Heyang
Liu
a,
Fengli
Wei
a,
Linlin
Huang
a,
Chenggong
Niu
a,
Zuyang
Luo
a,
Tayirjan Taylor
Isimjan
*b and
Xiulin
Yang
*a
aGuangxi Key Laboratory of Low Carbon Energy Materials, School of Chemistry and Pharmaceutical Sciences, Guangxi Normal University, Guilin 541004, China. E-mail: xlyang@gxnu.edu.cn
bSaudi Arabia Basic Industries Corporation (SABIC) at King Abdullah University of Science and Technology (KAUST), Thuwal 23955-6900, Saudi Arabia. E-mail: isimjant@sabic.com
First published on 13th October 2025
Dynamic surface reconstruction offers a promising route to enhance oxygen evolution reaction (OER) activity by optimizing the adsorption of key intermediates. Here, we construct a pearl-thread-like CoS2–CuxS heterostructure on copper foam (CoS2–CuxS/CF) to induce in situ formation of an active CoOOH–CuO–CuxS interface at low potentials. In situ Raman spectroscopy confirms the dynamic transformation, while operando electrochemical impedance spectroscopy reveals accelerated charge transfer. Density functional theory calculations show that the interface engineering shifts the d-band center, enhances electron density near the Fermi level, and lowers the free energy barrier for *O to *OOH conversion from 1.78 eV (CoOOH) to 1.48 eV (CoOOH–CuO–CuxS). Benefiting from the reconstructed interface, CoS2–CuxS/CF achieves an overpotential of 239 mV at 10 mA cm−2 and maintains stability for over 200 h in alkaline electrolyte. This work highlights a dynamic interface strategy to promote intrinsic OER kinetics and catalyst durability.
O bonds, resulting in high overpotentials and poor energy efficiency.3–5 Noble metal-based catalysts, such as RuO2 and IrO2, have been widely employed to mitigate these kinetic challenges by optimizing the adsorption energy of oxygenated intermediates.6 Nevertheless, noble metals’ limited availability and high cost severely restrict their widespread application in large-scale water splitting systems.7
To address these limitations, substantial efforts have focused on developing non-noble transition metal-based catalysts that are Earth-abundant and cost-effective and possess tunable electronic structures.8,9 Among these, copper-based materials have garnered significant attention in OER systems due to their unique electronic configuration and high natural abundance. For instance, Zhu's group demonstrated that CuNCo3 nanosheets achieved an overpotential of 260 mV at 10 mA cm−2.10 Chen's group electrodeposited CoS on Cu(OH)2 nanorods to form three-dimensional hierarchical core–shell electrocatalysts, achieving an overpotential of 296 mV at the same current density.11 However, Cu-based catalysts face inherent limitations, including high charge transfer resistance and sluggish reaction kinetics, which hinder their practical applications.12 However, intrinsic drawbacks, including sluggish charge transfer kinetics and suboptimal binding energies for key OER intermediates, limit Cu-based materials’ catalytic efficiency.13,14
Interface engineering has recently emerged as a powerful strategy to overcome these challenges by modulating the local electronic environment at heterostructured interfaces, thus enhancing electron transfer, tuning the d-band center, and facilitating optimal adsorption/desorption of oxygen species.15,16 Dynamic surface reconstruction of catalysts during the OER process involves breaking intrinsic chemical bonds and in situ generation of high-valence metal hydroxide active phases, the “true catalytic species” for alkaline OER. In previous studies, Li et al. researched the transformation process of NiOOH in Ni(Fe) MOF.17 Hao's group demonstrated that Ni(Fe)OOH species are the genuine active species for FeOOH-modified NiMoO4 electrodes.18 While significant progress has been made in designing heterostructures for the OER, achieving dynamic surface reconstruction at low overpotentials to form catalytically active phases with strong interfacial electronic coupling remains an outstanding challenge.19–21
Herein, we report a self-supported, pearl-thread-like CoS2–CuxS heterostructure grown on copper foam (CoS2–CuxS/CF) that enables dynamic surface reconstruction under OER conditions. In situ Raman spectroscopy reveals the low-potential formation of a CoOOH–CuO–CuxS active interface through partial oxidation of CoS2 and CuxS. Operando electrochemical impedance spectroscopy confirms enhanced charge transfer kinetics, and density functional theory (DFT) calculations show that the in situ generated interface optimizes the electronic structure, facilitating stronger *OOH adsorption and lowering the free energy barrier of the rate-determining step. Benefiting from the synergistic effects of dynamic reconstruction and interfacial coupling, the CoS2–CuxS/CF catalyst exhibits outstanding OER performance with a low overpotential of 239 mV at 10 mA cm−2 and excellent operational stability for over 200 h in alkaline media. This study highlights the importance of interface design in promoting dynamic surface reconstruction and provides new insights into the rational design of high-performance, non-noble-metal OER catalysts.
The morphology of CoS2–CuxS/CF was characterized by scanning electron microscopy (SEM). As shown in Fig. S3a, the untreated copper foam (CF) exhibits a smooth surface. After hydrothermal treatment, CoS2–CuxS/CF displays a uniformly distributed pearl-thread-like structure across the CF surface (Fig. 1c). Compared to the SEM image of CuxS/CF (Fig. S3b), introducing CoS2 significantly improves the morphology, transforming it from a fragmented granular structure into an ordered pearly thread-like structure. This architecture provides a larger specific surface area and enhances the contact between the catalyst surface and the electrolyte, thereby promoting OER activity.28 To further investigate the morphological evolution during synthesis, we monitored the catalyst structure as a function of hydrothermal reaction time (Fig. S4). Initially, only a few thread-like features were observed within 2 h. With a prolonged reaction time, these structures gradually developed into more ordered pearl-thread-like formations. However, after 4 h, excessive growth led to the appearance of aggregated granular structures that covered the underlying framework. These observations suggest that insufficient and excessive hydrothermal durations are detrimental to the formation of the optimal morphology, limiting active site exposure and hindering the adsorption of reactive oxygen species.29
Transmission electron microscopy (TEM) and high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) were further employed to visualize the structure in greater detail. As shown in Fig. 1d, TEM confirms the distinctive pearl-thread morphology observed in SEM. High-resolution TEM (HR-TEM) images (Fig. 1e) reveal dual lattice fringes with spacings of 0.247, 0.291, and 0.265 nm, corresponding to the (210) plane of CoS2, the (112) plane of Cu2S, and the (107) plane of CuS, respectively (Fig. 1f). Selected area electron diffraction (SAED) patterns (Fig. 1g) further corroborate these assignments, indicating a polycrystalline structure. Elemental analysis through energy-dispersive X-ray spectroscopy (EDX) (Fig. S5) and energy-dispersive spectroscopy (EDS) elemental mapping (Fig. 1h) reveals a homogeneous distribution of Co, Cu, and S throughout the structure. Together, these comprehensive characterization studies confirm the successful construction of the CoS2–CuxS/CF catalyst.
The surface elemental composition and valence states are explored by X-ray photoelectron spectroscopy (XPS). The high-resolution C 1s spectrum of CoS2–CuxS/CF is employed as a calibration standard for other elements (Fig. S6). The peaks at around 284, 284.8, 286, and 288.4 eV are assigned to C
C, C–C, C–O, and C
O, respectively.30 The high-resolution Cu 2p spectrum demonstrated two peaks centered at 932.6 eV and 933.6 eV, assigned to Cu+ and Cu2+ in CoS2–CuxS/CF (Fig. 2a).31 Compared with CuxS/CF, the Cu+ peak in CoS2–CuxS/CF shifts 0.5 eV towards lower binding energy, accompanied by an increased Cu+/Cu2+ ratio. This indicates that CoS2 alters the local electronic environment of CuxS, facilitating the conversion of Cu+ into Cu2+ and accelerating the generation of active intermediates. Given the close binding energies of Cu0 and Cu+, Cu LMM Auger spectra were also collected (inset of Fig. 2a).32 A peak at 569.2 eV corresponds to Cu+, while a peak at 566.5 eV is attributed to Cu0, confirming the oxidation states.33
Within the S 2p spectral range for CoS2–CuxS/CF (Fig. 2b), the characteristic peaks at 161.6 and 162.3 eV are assigned to S 2p3/2 and S 2p1/2, with essentially no changes in the electronic environment.34 For the spectra of Co 2p, the peak at 778.4 eV corresponds to Co–S, while 781.5 and 784.2 eV are ascribed to Co2+ and satellite peaks, demonstrating that the cobalt is mainly present in the form of Co2+ for CoS2–CuxS/CF (Fig. 2c).35,36 Based on the comprehensive analysis of XPS results, the valence shift of Cu and the retention of S valence indicate electron transfer between CoS2 and CuxS, which improves the metal coordination environment, thus enhancing catalytic efficiency. Besides, we performed contact angle tests to investigate the impact of interfacial interactions on water adhesion. As shown in Fig. 2d, CoS2–CuxS/CF exhibits a contact angle of 29.2°, much smaller than the 56.4° of CuxS/CF, meaning more intimate contact between the catalytic surface and electrolyte for the CoS2–CuxS heterostructure, which modifies the catalyst surface wettability, promoting the transportation of the catalyst to the electrolyte.37
