Synergistic interlayer confinement and built-in electric field construct reconstruction-inhibited cobalt selenide for robust oxygen evolution at high current density

Yuanhua Xiao Jinhui Shou Shiwei Zhang Ya Shen Junwei Liu Dangcheng Su Yang Kong Xiaodong Jia Qingxiang Yang Shaoming Fang Xuezhao Wang

Citation:  Yuanhua Xiao, Jinhui Shou, Shiwei Zhang, Ya Shen, Junwei Liu, Dangcheng Su, Yang Kong, Xiaodong Jia, Qingxiang Yang, Shaoming Fang, Xuezhao Wang. Synergistic interlayer confinement and built-in electric field construct reconstruction-inhibited cobalt selenide for robust oxygen evolution at high current density[J]. Chinese Chemical Letters, 2025, 36(11): 111441. doi: 10.1016/j.cclet.2025.111441 shu

Synergistic interlayer confinement and built-in electric field construct reconstruction-inhibited cobalt selenide for robust oxygen evolution at high current density

English

  • Electrocatalytic water splitting has garnered significant interest as a sustainable technology to produce "green" hydrogen without carbon emissions [1,2]. This process comprises two half-reactions on the electrodes which are the hydrogen evolution reaction (HER) and the oxygen evolution reaction (OER), respectively [3,4]. In particular, OER is more energy-intensive and considered to be the primary bottleneck for efficient processes due to its slow kinetics [5]. Consequently, developing electrocatalysts with low overpotential for OER is imperative. Transition metal chalcogenides (TMCs) have emerged as promising OER catalysts owing to their high activity, cost-effectiveness, controllable composition, and morphology, offering potential solutions to challenges in the OER process [6]. Compared to sulfur, selenium (Se) exhibits a larger atomic radius, lower ionization energy, and abundant d-electrons that facilitate charge transfer [7]. Additionally, the presence of Se atoms in metal selenides enhances the electrophilicity of metal atoms and augments the affinity of selenides for hydroxyl species [8].

    However, recent studies indicate that under high anodic potentials, transition metal selenides (TMSe) undergo surface oxidation and reconstruction, forming amorphous transition-metal oxyhydroxides (TMOOH), which are the actual active species instead of TMSe [9]. This reconstruction complicates the study of the electrochemical catalytic mechanism and leads to the dissolution of active metal species, resulting in poor catalyst stability [10]. Preventing TMSe surface reconstruction is thus critical for understanding structure-activity relationships and improving catalytic durability. Nonetheless, inhibiting this reconstruction during OER remains a significant challenge. Additionally, most reported TMSe catalysts exhibit high activity only at lower current densities (<100 mA/cm2), falling short of industrial-scale requirements due to high overpotentials associated with Gibbs free energies and rate-determining steps in *OOH intermediate formation [11]. A promising approach for enhancing catalyst durability is to develop the 2D cover-confined nanomaterials, also referred to as chainmail catalysts, which effectively shield enclosed nanometals from damage during rigorous reaction processes [12-14]. This innovative strategy not only improves catalytic activity but also ensures the stability of inner nanocatalysts, earning it the moniker ‘chainmail’ for catalyst [15].

    Among TMSe, cobalt-based selenides such as Co0.85Se, CoSe, CoSe2, and Co3Se4 have shown potential. Non-stoichiometric Co0.85Se, in particular, is expected to exhibit superior electrochemical performance due to the interaction between Co 3d and Se 4p spin electrons, enhancing conductivity [16,17]. MXenes, a category of 2D transition metal carbides, have attracted significant interest for their remarkable electrical conductivity, mechanical rigidity, large surface area, and outstanding electrochemical properties [18]. Synthesized by chemically etching the MAX phase (Mn+1AXn, in which M represents an early transition metal, A stands for a group Ⅲ/ ⅣA element, and X denotes either carbon or nitrogen), MXenes feature van der Waals gaps with layer spacing of 0.6–1.2 nm, providing an ideal confined platform for nanocatalysts [19].

    In this work, Co0.85Se@Ti3C2 was synthesized by first etching Ti3AlC2 to obtain Ti3C2−MXene, followed by solvothermal growth of Co0.85Se nanoparticles within its interlayer space. The interlayer confinement stabilizes Co0.85Se by preventing surface reconstruction and generates an internal electric field that accelerates charge transfer rate at the electrode interface. The material exhibits superhydrophilicity (water contact angle of 9.1°) and underwater superaerophobicity (bubble contact angle of 151.1°). The catalyst demonstrates low overpotentials of 230 mV at 100 mA/cm2, 376 mV at 1000 mA/cm2, and 417 mV at 1500 mA/cm2, with a Tafel slope of 59 mV/dec in 1 mol/L KOH, maintaining excellent stability over 200 h at 200 mA/cm2. Density functional theory (DFT) is introduced, which revealed that the Co0.85Se@Ti3C2 interaction reduces the energy barrier for *OOH formation from *O, enabling exceptional performance in high-current OER processes. This work provides a foundation for designing non-reconstructed OER catalysts with high activity at elevated current densities through confined structures with internal electric fields.

