Kai Zeng†
*acd,
Ming Chao†e,
Lanxiang Huang
b,
Hongwei Tao*a,
Zhengyou Hea,
Yibing Li*a and
Zhihui Sun*f
aSchool of Electrical Engineering, Institute of Smart City and Intelligent Transportation, Southwest Jiaotong University, Chengdu 610032, China. E-mail: zengkai@swjtu.edu.cn; hwtao@swjtu.edu.cn; yibing.li@swjtu.edu.cn
bLeshan West Silicon Materials Photovoltaic New Energy Industry Technology Research Institute, Leshan Normal University, Leshan 614000, China
cHuzhou Key Laboratory of Smart and Clean Energy, Yangtze Delta Region Institute, University of Electronic Science and Technology of China, Huzhou 313001, China
dSchool of Chemistry and Chemical Engineering, Nanjing University, Nanjing 210023, China
eCollege of Chemistry and Molecular Engineering, Peking University, Beijing 100083, China
fSchool of Mines, China University of Mining and Technology, Xuzhou 221116, China. E-mail: zhsun@cumt.edu.cn
First published on 27th August 2025
The synergistic integration of high-valence metal sites and interfacial bridge-oxygen bonding plays a pivotal role in the construction of valid bifunctional electrocatalysts for accelerating oxygen electrode redox kinetics and promoting the practical implementation of lithium–oxygen batteries. Herein, high-valence tungsten dopants induced the formation of 2H-MoSe2, which was successfully anchored on the layered Ti3C2 MXene matrix (W-MoSe2@MXene) with an interfacial bridge-oxygen bonding structure, affording a porous and vertical staggered nanosheet array network architecture. Consequently, the Li–O2 battery assembled with the as-prepared W-MoSe2@MXene cathode delivers high discharge specific capacity (12442.6 mA h g−1) and favorable cycling lifespan (over 194 cycles) at 1 A g−1. Notably, stable operation is also maintained over 61 cycles at an ultra-high current density of 5 A g−1. Experimental analysis in combination with density functional theory (DFT) calculation reveals that the synergistic interaction between high-valence W dopants and bridge-oxygen bonds facilitates spatial charge redistribution and accelerates charge transfer, thereby lowering the theoretical discharge–charge overpotential and enhancing electrode reaction kinetics. This work offers a feasible design paradigm for the construction of MXene-based oxygen electrode catalysts toward high-rate Li–O2 batteries.
Two-dimensional (2D) materials, such as transition metal chalcogenides and MXenes, have emerged as a vibrant class of nanomaterials owing to their high aspect ratios, tunable surface chemistries and remarkable mechanical flexibility.14–16 Nevertheless, their practical implementation is often compromised by pronounced self-restacking behavior and intrinsic limitations to individual material systems. Constructing 2D/2D heterostructures is pivotal for circumventing the inherent constraints of single-component systems. In particular, interfacial bridge-oxygen bonds with M1–O–M2 configuration have recently attracted growing attention due to their superior electrical conductivity and structural stability, which can effectively modulate Li2O2 formation/decomposition pathways and promote redox reaction kinetics.17,18 For example, a 2D/2D MoS2/Ti3C2 MXene heterostructure with an interconnected network was successfully fabricated, leveraging the synergistic effect of high-capacitance MoS2 and high-rate Ti3C2 MXene to deliver excellent electrochemical performance in a flexible capacitance device.19 Similarly, ample Ir–O–Mn bonds were successfully incorporated into an Ir/MnOx catalyst, serving as effective charge-transfer channels between IrOx clusters and the MnOx matrix.20 Sun et al. reported interfacial oxygen bridge engineering between MoOx and an MXene matrix (MoOx@Ti3C2 MXene), which functioned as a promising cathode catalyst for Li–O2 batteries. Consequently, the Li–O2 battery assembled with the MoOx@Ti3C2 MXene catalyst exhibited favorable discharge–charge polarization (0.75 V) and remarkable cycling stability (over 300 cycles). Density functional theory (DFT) analysis revealed that the bridge-oxygen bonding maximizes the advantages of electronic structure manipulation and Li2O2 formation.21
The high-valence metal doping (e.g. Zr4+, Nb5+, and Mo6+) strategy can effectively tune local electron density and induce spatial charge redistribution, while simultaneously acting as active sites for the oxygen electrode reaction and accelerating the reaction kinetics via optimizing intermediate adsorption and lowering energy barriers.22,23 Zeng and co-workers reported that the incorporation of high-valence metal sites (e.g., Zr4+ and Mo6+) generates sufficient unfilled antibonding orbitals through optimized d–p orbital hybridization between metal and non-metal atoms. This electronic configuration is conducive to diminish charge transfer barriers and lower the free energy of the rate-determining step (O* → OOH*).24 Li et al. designed a conductive nickel catecholate framework with strengthened Ni–O covalency and high-valence Ni3+ species (NiIII-NCF), which facilitated electron transport and favored the formation of nanosheet-like Li2O2, achieving a superior rate capability and cycling stability.25 Consequently, the synergistic modulation of high-valence metal centers and interfacial oxygen bridge bonds is poised to reshape the electronic structure and local coordination environment. However, the correlation and underlying cooperative interaction mechanism between the specific high-valence metals and interfacial bridge-oxygen bonding are yet to be unveiled, particularly in 2D/2D metal selenides and MXene heterostructures.