The OER performance of CoS2–CuxS/CF was evaluated in 1.0 M KOH using a standard three-electrode system. As shown in Fig. 3a, CoS2–CuxS/CF exhibits outstanding activity, requiring an overpotential of only 239 mV at 10 mA cm−2 and 315 mV at 100 mA cm−2. Compared with CoS2–Co3S4/CC and CuxS/CF, CoS2–CuxS/CF demonstrates superior catalytic performance, surpassing even RuO2/CF (η = 259 mV at 10 mA cm−2) under alkaline conditions. Notably, a high current density of 500 mA cm−2 is achieved with a relatively low overpotential of 390 mV, demonstrating the potential of CoS2–CuxS/CF for high-current-density applications. Tafel plots (Fig. 3b) highlight the favorable reaction kinetics of CoS2–CuxS/CF, with a Tafel slope of 60.3 mV dec−1, significantly lower than those of CoS2–Co3S4/CC (71.7 mV dec−1) and CuxS/CF (96.7 mV dec−1), indicating accelerated charge transfer and reduced energy loss during the OER.38,39 The catalytic performance of CoS2–CuxS/CF is influenced by synthetic parameters such as the amount of Co(NO3)2, hydrothermal time, and temperature. Optimization studies (Fig. S7) reveal that using 1.5 mmol Co(NO3)2 achieves the best performance, with faster reaction kinetics and improved charge transfer (Fig. S7). Further investigations (Fig. S8 and S9) show that a hydrothermal time of 4 h and a temperature of 150 °C yield the optimal morphology and catalytic activity. The enhanced performance is attributed to the synergistic effects between CoS2 and CuxS, which improve interfacial electron transfer and facilitate dynamic surface reconstruction by reducing charge transfer resistance (Rct). Additional cyclic voltammetry (CV) analyses supporting these observations are shown in Fig S10–S12. As summarized in Fig. 3c and Table S2, CoS2–CuxS/CF ranks among the best-performing sulfide and copper-based OER catalysts reported. The double-layer capacitance (Cdl) obtained from CV measurements (Fig. 3d) indicates the electrochemical surface area.40 The Cdl for CoS2–CuxS/CF is 93.2 mF cm−2, surpassing the comparison samples by several times, such as CuxS/CF (28.6 mF cm−2), CoS2–Co3S4/CC (20.4 mF cm−2), and CF (19.7 mF cm−2). The electrochemical surface area (ECSA) is determined using Cdl (Fig. 3e). A specific capacitance of 40 mF cm−2 is selected under normal circumstances.41 The electrochemical surface area (ECSA) is calculated to be 1560 cm2 for CoS2–CuxS/CF, nearly five times larger than that of CuxS/CF (340 cm2) (Fig. 3e). The pearl-thread-like morphology is responsible for the expanded surface area, increasing active site exposure and enhancing catalyst–electrolyte interactions.42 The larger specific surface increases the exposure of the interface in solution, which achieves higher density interfaces and facilitates the bonding of the active site. In order to demonstrate the intrinsic activity of CoS2–CuxS/CF more efficiently, ECSA-normalized LSV curves are displayed in Fig. 3e as well, which prove that CoS2–CuxS/CF has the highest electrocatalytic performance and intrinsic activity for the OER in an alkaline environment. The turnover frequency (TOF) is derived from the results of inductively coupled plasma mass spectrometry (ICP-MS), invariably correlating with the intrinsic activity of electrocatalysts (Table S1).43 As depicted in Fig. 3f, the TOF of CoS2–CuxS/CF exhibits the most rapid increasing trend with rising applied voltage. When the applied voltage reaches 259 mV, the TOF value of CoS2–CuxS/CF is 0.03 s−1, further demonstrating that the formation of the CoS2–CuxS interface improves the intrinsic activity. The faradaic efficiency (FE) of the OER process is further calculated, representing the ratio of actual gas evolved to the theoretical amount (Fig. 3g).44 As a result, actual oxygen production and theoretical oxygen production exhibit an extremely high degree of fit. More than 10 mL of O2 is generated by 1 cm2 of CoS2–CuxS/CF within 30 min, which is close to the theoretical oxygen production. In particular, the high activity of CoS2–CuxS/CF can be maintained for over 200 h at a current density of 100 mA cm−2 (Fig. 3h), implying the extraordinary cycling stability of CoS2–CuxS/CF.
The surface evolution of CoS2–CuxS/CF during the OER process was investigated using transmission electron microscopy (TEM) after long-term stability testing. As shown in Fig. S13 and 14, the reaction process induced significant alterations in the microscopic morphology of the catalyst surface. Additionally, high-angle annular dark-field scanning TEM (HAADF-STEM) combined with energy-dispersive X-ray (EDX) mapping revealed the uniform distribution of Cu, Co, O, and S throughout the structure (Fig. S15 and S16), further validating the catalyst's composition and architecture after the reaction. Notably, the oxygen content increased significantly after the OER process, accompanied by a corresponding decrease in sulfur content, implying the partial transformation of sulfides into oxides. TEM analysis revealed a lattice fringe spacing of 0.186 nm, corresponding to the (202) plane of CuO, confirming the formation of CuO (Fig. 4a and b). Additionally, a spacing of 0.230 nm is assigned to the (006) plane of CoOOH, indicating the generation of CoOOH during the OER. These findings are consistent with post-reaction XPS analysis (Fig. 4c), where an increase in the Cu2+ signal corroborates the formation of CuO. In addition, residual sulfur signals were still detected by XPS (Fig. 4d), which may have come from residual SO42−. To verify this, a residual SO42− detection experiment was conducted (Fig. S17). The reacted catalyst was placed in deionized water and subjected to 12 h of ultrasonic treatment. Then, the solution was acidified with dilute hydrochloric acid, and BaCl2 solution was added. The result showed that a small amount of white precipitate appeared, indicating the presence of residual SO42− on the surface of the catalyst after the reaction.45
In situ Raman spectroscopy was further employed to monitor structural evolution during the OER. Compared to the dry state, Raman scattering peaks at 390 cm−1 weakened after immersion in alkaline electrolyte due to environmental changes. In addition, copper foam (CF) oxidizes readily upon exposure to water, leading to observable Cu–O vibrations at open-circuit potential (OCP). As shown in Fig. 4e, characteristic Raman peaks at 270 cm−1, 348 cm−1, and 550 cm−1 correspond to Cu–O, Cu–S, and CoOOH, respectively.25,26,46 With increasing applied potential, the intensity of the Cu–S peak gradually decreases, while the Cu–O peak intensifies, indicating progressive oxidation of CuxS. Importantly, CoOOH formation becomes significant only after substantial CuO generation, suggesting that the disruption of the Cu–S bond and the formation of Cu–O bonds redistribute the local charge density, thereby facilitating the activation of Co sites into highly active CoOOH species. As shown in Fig. S18, the Raman peaks at 470 and 527 cm−1 belong to Co–S.47 As the applied voltage increases, the signal corresponding to Co–S disappears. When the voltage rises to 1.5 V, a characteristic peak attributed to CoOOH begins to emerge. The overpotential required for the surface reconstruction of the single CoS2 interface is much higher than that of the CoS2–CuxS interface, which indicates that the CoS2–CuxS heterostructure promotes the adsorption of OH− onto active intermediates, effectively improving the reaction kinetics. Combined analysis of these results confirms that a CoOOH–CuO–CuxS heterointerface forms during the OER process. CoS2 and CuxS are partially transformed into CoOOH and CuO, respectively. This hybrid structure benefits from the high conductivity of residual sulfides, lowering charge transfer resistance and enhancing reaction kinetics. Notably, in situ Raman spectroscopy showed that the Cu–S bond did not disappear even when the applied voltage reached 1.7 V, indicating the presence of residual CuSx. In summary, the in situ generated CoOOH–CuO–CuxS interface replaces the original sulfide phase, significantly boosting catalytic activity.48
Electrochemical impedance spectroscopy (EIS) was performed to further investigate the dynamic evolution of active species on the catalyst surface.49In situ EIS measurements are conducted at various potentials to further elucidate the role of Co sites in promoting the adsorption and desorption of oxygen-containing reactants.50Fig. 4f and g and present the Nyquist plots of CoS2–CuxS/CF and CuxS/CF at different potentials. As the potential increased, the charge transfer resistance (Rct) decreased significantly, indicating a notable acceleration in reaction kinetics.51 Moreover, the charge transfer rate of CoS2–CuxS/CF is consistently higher than that of CuxS/CF, which can be attributed to the dynamic evolution of Co sites at the interfaces of the CuxS and CoS2 species. The relevant bond plots are presented in Fig. S19. As the applied voltage increases, the position of the phase angle changes continuously, which reveals that the phase angles of CoS2–CuSx are smaller than those of CuSx at each identical applied potential. This indicates that more electrons participate in OER catalysis, thereby enhancing the kinetics of the OER. Fig. 4h illustrates the decreasing trend of Rct between CoS2–CuxS/CF and CuxS/CF. Under low-potential conditions, the Co active sites exhibit enhanced adsorption of oxygen-containing intermediates. These intermediates are rapidly accumulated upon the introduction of the Co sites, activating the OER performance of catalysts at low potentials.52
To elucidate electronic interactions of key reaction intermediates at the interface, we modeled the system using density functional theory (DFT).53 To align with system characterization, CoOOH–CuO–CuxS is selected as the computational model, with CoOOH serving as the contrast sample. Fig. 5a shows the optimized model configuration. The differential charge density is analyzed in Fig. 5b, with the yellow region indicating electron accumulation and the green region representing electron depletion. This suggests that there is localized charge accumulation near CoOOH and CuO regions. Significantly, charge aggregation at Cu sites is markedly weaker than that at Co sites, implying that CuO regulates interfacial electron transfer.54 This electronic coupling is further clarified in the two-dimensional charge density projection (Fig. 5c), where the blue region represents electron accumulation, and the red region indicates electron depletion. Density of states (DOS) calculations (Fig. 5d) reveal enhanced electron density near the Fermi level for CoOOH–CuO–CuxS compared to CoOOH, demonstrating that CoOOH–CuO interactions promote interfacial electron mobility and improve material conductivity.55 The d-band center shifts from −1.27 eV (CoOOH) to −1.17 eV (CoOOH–CuO–CuxS), indicating higher antibonding orbital occupancy above the Fermi level and reduced bonding orbital occupancy below the Fermi level (Fig. 5e), which is consistent with weakened Co–O bonds that favor oxygen species desorption.56Fig. 5f illustrates the four-electron OER pathway on CoOOH–CuO–CuxS, where Co sites remain the primary adsorption centers. Hydroxide adsorption forms *OH intermediates on Co sites, followed by sequential deprotonation to *O and *OOH species, culminating in O2 release.57 Gibbs free energy (ΔG) analysis (Fig. 5g) identifies the transition from *O to *OOH (ΔG3) as the rate-determining step. CoOOH–CuO–CuxS exhibits a significantly lower ΔG3 (1.48 eV) compared to CoOOH (1.79 eV), which reduces energy barriers of *O to *OOH, further facilitating the formation of CoOOH. In summary, compared to single CoOOH, the in situ generated active CoOOH–CuO–CuxS interfaces optimize the electronic structure and facilitate the desorption of oxygen-containing intermediates *O during the OER process, resulting in excellent OER performance.
The excellent OER performance motivates us to apply the catalyst in overall water splitting. Fig. S20a shows the model of the overall water splitting device. Electrochemical analyses revealed that the CoS2–CuxS/CF(+)||Pt/C(−) system demonstrated superior overall water splitting performance relative to the control in 1.0 M KOH. As shown in the polarization curve (Fig. S20b), the CoS2–CuxS/CF(+)||Pt/C(−) electrolyzer required only 1.62 V to achieve a current density of 100 mA cm−2, substantially lower than the 1.82 V required for the RuO2(+)||Pt/C(−) system. Furthermore, the CoS2–CuxS/CF(+)||Pt/C(−) system exhibited competitive performance at 100 mA cm−2 compared with numerous reported electrocatalysts (Fig. S20c and Table S3). As demonstrated in Fig. S20d, the CoS2–CuxS/CF(+)||Pt/C(−) system exhibits outstanding stability at a current density of 100 mA cm−2 for over 60 h, indicating its certain potential for industrial applications.
| This journal is © the Partner Organisations 2026 |