    Fig. 1a presents a schematic diagram illustrating the preparation process of the Co0.85Se@Ti3C2 catalyst. Fig. 1b displays the XRD patterns of both raw materials and synthesized products. The diffraction peaks for the raw Ti3AlC2 align with its standard card JCPDS No. 52–0875. Followed by the etching of Ti3AlC2 using a concentrated hydrochloric acid and lithium fluoride mixture, the (002) peak shifts to lower angles of 7.2°, indicating an increase in interlayer distance due to the dissolution of aluminum layers within the crystal structure, ultimately leading to the formation of two-dimensional layered Ti3C2 material. Furthermore, the characteristic peaks for synthesized Co0.85Se and Co0.85Se@Ti3C2 in Fig. 1b closely match well with the standard card for Co0.85Se (JCPDS No. 52–1008), with no additional impurity peaks [20,21]. Notably, the Ti3C2 characteristic peak in Co0.85Se@Ti3C2 is shifted further left to 5.8° compared to that of pure Ti3C2 (the partial XRD within the red dashed box was magnified and is presented in Fig. 1b, resulting in an increase in interlayer distance from 1.21 nm to 1.52 nm, which manifests the successful incorporation of Co0.85Se nanoparticles into the interlayer space of Ti3C2, thereby causing further expansion.

    Figure 1

    Figure 1.  (a) Schematic illustration of the process for the catalyst Co0.85Se@Ti3C2 preparation. (b) XRD pattern of the catalysts of Ti3AlC2, Ti3C2, Co0.85Se and Co0.85Se@Ti3C2 and the corresponding local enlarged view. (c) SEM image of Ti3C2. (d) SEM image of Co0.85Se@Ti3C2. (e) Distribution of C, Ti, Co, and Se in Co0.85Se@Ti3C2 catalyst. (f) TEM image and SAED pattern of the Co0.85Se@Ti3C2. (g) HRTEM image of the Co0.85Se@Ti3C2.

    Fig. 1c depicts the SEM image of Ti3C2, revealing a two-dimensional accordion-like multilayer structure with a smooth surface. In Fig. 1d, the SEM image of the Co0.85Se@Ti3C2 composite material illustrates a rough surface due to the distribution of the Co0.85Se nanoparticles on the surface and layer space of Ti3C2. Fig. S1 (Supporting information) reveals that pure Co0.85Se demonstrates obvious agglomeration. The EDS mapping (Fig. 1e) of the Co0.85Se@Ti3C2 catalyst shows uniform element distribution, indicating the uniformly dispersed Co0.85Se nanoparticles on the surface of layered Ti3C2. Fig. 1f presents TEM images and electron diffraction patterns for the ultrasonically agitated Co0.85Se@Ti3C2 sample. The TEM images reveal uniformly sized Co0.85Se particles with ultrafine dimensions, while electron diffraction rings correspond to (101), (102), and (110) crystal planes of Co0.85Se and (110) crystal plane of Ti3C2. The HRTEM image (Fig. 1g) demonstrates that the Co0.85Se@Ti3C2 material possesses a heterostructure with varying interlayer spacings in two distinct regions: one layer spacing at 0.27 nm belonging to (101) plane of Co0.85Se and the another layer spacing at 1.52 nm belonging to (002) plane of Ti3C2.

    The full XPS spectrum of the composite (Fig. S2 in Supporting information) confirms the presence of Co, Se, Ti, C, and O. The high-resolution XPS spectrum of C 1s (Fig. 2a) for Co0.85Se@Ti3C2 shows peaks at 281.6, 283.8, 284.8, and 287.8 eV, corresponding to C-Ti, C-Ti-O, C—C, and C=O, respectively [22]. The O 1s spectrum (Fig. 2b) features a prominent peak at 530.5 eV and a weaker peak at 532.1 eV, attributed to Ti-O and Ti-OH groups, indicating oxygen primarily as a terminating group in Ti3C2 [23]. The Co 2p spectrum (Fig. 2c) displays peaks at 778.3 and 780.9 eV for the Co 2p3/2, and at 793.3 and 796.9 eV for Co 2p1/2, with satellite peaks at 784.6 and 802.6 eV [24]. Fig. 2d presents the detailed spectrum of Se 3d in Co0.85Se@Ti3C2 material, with binding energies of 54.2 and 55.1 eV corresponding to the Se 3d5/2 and Se 3d3/2 peaks, respectively. The peaks observed at 58.86 and 60.2 eV are attributed to SeOx, suggesting that the surface of the material has undergone oxidation in ambient air [25].

    Figure 2

    Figure 2.  (a, b) The XPS spectra of C 1s, O 1s. (c, d) The XPS spectra of Co 2p, Se 3d for the catalyst Co0.85Se and Co0.85Se@Ti3C2. (e, f) The UPS spectra of Ti3C2, Co0.85Se. (g) Schematic diagram of the formation of built-in electric field between Ti3C2 and Co0.85Se. EVac: vacuum level, EF: fermi level, EV: valence band edge, EC: conduction band edge.

    Compared to pure Co0.85Se, the Co and Se peaks in Co0.85Se@Ti3C2 shift to lower binding energies due to enhanced electron screening from charge density redistribution at the interface [26-28], suggesting charge transfer from Ti3C2 to Co0.85Se. Furthermore, the comparative investigation of the peak shifts of Ti 2p between Ti3C2 and Co0.85Se@Ti3C2 further substantiated the aforementioned discourse (Fig. S3 in Supporting information). The semiconductor type was determined using Mott-Schottky (MS) curves, revealing p-type characteristics for Co0.85Se (Fig. S4). Ultraviolet photoelectron spectroscopy (UPS) measured the work functions (Φ), with Φ(Ti3C2) calculated as 2.79 eV and Φ(Co0.85Se) as 3.62 eV (Figs. 2e and f). The energy level diagram (EV vs. vacuum level) and the conduction band diagram (EC vs. vacuum level) are presented in the schematic diagram of the energy band structure in Fig. 2g, visualizing the effect of the built-in electric field formed between Co0.85Se and Ti3C2. It can be observed that once Co0.85Se comes into contact with Ti3C2, due to the corresponding disparity in Fermi levels, a Schottky junction with built-in electric field will be established, and electrons will spontaneously flow from Ti3C2 to p-type semiconductor Co0.85Se unobstructedly. This unidirectional electron transfer is conducive to enriching electrons in the Co0.85Se region and depleting electrons in the Ti3C2 region, thereby further accelerating the electrocatalytic process. Consequently, the BEF structure of Co0.85Se@Ti3C2 is beneficial for providing optimized active sites, thereby facilitating the adsorption of oxygen intermediates [29-32].

    The synthesized samples were subsequently investigated for their electrocatalytic performance in the OER under alkaline conditions with iR-compensation. As shown in Fig. 3a, the LSV curves of Co0.85Se, Ti3C2, Co0.85Se@Ti3C2, and RuO2 were obtained for electrochemical OER analysis. It is evident that Co0.85Se@Ti3C2 demonstrated superior electrocatalytic activity of the lowest onset potential. Fig. 3b was the comparison of overpotentials, in which Co0.85Se@Ti3C2 exhibited the lowest overpotential for OER at high current densities of 100, 1000 and 1500 mA/cm2, requiring only 230, 376 and 417 mV overpotentials, respectively, while the pure Co0.85Se displayed significantly higher overpotentials of 415, 603 and 729 mV at the respective current densities. Furthermore, Ti3C2 and RuO2 attained overpotentials of 600 and 490 mV respectively at 100 mA/cm2, but failed to reach the high current densities mentioned above. Table S1 (Supporting information) highlights the superior OER performance of Co0.85Se@Ti3C2 compared to other catalysts.

    Figure 3

    Figure 3.  (a) LSV measurements of different samples for OER in 1.0 mol/L KOH electrolyte. (b) The comparison of the overpotential to Ti3C2, RuO2, Co0.85Se and Co0.85Se@Ti3C2. (c) Tafel slope derived from (a). (d) EIS Nyquist plots of different samples. Inset: the equivalent circuit for the EIS data fitting. (e) Capacitive currents at 1.17 V (vs. RHE) as a function of scan rate for Ti3C2, Co0.85Se, and Co0.85Se@Ti3C2. (f) The TOF to Ti3C2, Co0.85Se, Co0.85Se@Ti3C2. (g) Multi-current testing of Co0.85Se@Ti3C2 catalysts in the range of 200–1000 mA/cm2. (h) Chronopotential curves for Co0.85Se@Ti3C2 at 200 mA/cm2.

    In Fig. 3c, Co0.85Se@Ti3C2 exhibits the smallest Tafel slope (59 mV/dec), which is significantly surpasses the Co0.85Se (94 mV/dec), Ti3C2 (111 mV/dec), and RuO2 (102 mV/dec), suggesting its more rapid reaction kinetics. Moreover, the Tafel plot indicates that Co0.85Se@Ti3C2 possesses the highest exchange current density (j0) of 6.67 mA/cm2, as detailed in Table S2 (Supporting information). Additionally, the obtained Tafel slope is in alignment with the slope derived from electrochemical impedance spectroscopy (EIS) (Fig. 3d). The charge transfer resistance (Rct) was determined by fitting the Nyquist plots with resistance-capacitance circuits (inset in Fig. 3d, solution resistance (Rs), constant phase element (CPE)). Co0.85Se@Ti3C2 exhibits an Rct of 4.52 Ω, which is significantly lower than that of Ti3C2 (Rct = 20.67 Ω), and Co0.85Se (Rct = 6.12 Ω), indicating a superior electron transfer efficiency and reduced charge transfer resistance than other electrocatalysts.

    To acquire the actual alkaline OER activities of the fabricated electrodes, the electrochemical surface active area (ECSA) was ascertained through double-layer capacitance (Cdl) [33]. The Cdl were obtained from cyclic voltammetry (CV) tests conducted at a non-faradic potential region (Fig. S5 in Supporting information), which represent the ECSA of the catalysts. The current density exhibits a linear correlation with the scan rates, and the value of Cdl is calculated based on the slope of this linear relationship.

    As depicted in Fig. 3e, the Co0.85Se@Ti3C2 composite exhibits a significantly higher Cdl of 28.98 mF/cm2 compared to the individual components, Co0.85Se (9.73 mF/cm2) and Ti3C2 (4.25 mF/cm2). This elevated Cdl corresponds to the highest electrochemically active surface area (ECSA) of 724.5 cm2 for Co0.85Se@Ti3C2, outperforming both Co0.85Se (243.3 cm2) and Ti3C2 (106.3 cm2), as well as other materials reported in prior studies [34-36]. These findings suggest that a well-designed composite structure can enhance the adsorption of surface species, resulting in a dense Helmholtz double layer between the interface of electrolyte and catalyst with high surface capacitance, which subsequently accelerates the OER process. Furthermore, the BEF at the Co0.85Se@Ti3C2 interface, arising from electron transfer from MXene to Co0.85Se, promotes electron transfer rate and enhances the charge transfer capacity during OER. This is corroborated by the electrochemical impedance spectroscopy (EIS) results shown in Fig. 3d. In Fig. 3f, Co0.85Se@Ti3C2 attains significantly higher turnover frequency (TOF) values at diverse overpotentials compared with those of Ti3C2 and Co0.85Se, which further substantiates the remarkable electrocatalytic performance of Co0.85Se@Ti3C2.

    The Faradaic efficiency (FE) of Co0.85Se@Ti3C2 for OER was determined by measuring the oxygen volume obtained at a constant current of 0.05 A using a water drainage method (Fig. S6 in Supporting information) [37]. The measured O2 vol-time curve closely aligns with the theoretical value (Fig. S7 in Supporting information), indicating a high FE (>98%) for the OER process on Co0.85Se@Ti3C2. Long-term stability is a critical factor in evaluating the OER properties of a catalyst during practical application. As illustrated in Fig. 3g, a chronoamperometry curve with multiple current steps was performed, and the immediate response demonstrated outstanding mass transport capability. Furthermore, there is minimal change in potential even at a relatively high current density of 1000 mA/cm2, suggesting exceptional durability for OER. Additionally, the chronopotentiometric curves show that the material can sustain 200 mA/cm2 for at least 200 h with only decay by 4% (Fig. 3h), indicating its potential for industrial applications. Despite significant gas evolution during extended electrocatalysis on Co0.85Se@Ti3C2, oxygen bubbles quickly dissipate into the electrolyte solution without accumulating on the material surfaces due to its excellent gas release performance, which will be further discussed later.

    High oxidation potentials often result in significant deterioration to the physical phase and structure of the catalyst during OER, particularly metal selenides. In order to investigate the structural reconstruction of Co0.85Se@Ti3C2, in-situ XRD and Raman were conducted. The schematic diagrams for in-situ XRD and Raman measurements are depicted in Figs. 4a and b. During the XRD detection, continuous CV scanning was performed from 1.13 V to 1.53 V (vs. RHE). As shown in Fig. 4c, the in-situ XRD showed no significant change in characteristic peaks for the lattice plane of (311) and (222) of Co0.85Se (JCPDS No. 52–1008), suggesting that there was no structural reconstruction during OER. Additionally, no diffraction peaks corresponding to CoOOH and Co(OH)2 were observed in the XRD patterns, further confirming that the restructuring was hindered [38-40]. The in-situ Raman spectra of the Co0.85Se@Ti3C2 catalyst at various potentials were given in Fig. 4d. Surprisingly, even at a high voltage of 1.8 V, there is no discernible characteristic peak of CoOOH appeared (range: 400–750 cm-1) in the spectrum. In contrast, the pure Co0.85Se sample without Ti3C2 displayed distinct surface reconstruction peaks at a voltage of 1.23 V during in-situ Raman, which were identified by the generation of CoOOH and Co(OH)2 in Fig. 4e [41-44]. Furthermore, the ex-situ XPS spectrum (Figs. S8a and b in Supporting information) of Co0.85Se@Ti3C2 after the OER test still presents the intense characteristic peaks of cobalt selenide, encompassing distinct Co 2p1/2, Co 2p3/2, Se 3d3/2, and Se 3d5/2 peaks. Meanwhile, the SeOx peak does not exhibit a significant increase. This is strikingly dissimilar from the situation where the Se 3d feature peaks of other metal selenides weaken or even vanish after the OER test and reconstruction [45-47]. Additionally, the O 1s (Fig. S8c in Supporting information) spectrum mainly corresponds to the oxygen-terminated groups on the surface of Ti3C2 [48], which is almost the same as that of the sample before the OER test.

