef CAS PubMed.
  75. Q. Wu, Y. Jia, Q. Liu, X. Mao, Q. Guo, X. Yan, J. Zhao, F. Liu, A. Du and X. Yao, Ultra-dense carbon defects as highly active sites for oxygen reduction catalysis, Chem, 2022, 8, 2715–2733 CAS.
  76. Y. Wang, X. Ge, Q. Lu, W. Bai, C. Ye, Z. Shao and Y. Bu, Accelerated deprotonation with a hydroxy-silicon alkali solid for rechargeable zinc-air batteries, Nat. Commun., 2023, 14, 6968 CrossRef CAS PubMed.
  77. O. van der Heijden, S. Park, R. E. Vos, J. J. J. Eggebeen and M. T. M. Koper, Tafel slope plot as a tool to analyze electrocatalytic reactions, ACS Energy Lett., 2024, 9, 1871–1879 CrossRef CAS PubMed.
  78. G. Gao, Z. Sun, X. Chen, G. Zhu, B. Sun, Y. Yamauchi and S. Liu, Recent advances in Ru/Ir-based electrocatalysts for acidic oxygen evolution reaction, Appl. Catal. B Environ. Energy, 2024, 343, 123584 CrossRef CAS.
  79. S. Wang, J. Zhang, O. Gharbi, V. Vivier, M. Gao and M. E. Orazem, Electrochemical impedance spectroscopy, Nat. Rev. Methods Primers, 2021, 1, 41 CrossRef CAS.
  80. F. Ciucci, Modeling electrochemical impedance spectroscopy, Curr. Opin. Electrochem., 2019, 13, 132–139 CrossRef CAS.
  81. Z. Li, K. Werner, K. Qian, R. You, A. Płucienik, A. Jia, L. Wu, L. Zhang, H. Pan, H. Kuhlenbeck, S. Shaikhutdinov, W. Huang and H. J. Freund, Oxidation of reduced ceria by incorporation of hydrogen, Angew. Chem., Int. Ed., 2019, 58, 14686–14693 CrossRef CAS PubMed.
  82. C. Copéret, D. P. Estes, K. Larmier and K. Searles, Isolated surface hydrides: formation, structure, and reactivity, Chem. Rev., 2016, 116, 8463–8505 CrossRef PubMed.
  83. K. Deng, Z. Lian, W. Wang, J. Yu, Q. Mao, H. Yu, Z. Wang, L. Wang and H. Wang, Hydrogen spillover effect tuning the rate-determining step of hydrogen evolution over Pd/Ir hetero-metallene for industry-level current density, Appl. Catal. B Environ. Energy, 2024, 352, 124047 CrossRef CAS.
  84. A. Zhou, D. Song, Z. Hui, J. Yang, W. An, X. Zuo, X. Ye and M. Jiang, Coupling NiFe-MOF with antiperovskite nickel-based nitrides as efficient electrocatalysts for overall water splitting, Results Phys., 2024, 59, 107611 CrossRef.
  85. X. Qian, L. Jiang, J. Fang, J. Ye, G. He and H. Chen, Constructing a self-supported bifunctional multiphase heterostructure for electrocatalytic overall water splitting, Inorg. Chem., 2024, 4c01963 Search PubMed.
  86. A. Esmaeili, F. Keivanimehr, M. Mokhtarian, S. Habibzadeh, O. Abida and M. Moghaddamian, 2D Ni2P/N-doped graphene heterostructure as a Novel electrocatalyst for hydrogen evolution reaction: A computational study, Heliyon, 2024, 10, e27133 CrossRef CAS PubMed.
  87. F. Yang, X. Zhu, B. Jia, J. Hao, Y. Ma, P. Lu and D. Li, Trifunctional electrocatalysts of single transition metal atom doped g-C4N3 with high performance for OER, ORR and HER, Int. J. Hydrogen Energy, 2024, 51, 195–206 CrossRef CAS.
  88. Y. Jiang, J. Leng, S. Zhang, T. Zhou, M. Liu, S. Liu, Y. Gao, J. Zhao, L. Yang, L. Li and W. Zhao, Modulating water splitting kinetics via charge transfer and interfacial hydrogen spillover effect for robust hydrogen evolution catalysis in alkaline media, Adv. Sci., 2023, 10, 2302358 CrossRef CAS.
  89. X. Zheng, L. Li, M. Deng, J. Li, W. Ding, Y. Nie and Z. Wei, Understanding the effect of interfacial interaction on metal/metal oxide electrocatalysts for hydrogen evolution and hydrogen oxidation reactions on the basis of first-principles calculations, Catal. Sci. Technol., 2020, 10, 4743–4751 RSC.
  90. K. K. Mandari and M. Kang, Ni2P/Co2P as a highly efficient catalyst for enhancing the electrochemical performance of P-NiMoO4 for oxygen evolution reaction, J. Alloys Compd., 2024, 976, 172973 CrossRef CAS.
  91. P. Li, S. Luo, Z. Xiong, H. Xiao, X. Wang, K. Peng, X. Xie, Z. Zhang, G. Deng, M. Yang and C. Wang, RuxMoS2 interfacial heterojunctions achieve efficient overall water splitting and stability in both alkaline and acidic media under large current density exceeding 100 mA cm−2, Mol. Catal., 2025, 570, 114710 CAS.
  92. Z. Dong, C. Zhou, W. Chen, F. Lin, H. Luo, Z. Sun, Q. Huang, R. Zeng, Y. Tan, Z. Xiao, H. Huang, K. Wang, M. Luo, F. Lv and S. Guo, Cerium dioxide as an electron buffer to stabilize iridium for efficient water electrolysis, Adv. Funct. Mater., 2024, 2302358 Search PubMed.
  93. R. Ram, L. Xia, H. Benzidi, A. Guha, V. Golovanova, A. G. Manjón, D. L. Rauret, P. S. Berman, M. Dimitropoulos, B. Mundet, E. Pastor, V. Celorrio, C. A. Mesa, A. M. Das, A. Pinilla-Sánchez, S. Giménez, J. Arbiol, N. López and F. P. G. D. Arquer, Water-hydroxide trapping in cobalt tungstate for proton exchange membrane water electrolysis, Science, 2024, 384, 1373–1380 CrossRef CAS PubMed.
  94. J. Zhu, Y. Guo, F. Liu, H. Xu, L. Gong, W. Shi, D. Chen, P. Wang, Y. Yang, C. Zhang, J. Wu, J. Luo and S. Mu, Regulative electronic states around ruthenium/ruthenium disulphide heterointerfaces for efficient water splitting in acidic media, Adv. Funct. Mater., 2021, 60, 12328–12334 CAS.

This journal is © the Partner Organisations 2026
Click here to see how this site uses Cookies. View our privacy policy here.