Inspired by the aforementioned considerations, high-valence W-doped MoSe2 nanosheets anchored on the 2D layered Ti3C2 MXene matrix (W-MoSe2@MXene) with ample bridge-oxygen bonds, featuring a porous and vertical staggered nanosheet array network architecture, is successfully constructed and acts as a favorable cathode for Li–O2 batteries. As a result, the Li–O2 battery assembled with the W-MoSe2@MXene cathode possesses a promising specific capacity of 12442.6 mA h g−1 and excellent cycling lifespan (over 194 cycles) at 1 A g−1, which is superior to that of MoSe2@MXene (5568.3 mA h g−1, 104 cycles) and MXene (4823.2 mA h g−1, 67 cycles). Notably, the W-MoSe2@MXene-based cathode can even stably cycle over 108 and 61 cycles at ultra-high current densities of 3 and 5 A g−1, respectively. DFT calculation results reveal that the synergistic interaction between high-valence metal sites and bridge-oxygen bonds is beneficial to optimize the adsorption/desorption of Li2O2 and accelerate the charge transfer kinetics, thereby lowering the theorical overpotentials of the discharge and charge process.
Scanning electron microscopy/transmission electron microscopy (SEM/TEM) were employed to investigate the morphology and microstructure of the as-prepared catalysts. As shown in Fig. 2a and b, SEM images reveal the interconnected and vertical W-MoSe2 nanosheets randomly distributed on the surface of Ti3C2 MXene, which is conducive to forming a porous architecture. TEM observations further confirm the three-dimensional porous and interconnected nanosheet array structure in W-MoSe2@MXene (Fig. 2c and d), which is favourable for facilitating mass diffusion and charge transport. High-resolution TEM (HRTEM) images show a distinct multi-phase heterostructure in W-MoSe2@MXene, implying the existence of sufficient heterointerfaces (Fig. 2e and S2). Additionally, the characteristic lattice spacings of 0.64 and 0.26 nm are observed, corresponding to the (002) lattice plane of 2H-MoSe2 and the (002) lattice plane of Ti3C2 MXene (Fig. 2f–h), respectively. Notably, the hexagonal lattice region can be visualized as depicted in Fig. 2i, proving the formation of 2H phase MoSe2 within the W-MoSe2@MXene heterostructure. High-angle annular dark field (HAADF) images and scanning TEM (STEM) mapping demonstrate the homogeneous distribution of Mo, W, Se, Ti and C elements throughout the W-MoSe2@MXene matrix (Fig. 2j and k).