    Figure 4

    Figure 4.  (a) Schematic diagram of the principle of in-situ X-ray diffraction. (b) Schematic diagram of the in-situ Raman device. (c) In-situ XRD plot of the catalyst Co0.85Se@Ti3C2 CV cycling at 1.098–1.498 (V vs. RHE) potential. (d, e) In-situ Raman plot of catalyst Co0.85Se@Ti3C2 and Co0.85Se performed at OCP-1.73 V (V vs. RHE) potential. OCP: circuit potential.

    The wettability of the catalyst significant influence the exposure of surface active sites to the electrolyte [49]. Moreover, the accumulation of bubbles on the catalyst surface would lead to a "bubble shielding effect", which impeded the contact between the electrolyte and catalyst [50]. Therefore, constructing catalysts with exceptional hydrophilicity and aerophobicity properties would enhance OER processes. To analyze the wettability properties of various catalysts, contact angle (CA) measurements are performed (Fig. 5a). The results indicate that Co0.85Se@Ti3C2 exhibits the smallest water CA at 9.1°, demonstrating remarkable hydrophilicity compared to Co0.85Se (26.7°) and Ti3C2 (32.6°). Furthermore, Co0.85Se@Ti3C2 demonstrates superior aerophobicity with a larger bubble CA (151.1°) compared to Co0.85Se (121.8°) and Ti3C2 (96.3°) (Fig. 5b). The above results demonstrate that the promoted efficient mass transfer and gas bubble release on the surface of Co0.85Se@Ti3C2. The enhanced capability of Co0.85Se@Ti3C2 to attract water is primarily attributed to its composite structure, which facilitates electron redistribution, amplifies the material’s polarity, and consequently enhances its affinity for polar water molecules. Additionally, in accordance with the solid-liquid-gas interface theory, the uniformly distributed of Co0.85Se particles on the surface of Ti3C2 will also enhance the roughness of the surface, which will give rise to the formation of a three-phase interface with discontinuous contact, leading to an extremely small contact area between the bubble and the electrode surface and thereby facilitating the desorption of gaseous products [51]. Consequently, it can be inferred that the composite structure promotes the adsorption of H2O molecules nevertheless, and improves the desorption capacity of O2 molecules on the surface of catalysts, resulting in heightened reactivity at active sites and improved performance in OER.

    Figure 5

    Figure 5.  The water droplet (a) and air bubble (b) contact angles measurement of Ti3C2, Co0.85Se, Co0.85Se@Ti3C2. The densities of states of Ti3C2 (c), Co0.85Se (d), and Co0.85Se@Ti3C2 (e). (f) Adsorption energy of H2O molecules. (g) The four step processes of OER. (h) Step diagram of Co0.85Se and Co0.85Se@Ti3C2 OER reaction process.

    The density functional theory (DFT) approach was employed to further clarify the mechanism accountable for the improved OER performance. The calculation models of Ti3C2, Co0.85Se, and Co0.85Se@Ti3C2 heterojunction are shown in Fig. S9 (Supporting information). The density of states (DOS) was computed (Figs. 5c-e), suggesting that the Co0.85Se@Ti3C2 heterojunction offers a higher density of electronic cloud close to the Fermi level, which is beneficial for lowering the reaction energy barrier of water splitting [52] and thereby enhancing the catalytic activity. Further d-band center computations reveal that the d-band centers of Ti3C2, Co0.85Se, and Co0.85Se@Ti3C2 are −0.89, −1.08, and −0.40 eV respectively. This indicates that, in line with the d-band theory, Co0.85Se@Ti3C2 has higher valence states and more antibonding states [53]. Therefore, Co0.85Se@Ti3C2 is beneficial for adsorption, thus enhancing the electrocatalytic activity. Fig. 5f shows that the adsorption energy of H2O molecules on Co0.85Se@Ti3C2 is −0.72 eV, which is lower than that of Co0.85Se (−0.25 eV) and Ti3C2 (−0.057 eV). Consequently, it can be inferred that the heterostructure promotes the adsorption of H2O molecules on the surface of the catalysts, thereby increasing the reactivity at active sites and improving the OER performance. Fig. 5g discloses that the OER process involves four consecutive proton-coupled electron transfer steps, including the intermediates of *OH, *O, and *OOH [54]. As presented in Fig. 5h, the transfer of electrons from *O to *OOH (Step Ⅲ) has the most significant energy barrier, suggesting that it is the rate-determining step governing the rate of the OER process for Co0.85Se@Ti3C2, which has a lower energy barrier of 2.50 eV compared to Co0.85Se (3.49 eV). Therefore, the DFT analysis verifies that the Co0.85Se@Ti3C2 heterostructure exhibits superior OER activity, which is in accordance with the experimental results.

    In summary, a novel electrocatalyst (Co0.85Se@Ti3C2) featuring a confined interlayer structure was proposed for alkaline OER, which demonstrates remarkable catalytic activity (376 mV@1000 mA/cm2, and 417 mV@1500 mA/cm2) and exceptional long-term durability up to 200 h at a high current density of 200 mA/cm2. Both experimental and theoretical studies validated that the confined structure, the built-in electric field, and the superhydrophilic/oleophobic properties synergistically optimize the spatial environment and electronic configuration of Co0.85Se, thereby preventing the reconstruction of Co0.85Se and reducing the energy barrier for the OER process and enhancing the intrinsic activity. This investigation establishes the foundation for designing efficient and non-reconstructive transition metal chalcogenide electrocatalysts and promotes progress in the OER field.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    Yuanhua Xiao: Writing – review & editing. Jinhui Shou: Writing – original draft. Shiwei Zhang: Software. Ya Shen: Software. Junwei Liu: Resources. Dangcheng Su: Software. Yang Kong: Resources. Xiaodong Jia: Resources. Qingxiang Yang: Resources. Shaoming Fang: Resources. Xuezhao Wang: Resources.

    The authors are grateful to the financial support from the National Natural Science Foundation of China (NSFC, Nos. U23A20579, U1904190, 52272243), the Natural Science Foundation of Henan Province (No. 242300421467), the Program for Science and Technology Innovation Talents in Universities of Henan Province (No. 22HASTIT005), Key Scientific Research Projects in Higher Education Institutions of Henan Province (Nos. 24A430047, 24A430029), China Postdoctoral Science Foundation (No. 2023M741083), Research Foundation for Talented Scholars (No. 21010744).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2025.111441.


    1. [1]

      N. Zhao, L. Mu, W. Liao, et al., J. Colloid Interface Sci. 682 (2025) 148–156. doi: 10.1016/j.jcis.2024.11.193

    2. [2]

      Y. Xiao, S. Zhang, Y. Shen, et al., J. Colloid Interface Sci. 664 (2024) 329–337. doi: 10.1016/j.jcis.2024.03.041

    3. [3]

      L. Mu, S. Qiu, G. Zhao, et al., J. Mater. Chem. A 12 (2024) 1714–1724. doi: 10.1039/d3ta06701g

    4. [4]

      Y. Xiao, Y. Shen, D. Su, et al., J. Mater. Sci. Technol. 154 (2023) 1–8. doi: 10.1016/j.jmst.2022.12.042

    5. [5]

      G. Zhao, L. Mu, L. Chen, et al., Int. J. Hydrog. Energy 88 (2024) 432–440. doi: 10.1016/j.ijhydene.2024.09.114

    6. [6]

      M. Li, H. Li, H. Fan, et al., Nat. Commun. 15 (2024) 6154. doi: 10.1038/s41467-024-50535-2

    7. [7]

      A. Du, Y. Zhao, Z. Zhang, et al., Energy Storag. Mater. 26 (2020) 23–31. doi: 10.1016/j.ensm.2019.12.030

    8. [8]

      Y. Cheng, L. Zhu, Y. Gong, Int. J. Hydrog. Energy 92 (2024) 1049–1058. doi: 10.1016/j.ijhydene.2024.09.267

    9. [9]

      J. Zhao, F. Wang, X. Lu, et al., J. Colloid Interface Sci. 642 (2023) 532–539. doi: 10.3390/agronomy13020532

    10. [10]

      Q. Wang, Y. Cheng, H.B. Tao, et al., Angew. Chem. Int. Ed. 62 (2023) e202216645. doi: 10.1002/anie.202216645

    11. [11]

      X. Wang, S. Xi, P. Huang, et al., Nature 611 (2022) 702–708. doi: 10.1038/s41586-022-05296-7

    12. [12]

      Y. Xiao, H. Jiang, K. Zhang, et al., Chem. Eng. J. 492 (2024) 152274. doi: 10.1016/j.cej.2024.152274

    13. [13]

      Y. Xiao, Y. Kong, X. Wang, et al., J. Colloid Interface Sci. 677 (2025) 577–586. doi: 10.1016/j.jcis.2024.07.245

    14. [14]

      Y. Xiao, Q. Le, Y. Kong, et al., Chem. Eng. J. 505 (2025) 159268. doi: 10.1016/j.cej.2025.159268

    15. [15]

      L. Tang, X. Meng, D. Deng, et al., Adv. Mater. 31 (2019) 1901996. doi: 10.1002/adma.201901996

    16. [16]

      L. Li, S. Wang, J. Peng, et al., Molecules 29 (2024) 3083. doi: 10.3390/molecules29133083

    17. [17]

      Z. Jiao, Y. Chen, M. Demir, et al., J. Energy Storag. 52 (2022) 104929. doi: 10.1016/j.est.2022.104929

    18. [18]

      A. Hanan, M. Nazim, R. Walvekar, et al., Chem. Eng. J. 483 (2024) 149107. doi: 10.1016/j.cej.2024.149107

    19. [19]

      F. Bu, M.M. Zagho, Y. Ibrahim, et al., Nano. Today 30 (2020) 100803. doi: 10.1016/j.nantod.2019.100803

    20. [20]

      X. Sun, N. Habibul, H. Du, Chin. J. Catal. 42 (2021) 235–243. doi: 10.1016/S1872-2067(20)63632-4

    21. [21]

      C. Atangana, H. Huang, H. Hong, et al., Energ. Storag. Mater. 24 (2020) 167–176. doi: 10.1016/j.ensm.2019.08.022

    22. [22]

      R. Zhao, H. Di, X. Hui, et al., Energ. Environ. Sci. 14 (2021) 5096 -5096. doi: 10.1039/d1ee90043a

    23. [23]

      C. Sun, X. Shi, Y. Zhang, et al., ACS Nano 14 (2020) 1176–1184. doi: 10.1021/acsnano.9b09541

    24. [24]

      J. Liang, S. Li, F. Li, et al., J. Colloid Interface Sci. 655 (2024) 296–306. doi: 10.1016/j.jcis.2023.11.010

    25. [25]

      K. Li, R. Cheng, Q. Xue, et al., Chem. Eng. J. 450 (2022) 137991. doi: 10.1016/j.cej.2022.137991

    26. [26]

      L. Jin, H. Xu, K. Wang, et al., Chem. Sci. 16 (2025) 329–337. doi: 10.1039/d4sc06673a

    27. [27]

      L. Mu, G. Zhao, B. Zhang, et al., J. Colloid Interface Sci. 677 (2025) 68–78.

    28. [28]

      H. Xu, L. Jin, K. Wang, et al., Fuel 369 (2024) 131716. doi: 10.1016/j.fuel.2024.131716

    29. [29]

      K. Wang, H. Xu, H. Xing, et al., Chem. Eng. J. 500 (2024) 157316. doi: 10.1016/j.cej.2024.157316

    30. [30]

      B. Zhang, N. Zhang, G. Zhao, et al., J. Colloid Interface Sci. 665 (2024) 1054–1064. doi: 10.1016/j.jcis.2024.04.002

    31. [31]

      H. Xu, L. Yang, Y. Liu, et al., Fuel 377 (2024) 132796. doi: 10.1016/j.fuel.2024.132796

    32. [32]

      X. Luan, Z. Zheng, S. Zhao, et al., Adv. Funct. Mater. 32 (2022) 2202843. doi: 10.1002/adfm.202202843

    33. [33]

      X. Zhou, Y. Mo, F. Yu, et al., Adv. Funct. Mater. 33 (2023) 2209465. doi: 10.1002/adfm.202209465

    34. [34]

      B. Lu, Z. Li, J. Yin, et al., Appl. Catal. B: Environ. 350 (2024) 123940. doi: 10.1016/j.apcatb.2024.123940

    35. [35]

      Y. Yu, Y. Wu, W. Luo, et al., Chem. Eng. J. 477 (2023) 147076. doi: 10.1016/j.cej.2023.147076

    36. [36]

      L. He, N. Wang, M. Xiang, et al., Appl. Catal. B 345 (2024) 123686. doi: 10.1016/j.apcatb.2023.123686

    37. [37]

      Y. Song, M. Sun, S. Zhang, et al., Adv. Funct. Mater. 33 (2023) 2214081. doi: 10.1002/adfm.202214081

    38. [38]

      Q. Ji, Y. Kong, H. Tan, et al., ACS Catal. 12 (2022) 4318–4326. doi: 10.1021/acscatal.2c01090

    39. [39]

      S. Ye, J. Wang, J. Hu, et al., ACS Catal. 11 (2021) 6104–6112. doi: 10.1021/acscatal.1c01300

    40. [40]

      F. Reikowski, F. Maroun, I. Pacheco, et al., ACS Catal. 9 (2019) 3811–3821. doi: 10.1021/acscatal.8b04823

    41. [41]

      K. Wang, Y. Li, J. Hu, et al., Chem. Eng. J. 447 (2022) 137540. doi: 10.1016/j.cej.2022.137540

    42. [42]

      M. Chen, D. Liu, J. Feng, et al., Chem. Eng. J. 443 (2022) 136432. doi: 10.1016/j.cej.2022.136432

    43. [43]

      Y. Zhi, Z. Li, Y. Tang, et al., Nano Energ. 134 (2025) 110565. doi: 10.1016/j.nanoen.2024.110565

    44. [44]

      Z. Ji, W. Yuan, S. Zhao, et al., Chem. Catal. 3 (2023) 100501.

    45. [45]

      Y. Huang, L. Zhang, L.W. Jiang, et al., Small 19 (2023) 2302970. doi: 10.1002/smll.202302970

    46. [46]

      S. Feng, P. Rao, Y. Yu, et al., Electrochim. Acta 463 (2023) 142870. doi: 10.1016/j.electacta.2023.142870

    47. [47]

      M. Chen, Y. Zhang, R. Wang, et al., J. Energy Chem. 84 (2023) 173–180. doi: 10.1016/j.jechem.2023.05.009

    48. [48]

      T. Hou, Q. Luo, Q. Li, et al., Nat. Commun. 11 (2020) 4251. doi: 10.1038/s41467-020-18091-7

    49. [49]

      X. Xu, H. Liao, L. Huang, et al., Appl. Catal. B 341 (2024) 123312. doi: 10.1016/j.apcatb.2023.123312

    50. [50]

      J.W. Yu, K.D. Hu, Z.Y. Zhang, et al., Energy Environ. Sci. 16 (2023) 2068–2079. doi: 10.1039/d2ee04143j

    51. [51]

      X. Shan, J. Liu, H. Mu, et al., Angew. Chem. Int. Ed. 59 (2020) 1659–1665. doi: 10.1002/anie.201911617

    52. [52]

      H. Li, Q. Yu, X. Zhu, et al., Chem. Eng. J. 493 (2024) 152860. doi: 10.1016/j.cej.2024.152860

    53. [53]

      Z. Shi, C. Mao, L. Zhong, et al., Appl. Catal. B: Environ. 339 (2023) 123123. doi: 10.1016/j.apcatb.2023.123123

    54. [54]

      X. Shen, H. Li, Y. Zhang, et al., Appl. Catal. B: Environ. 319 (2022) 121917. doi: 10.1016/j.apcatb.2022.121917

  • Figure 1  (a) Schematic illustration of the process for the catalyst Co0.85Se@Ti3C2 preparation. (b) XRD pattern of the catalysts of Ti3AlC2, Ti3C2, Co0.85Se and Co0.85Se@Ti3C2 and the corresponding local enlarged view. (c) SEM image of Ti3C2. (d) SEM image of Co0.85Se@Ti3C2. (e) Distribution of C, Ti, Co, and Se in Co0.85Se@Ti3C2 catalyst. (f) TEM image and SAED pattern of the Co0.85Se@Ti3C2. (g) HRTEM image of the Co0.85Se@Ti3C2.

    Figure 2  (a, b) The XPS spectra of C 1s, O 1s. (c, d) The XPS spectra of Co 2p, Se 3d for the catalyst Co0.85Se and Co0.85Se@Ti3C2. (e, f) The UPS spectra of Ti3C2, Co0.85Se. (g) Schematic diagram of the formation of built-in electric field between Ti3C2 and Co0.85Se. EVac: vacuum level, EF: fermi level, EV: valence band edge, EC: conduction band edge.

    Figure 3  (a) LSV measurements of different samples for OER in 1.0 mol/L KOH electrolyte. (b) The comparison of the overpotential to Ti3C2, RuO2, Co0.85Se and Co0.85Se@Ti3C2. (c) Tafel slope derived from (a). (d) EIS Nyquist plots of different samples. Inset: the equivalent circuit for the EIS data fitting. (e) Capacitive currents at 1.17 V (vs. RHE) as a function of scan rate for Ti3C2, Co0.85Se, and Co0.85Se@Ti3C2. (f) The TOF to Ti3C2, Co0.85Se, Co0.85Se@Ti3C2. (g) Multi-current testing of Co0.85Se@Ti3C2 catalysts in the range of 200–1000 mA/cm2. (h) Chronopotential curves for Co0.85Se@Ti3C2 at 200 mA/cm2.

    Figure 4  (a) Schematic diagram of the principle of in-situ X-ray diffraction. (b) Schematic diagram of the in-situ Raman device. (c) In-situ XRD plot of the catalyst Co0.85Se@Ti3C2 CV cycling at 1.098–1.498 (V vs. RHE) potential. (d, e) In-situ Raman plot of catalyst Co0.85Se@Ti3C2 and Co0.85Se performed at OCP-1.73 V (V vs. RHE) potential. OCP: circuit potential.

    Figure 5  The water droplet (a) and air bubble (b) contact angles measurement of Ti3C2, Co0.85Se, Co0.85Se@Ti3C2. The densities of states of Ti3C2 (c), Co0.85Se (d), and Co0.85Se@Ti3C2 (e). (f) Adsorption energy of H2O molecules. (g) The four step processes of OER. (h) Step diagram of Co0.85Se and Co0.85Se@Ti3C2 OER reaction process.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  134
  • HTML全文浏览量:  10
文章相关
  • 发布日期:  2025-11-15
  • 收稿日期:  2025-01-27
  • 接受日期:  2025-06-10
  • 修回日期:  2025-06-05
  • 网络出版日期:  2025-06-11
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章