![]() | ||
Fig. 2 (a and b) SEM images, (c and d) TEM images, (e–i) HRTEM images, (j) HAADF-STEM images and (k) HAADF-STEM image and EDS mapping of W-MoSe2@MXene. |
As a counterpart, the MoSe2@MXene catalyst was fabricated by the same method without the addition of the W dopant precursor. The thus-prepared sample shows a similar three-dimensional porous structure composed of interconnected nanosheet arrays (Fig. S3 and S4). As displayed in Fig. S5, a heterophase hybrid structure is observed in MoSe2@MXene. HRTEM analysis demonstrates interplanar spacings of 0.82 and 0.26 nm, corresponding to the (002) lattice plane of 1T/2H-MoSe2 and (002) lattice plane of Ti3C2 MXene,27,28 respectively. Notably, W doping promotes the phase transformation from the metastable 1T/2H structure to the thermodynamically favorable 2H phase, thereby enhancing the structural stability of W-MoSe2@MXene under Li–O2 battery operating conditions. Energy dispersive spectroscopy (EDS) mapping manifests that the Mo, Se, Ti and C are homogeneously distributed in MoSe2@MXene (Fig. S6).
The crystal structure of MXene, MoSe2@MXene and W-MoSe2@MXene was assessed by X-ray diffraction (XRD) as shown in Fig. 3a. For pristine MXene, the diffraction peaks at 8.9°, 18.6°, 27.9°, 36.3°, 42.1° and 62.7°, are assigned to the (002), (004), (008), (111), (200) and (220) planes of typical Ti3C2 MXene,29 respectively. Additionally, MoSe2@MXene demonstrates four characteristic diffraction peaks at 12.3°, 32.4°, 38.2° and 56.3°, indexed to the (002), (100), (103) and (110) planes of MoSe2 (JCPDS#29-0914), respectively. W-MoSe2@MXene possesses identical diffraction peaks to MoSe2@MXene with a slight negative shift, indicating lattice expansion due to successful W incorporation rather than secondary phase formation. This result further proves that the W-doped MoSe2 was successfully anchored on the layered MXene.
The surface chemical composition and valence states of MoSe2@MXene and W-MoSe2@MXene samples were probed by X-ray photoelectron spectroscopy (XPS). The full XPS spectra of W-MoSe2@MXene catalysts display the existence of W, Mo, Se, Ti, and C elements (Fig. S7). The high-resolution Mo 3d XPS spectra of MoSe2@MXene and W-MoSe2@MXene can be deconvoluted into two paired peaks (Fig. 3b), related to the Mo4+ (228.7 and 231.8 eV) and Mo6+ (229.5 and 232.2 eV),30,31 respectively. Remarkably, the Mo 3d orbital binding energy of W-MoSe2@MXene shifted positively by around 0.3 eV after the incorporation of the W dopant, indicating that W would capture the electron cloud redistribution around Mo atoms and induce spatial charge redistribution, which is mainly attributed to the electronegativity of W (2.36) being relatively higher than that of Mo (2.16).32,33 In Se 3d XPS spectra (Fig. 3c), the two characteristic peaks at 54.2 and 55.1 eV are in good agreement with the Se 3d5/2 and Se 3d3/2, respectively.34 In the W 4f XPS spectrum of W-MoSe2@MXene (Fig. 3c), the peaks fitted at 37.8 and 35.9 eV can be attributed to the W 4f7/2 and W 4f5/2 of W6+ species (Fig. 3d), respectively, indicating the incorporation of high-valence W sites.35 As shown in Fig. 3e, the Ti 2p XPS spectra demonstrate three characteristic peaks at 453.9, 458.7 and 464.5 eV, corresponding to the Ti–C, Ti–O 2p3/2 and Ti–O 2p1/2,36 respectively. In the O 1s XPS spectra (Fig. 3f), the four peaks at 529.8, 530.8, 531.5 and 533.2 eV can be attributed to Ti–O, Ti–OH, Mo–O–Ti and surface adsorbed H2O,37,38 respectively. Notably, the presence of the Mo–O–Ti bond indicates the successful formation of interfacial metal–oxygen bridge sites between W-MoSe2 and MXene. This heterointerface is expected to boost interfacial charge transfer and lower the polarization overpotential, thereby improving the electrochemical performance of Li–O2 batteries.
The catalytic performance of the W-MoSe2@MXene cathode was evaluated in comparison with its MXene and MoSe2@MXene counterparts via the assembled coin type Li–O2 batteries. The cyclic voltammetry curves of samples are shown in Fig. 4a, where W-MoSe2@MXene possesses higher oxygen reduction reaction (ORR) onset potential (2.89 V) and lower oxygen evolution reaction (OER) onset potential (3.61 V) than those of MXene and MoSe2@MXene, indicative of superior bifunctional catalytic performance. Additionally, the larger integrated areas of the cathodic and anodic peaks imply a better specific capacity in the W-MoSe2@MXene-based system. Galvanostatic discharge/charge curves demonstrate that W-MoSe2@MXene exhibits a favorable discharge and charge overpotential of 0.61 V at 100 mA g−1 (Fig. 4b), which is markedly lower than that of MoSe2@MXene (0.87 V) and MXene (1.11 V). W-MoSe2@MXene manifests the smallest charge transfer resistance, indicative of rapid charge transport enabled by the synergistic effects of interfacial bridge-oxygen bonds and high-valence W doping (Fig. 4c). The rate performance of samples was also evaluated at a fixed specific capacity of 1000 mA h g−1. As the current density increases, W-MoSe2@MXene exhibits a significant decrease charge potential plateau than other counterparts, which is beneficial to induce a relatively low overpotential (Fig. 4d). Additionally, the W-MoSe2@MXene cathode assembled Li–O2 battery exhibits a favorable specific capacity of 12442.6 mA h g−1 at 1 A g−1, which is ∼2.23 and ∼2.56 times greater than that of MoSe2@MXene (5568.3 mA h g−1) and MXene (4823.2 mA h g−1), respectively.
Meanwhile, to evaluate the high-rate performance of the thus-prepared catalyst, the assembled Li–O2 battery with the W-MoSe2@MXene cathode was evaluated at high current densities of 1–5 A g−1. As shown in Fig. 4f, the charge potential plateau increases only slightly from 4.12 to 4.31 V and eventually returns to 4.18 V, implying an advantageous high-rate performance. Specifically, it can run continuously for 127 cycles at 1 A g−1. Additionally, the W-MoSe2@MXene cathode assembled Li–O2 battery at high current densities retains a stable discharge–charge overpotential from ∼0.95 V (1 A g−1) to ∼1.32 V (5 A g−1) and returns to ∼0.98 V (1 A g−1). However, other counterparts display relatively higher polarization overpotentials at ultra-high rates (Fig. 5a), implying the potential application of the as-prepared W-MoSe2@MXene cathode in high-rate Li–O2 batteries. The discharge–charge cycling life is a crucial parameter to assess the catalytic activity of Li–O2 batteries. As depicted in Fig. 5b and S9, the W-MoSe2@MXene-based Li–O2 battery can effectively operate over 194 cycles with a fixed specific capacity of 1000 mA h g−1 at a high current density of 1 A g−1, which is evidently superior to MoSe2@MXene (104 cycles) and MXene (67 cycles) (Fig. S10 and S11). Surprisingly, W-MoSe2@MXene can even stably cycle over 61 cycles at an ultra-high current density of 5 A g−1, indicating a favorable high current density catalytic activity.
To elucidate the nature of discharge products on the W-MoSe2@MXene cathode, SEM and XPS characterization studies as well as electrochemical analysis were performed on the electrodes after discharge and subsequent recharge. As shown in Fig. 5c, the surface of the W-MoSe2@MXene cathode is uniformly covered by densely accumulated discharge products, which is attributed to the three-dimensional porous architecture and interconnected nanosheet array of the W-MoSe2@MXene providing abundant nucleation sites and spatial accommodation for Li2O2 deposition. This structural configuration facilitates efficient Li2O2 decomposition into Li+ and O2, and enhances electron conductivity as well as supports higher specific capacities. Upon recharging (Fig. 4d), the discharge products are completely removed and the initial three-dimensional framework of the W-MoSe2@MXene cathode is reinstated, exposing active sites for subsequent electrochemical reactions. Fig. 4e presents the high-resolution XPS spectra of Li 1s and Se 3d obtained from the cathode after the initial discharge. The emergence of a new peak centered at 54.2 eV is in good agreement with the formation of Li2O2.39,40 Notably, the complete disappearance of the Li2O2 signal indicates full decomposition of the discharge product, underscoring the excellent reversibility of the electrochemical reaction (Fig. 4f). Additionally, high-resolution XPS spectra of Mo 3d, W 4f and Ti 2p demonstrate that the peaks after the first discharge and subsequent recharge are basically the same as in the pristine state, indicating the preservation of the chemical environment and confirming the excellent structural stability of the W-MoSe2@MXene cathode (Fig. S12 and S13). This robust framework is beneficial to the long-term cycling stability of the MXene-based Li–O2 battery. Electrochemical impedance spectroscopy (EIS) measurements (Fig. 5g) along with equivalent circuit modeling (Fig. S14) further elucidate the electrochemical changes during cycling.41,42 The substantial increase in charge transfer resistance (Rct) observed after the discharge process reflects the obstruction of ion diffusion pathways by poorly conductive and insoluble products. The Rct returns to values comparable to those of the pristine W-MoSe2@MXene cathode after the recharging process, indicating effective decomposition of the discharge products. The high donor number of DMSO (∼29.8 kcal mol−1) promotes a solution-mediated growth pathway of Li2O2. Strong Li+ solvation by DMSO stabilizes soluble Li+–O2− ion pairs, enabling their diffusion and disproportionation, ultimately producing the toroidal Li2O2 particles (Fig. 5c), which is beneficial to achieve high discharge capacities, as it is less passivating compared with thin-films Li2O2 typically observed in low donor number solvents.43 Nonetheless, a potential challenge associated with toroidal Li2O2 is its incomplete oxidation during charging, which underscores the importance of W-MoSe2@MXene catalyst design in mitigating the resulting high overpotentials. Therefore, the synergistic effect arising from the incorporation of high-valence W6+ and oxygen-bridge sites as well as the distinctive porous architecture facilitates the efficient decomposition of the in situ formed Li2O2 discharge product during recharging (Fig. 5h), thereby accelerating reaction kinetics and enhancing cycling performance.
To clarify the positive role of interfacial oxygen-bridge bonding and high-valence metal sites in the electronic structure and reaction free energy barrier, first-principles calculations based on the density functional theory (DFT) were performed. Density of states (DOS) analyses reveal that both W-MoSe2@MXene and MoSe2@MXene tend to possess the metallic conductor behavior with a continuous occupation of electrons across the Fermi level, indicative of a favorable intrinsic electronic conductivity (Fig. 6a and b). Projected DOS results further indicate that the pronounced electronic states near the Fermi level predominantly originate from Ti atoms, owing to the excellent electronic conductivity of the MXene component. Accordingly, charge density difference calculations show the apparent charge transfer between W-MoSe2 (MoSe2) and MXene in W-MoSe2@MXene (MoSe2@MXene) induced by the construction of the heterostructure with the favorable interfacial bridge-oxygen bonding (Fig. 6c and d), which is beneficial to accelerate the interface charge transfer kinetics at the catalyst–electrolyte interface.
To probe the origin of the catalytic activity and underlying reaction pathways, the adsorption energetics of relevant intermediates and products were investigated. Voltage gaps were calculated to quantitatively assess the overpotentials by using the definitions UORR = Ueq − UDC and UOER = UC − Ueq for the discharge (ORR) and charge (OER) process.44,45 Optimized structure models of key intermediates/products on W-MoSe2@MXene and MoSe2@MXene are depicted in Fig. 6g, S15 and S16, respectively. As shown in Fig. 6e, the MoSe2@MXene catalyst exhibits substantial ORR and OER theoretical overpotentials of 2.06 and 2.39 V, respectively, reflecting its limited catalytic activity. The as-prepared W-MoSe2@MXene catalyst possesses significantly lower theoretical overpotentials of 1.75 V and 1.91 V for the ORR and OER (Fig. 6f), respectively, confirming that W doping effectively enhances bifunctional oxygen electrocatalysis. The significantly reduced voltage gaps induced by the synergistic effect between high-valence W sites and interfacial bridge-oxygen bonding are expected to improve energy efficiency, suppress parasitic reactions and thereby ultimately prolong the cycling lifespan of lithium–oxygen batteries.
Footnote |
† K. Zeng and M. Chao contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |