Plasma-engineered sandwich-structured N-doped carbon@TiNb2O7 with vertical graphene skeletons for ultrahigh-rate and long-cycling lithium storage
Abstract
The rapid expansion and booming development of the lithium-ion battery market have raised escalating concerns over safety issues. Titanium niobium oxide (TiNb2O7, TNO) is a highly promising, safe anode material due to its intercalation reaction mechanism and high operating potential. However, its intrinsic low electronic conductivity severely hinders practical implementation. To address this, we developed a plasma-assisted interfacial engineering strategy to fabricate self-supported sandwich-structured N-doped carbon (N-C)@TNO composites. This unique “conductive skeleton || active core || protective shell” architecture comprises: (1) vertical graphene (VG) arrays acting as three-dimensional charge highways, (2) TNO nanoparticles (30-60 nm) serving as redox-active centers, and (3) uniform N-C shells (~3 nm). The synergistic coupling between the VG skeleton and the N-C coating establishes an all-around conductive network. The optimized N-C@TNO anode delivers exceptional rate capability (300.1 mAh g-1 at 0.2 C and 214.4 mAh g-1 at 40 C) and ultralong cycling stability (95.38% capacity retention after 5,000 cycles at 20 C), outperforming most reported TNO-based anodes. This work presents a novel concept for designing high-power storage electrodes, particularly multistage composite structures.
Keywords
INTRODUCTION
Lithium-ion batteries (LIBs) have expanded beyond portable electronics to power a diverse range of applications, including electric vehicles and grid-scale energy storage systems[1-3]. This rapid diversification has concurrently heightened demands for enhanced safety and ultrafast charge/discharge capabilities[4-6]. Among promising alternative anode materials, titanium niobium oxide (TiNb2O7, TNO), with a shear
To overcome these limitations, strategies including heteroatom doping, conductive compositing, and morphology engineering have been implemented[17-22]. Heteroatom doping enhances intrinsic electronic conductivity through band structure modulation, as demonstrated by Tian et al. in La-doped TNO, where doping induces localized energy levels via hybridization among O 2p, Ti 3d, and Nb 4d orbitals[23]. However, excessive doping may compromise structural integrity[24]. Conductive compositing effectively augments electronic conductivity without perturbing the host crystal structure[25]. For example, Wu et al. reported significantly enhanced electrical transport in MXene-coated TNO microspheres, and Yang et al. achieved comparable improvements using solvothermal synthesized carbon-coated porous TNO microspheres[26,27]. Ho et al. highlight the value of three-dimensional (3D) graphene macrostructures for building conductive, robust electrode frameworks, which supports our VG skeleton strategy[28]. In parallel, Folorunso et al. demonstrate that graphene coating serves as a universal strategy for interfacial engineering, enhancing cycling stability[29]. These conductive matrices establish extrinsic pathways, effectively mitigating limitations in electron transfer[30]. Critically, while both heteroatom doping and conductive compositing significantly enhance electronic conductivity, their capacity to improve ion diffusion kinetics remains constrained. Morphology engineering, particularly reducing active material particle size, shortens diffusion pathways for both ions and electrons, thereby enhancing kinetics[31]. Building on the work of Xu et al., who systematically investigated how the morphological dimensions (1D, 2D, 3D) of TNO anodes influence lithium-ion storage kinetics and identified the advantages of 2D structures for shorter Li+ diffusion paths[32]. Nonetheless, the concomitant increase in specific surface area amplifies susceptibility to extensive interfacial side reactions, potentially degrading coulombic efficiency and cycle life[33]. Consequently, developing TNO electrode materials with superior electrochemical performance requires a synergistic strategy that integrates morphological control with the design of conductive scaffolds[34-36].
In this study, we design a sandwich-structured N-doped carbon (N-C)@TNO integrated with vertical graphene (VG) skeletons. The fabrication involves three sequential steps: First, the 2D graphene nanosheets are deposited on carbon cloth vertically via plasma-enhanced chemical vapor deposition (PECVD) to form a 3D conductive scaffold. Subsequently, TNO nanoparticles were grown on VG nanosheets via a simple solvothermal process with the preserved 2D morphology [Supplementary Figure 1]. The porous structure of VG nanosheets provides TNO with faster electronic/ionic transport channels. Finally, an ultrathin N-C shell (~3 nm) is uniformly in situ coated on TNO via PECVD again. This synergistic configuration establishes rapid 3D electron transport channels while the N-C shell acts as a physical barrier to minimize the direct contact between the electrodes and the electrolyte to prolong the cycle life. The resultant self-supporting electrode delivers outstanding performance: 300.1 mAh g-1 at 0.2 C, 214.4 mAh g-1 at 40 C, and 95.38% capacity retention after 5,000 cycles at 20 C. Post-cycling Scanning Electron Microscope (SEM) confirms structural integrity following rigorous electrochemical testing. This work establishes a novel paradigm for designing high-rate, durable energy storage devices through the integration of morphological and conductive engineering.
EXPERIMENTAL
Materials synthesis
Preparation of VG arrays: The VG nanosheet array was constructed on carbon cloth substrates through a PECVD process. Specifically, the carbon cloth substrate was loaded into the PECVD furnace that had been preheated to 500 °C under a vacuum of approximately 5 Pa. A gas mixture containing 20 sccm Ar, 10 sccm H2, and 6 sccm CH4 was then introduced into the chamber, and the power was set to 500 W for a reaction time of 9 min to form the VG array.
Preparation of TNO arrays: The TNO composite was fabricated through a hydrothermal synthesis method. First, 0.71 g titanium isopropoxide (C12H28O4Ti, TIP; 99.9%, Macklin) and 1.35 g niobium chloride (NbCl5; 99%, Macklin) were dissolved in 120 mL of ethanol according to stoichiometric proportions. Then, the resulting solution was mixed with the pre-synthesized VG arrays and transferred into a Teflon-linked steel autoclave for hydrothermal treatment. The hydrothermal reaction was kept at 200 °C for 1 h. After being washed with ethanol and deionized water and dried at 60 °C, the sample was annealed in Ar at 800 °C for
Preparation of N-C@TNO arrays: The above TNO arrays acted as the skeleton for the growth of the N-C coating layer. First, the TNO arrays were loaded into the PECVD furnace that had been preheated to 600 °C under a vacuum of approximately 5 Pa. Ethylenediamine (EDA) was then used as N and C source by heating it to 40 °C to promote the volatilization rate. Following that, EDA gas was introduced; the flow of EDA vapor was regulated through a mechanical valve to maintain the chamber pressure at approximately 30 Pa. At this point, the power was set to 500 W for a reaction time of 6 min to form the N-C@TNO composite arrays.
Materials characterization
The microstructures and morphologies of the samples were examined using transmission electron microscopy (TEM)-high-resolution TEM (HRTEM) (HermoFisher Scientific, Talos F200X G2) and SEM (HermoFisher Scientific Apreo 2 S). The crystal and phase structures of all samples were confirmed using
Electrochemical measurements
Cyclic voltammetry (CV) measurements were conducted on a CHI 760E electrochemical workstation (CH Instruments Inc., Shanghai) with a potential window of 1.2-2.5 V at various scan rates ranging from
Electrode and cell preparation
The electrochemical measurements of electrodes were conducted using a CR2032 coin cell battery assembled in an Ar-filled glove box. The working electrode, counter electrode, and separator were prepared from samples, lithium foil, and microporous polypropylene membrane (Celgard, 2300), respectively. A
RESULTS AND DISCUSSION
Figure 1A illustrates the fabrication schemes of the sandwich electrode with a triple-layered “conductive skeleton || active core || protective shell” architecture. First, the VG arrays are deposited on carbon cloth via a facile PECVD method. Supplementary Figure 2A and B shows that after the deposition, a thin layer of interconnected VG nanosheets grows on the carbon fibers with a ~200-500 nm porous structure, which acts as the high-conductivity 3D scaffold. Then, TNO nanoparticles are uniformly anchored onto the VG skeleton to form a coral-like composite structure using the simple solvothermal method. Finally, an N-C shell layer was in situ capped on the TNO surface by PECVD using EDA as the carbon precursor. This process resulted in the formation of a uniform and continuous N-C protective layer on the TNO surface. SEM images clearly revealed the morphological evolution across synthesis stages. On the one hand,
Figure 1. Plasma-engineered synthesis and multi-scale characterization of the N-C@TiNb2O7 composite. (A) Schematic of the fabrication process; SEM images of (B) TNO array and (C) N-C@TNO array; HRTEM images of (D) TNO and (E) N-C@TNO; (F) EDX mapping of N-C@TNO showing elemental distribution. TNO: TiNb2O7, titanium niobium oxide; HRTEM: high-resolution transmission electron microscopy; EDX: energy-dispersive X-ray spectroscopy; PECVD: plasma-enhanced chemical vapor deposition; N-C: N-doped carbon; SEM: scanning electron microscope.
After TNO was synthesized on VG, the porous structure of VG remained intact, and the TNO array exhibited a coral-like morphology. TNO nanoparticles facilitate the reduction of the ions/electrons transport path, thereby accelerating the reaction kinetics [Figure 1B and Supplementary Figure 3A]. The thickness of the TNO nanosheets (about 40 nm) increased further after the N-C layer coating on the TNO surface; however, the N-C@TNO array could maintain an excellent 3D pore structure, indicating that the N-C layer has good compatibility with TNO [Figure 1C and Supplementary Figure 3B]. The detailed microstructure of the TNO and N-C@TNO composites was examined using TEM and HRTEM. The TEM image [Supplementary Figure 4A] shows that the particle size of TNO is about 30-60 nm. HRTEM images
The phases and crystal structures of the samples at different stages were analyzed using XRD. As shown in Figure 2A, TNO powder is the precipitate obtained after centrifugation of the solvothermal product, and its diffraction peaks match the standard pattern of TiNb2O7 (PDF#04-007-0513). Both TNO and N-C@TNO show peaks at 26.2° and 44.4° corresponding to the (002) and (101) planes of graphite. The remaining major peaks are observed at 17.3°, 23.9°, 26.0°, 32.5°, 44.5°, and 47.0°, corresponding to the (002), (110), (003), (203), (005), and (020) planes of TiNb2O7. The absence of impurity peaks also confirms the successful fabrication of TNO, and there was no shift in the peak position after the introduction of the N-C coating layer, indicating the good compatibility of the N-C shell layer with TNO. The Raman spectra analysis, plotted in Supplementary Figure 5, provides further insights into the phase evolution of TNO and
Figure 2. Structure and composition of the samples. (A) XRD patterns comparing the TNO array and N-C@TNO array; (B) Survey XPS spectra, and high-resolution XPS spectra of (C) C 1s, (D) N 1s, (E) Nb 3d, and (F) O 1s. TNO: TiNb2O7, titanium niobium oxide; N-C:
Furthermore, the elemental composition and valence distribution on the surface of TNO and N-C@TNO are characterized by XPS. The survey spectra of N-C@TNO demonstrate the main elements (C, N, Ti, Nb, and O) in the composite arrays [Figure 2B]. Deconvolution of the C1s spectra reveals four peaks at 284.8 eV (C-C), 285.7 eV (C-O), 286.4 eV (C-N), and 287.6 eV (C=O), whereas no C-N signal is detected in TNO [Figure 2C]. The N 1s spectra can be deconvoluted into pyridine-N (397.1 eV), pyrrole-N (398.8 eV), and graphitic-N (400.4 eV), respectively [Figure 2D]. It is believed that N-C materials enhance electronic conductivity through the higher electronegativity of N atoms[42]. There was no N signal observed in TNO, proving that the N-C layer was successfully coated on the surface of TNO nanoparticles. On the one hand, the higher electronegativity of nitrogen relative to carbon enhances the electrical conductivity of the N-C layer; on the other hand, the comparable atomic radii of N and C minimize structural distortion during substitution, thereby preserving the structural integrity of the carbon framework. As for the Ti 2p spectra
To systematically evaluate the reaction kinetics of the N-C@TNO composite electrode, CV tests were conducted within a voltage window of 1.2-2.5 V (vs. Li/Li+) at a scan rate of 1.0 mV s-1 [Figure 3A]. The redox peaks observed at approximately 1.3 V, 1.6-1.7 V, and 1.9 V correspond to the redox reactions of
Figure 3. Electrochemical properties of TNO and N-C@TNO electrodes in lithium-ion batteries (LIBs): (A) Comparison of CV curves at a scan rate of 1.0 mV s-1; (B) Voltage gap (|ΔEp|) between the redox peaks at different scan rates; (C) Galvanostatic charge-discharge profiles at 5 C; (D) High-rate capability; (E) Comparison of the rate capability of TNO-based electrodes in the literature; (F) UV-vis absorption spectra; (G) Cycling performance at 20 C. TNO: TiNb2O7, titanium niobium oxide; N-C: N-doped carbon; CV: cyclic voltammetry; UV-vis: ultraviolet–visible spectroscopy.
To further elucidate the conductivity enhancement mechanism of the material’s electron transport properties. The optical bandgap energy (Eg) of the synthesized materials was determined using UV-Visible absorption spectroscopy, with the corresponding data presented in Figure 3F, expressed as
where α denotes the absorption coefficient, hν is the photon energy, and C represents a proportionality constant. Notably, the N-C@TNO composite exhibits a reduced bandgap of 2.57 eV compared to 2.83 eV for pristine TNO, indicating that the N-C coating effectively enhances electronic conductivity. This facilitation of charge transport promotes improved kinetics for Li+ insertion/extraction, which is particularly advantageous for high-rate battery applications. More importantly, the N-C@TNO electrode exhibits superior cyclic stability, with a capacity retention of 95.38% at 20 C and a nearly 100% Coulombic efficiency after 5,000 cycles. This performance is superior to that of the TNO electrodes, which exhibit a capacity retention of 83.37% [Figure 3G].
Multiple electrochemical methods, including the galvanostatic intermittent titration technique (GITT), EIS, and pseudocapacitive contribution obtained by CV, have been implemented in order to gain deeper insights into the diffusion kinetics. Supplementary Figure 9 exhibits the GITT curves of TNO and N-C@TNO. Based on Fick’s second law of diffusion, the Li+ diffusion coefficients (
where ΔEs represents the steady state voltage change induced by the current pulse; ΔEt denotes the total voltage variation during the pulse; τ represents the constant current pulse time; mB is the quality of the active material, S refers to the electrode-electrolyte contact area, and ρ indicates the compacted density. As derived from the GITT data, N-C@TNO exhibits a higher
Figure 4. The calculated
Here, a is an adjustable parameter, while the magnitude of b serves to distinguish between different types of charge storage mechanisms. A b-value of 0.5 indicates that the lithium-ion intercalation/deintercalation process is predominantly governed by diffusion-controlled behavior, whereas a b-value of 1.0 is characteristic of surface-controlled capacitive processes. As shown in Figure 4C, the b-values derived from the analysis of the oxidation peaks are 0.915 for N-C@TNO and 0.870 for TNO, indicating that the electrochemical reactions of both electrodes involve a combination of pseudocapacitive and diffusion-controlled processes. To further quantify and distinguish between the capacitive and diffusion-controlled contributions, the expression can be reformulated as follows:
where k1v and k2v correspond to the current contributions from capacitive-controlled and diffusion-controlled processes, respectively. The pseudocapacitive contribution of the electrodes can be calculated for different sweep speeds, which can be written as follows:
Figure 4D and E displays the capacitive contribution of TNO with the value of 79.79% at a scan rate of
To verify the enhancement of N-C@TNO stability, we captured SEM, TEM, and HRTEM images after rate cycling and long cycling processes. Following the cycling process, the N-C@TNO electrode almost maintains its original architecture, still preserving a well-defined 3D porous framework without visible cracks or delamination [Figure 5A and B]. In contrast, the TNO electrode suffers severe particle agglomeration, extensive cracking, and complete collapse of its porous network [Supplementary Figure 11A and B]. These structural degradations directly account for its electrochemical performance decay, consistent with its inferior long-term cycling performance. TEM and HRTEM images show that the N-C coating layer on the surface maintains its intact morphology without degradation during the cycling process [Figure 5C and D]. EDX analysis of the images showed that the C and N signals are uniformly overlapped with Ti and Nb signals, indicating that the N-C layer remained tightly bound to TNO after rate cycling and long cycling processes [Supplementary Figure 12]. Furthermore, the high-resolution XPS spectra of the cycled
Figure 5. Morphology of the N-C@TNO array after rate and long-term cycling. (A and B) SEM images, (C) TEM image, and (D) HRTEM image. TNO: TiNb2O7, titanium niobium oxide; N-C: N-doped carbon; TEM: transmission electron microscopy; HRTEM: high-resolution TEM; SEM: scanning electron microscope.
The excellent electrochemical performance of our N-C@TNO electrode can be attributed to the following factors: Firstly, the VG skeleton and the N-C layer form a 3D continuous conductive network, which can accelerate electron transfer and provide a sandwiched structure to enhance the reaction kinetics and
CONCLUSIONS
In summary, we constructed a sandwich-structured electrode through a combined PECVD strategy. TNO nanoparticles were uniformly anchored on a VG skeleton and coated by a N-C shell, establishing a distinctive “conductive skeleton || active core || protective shell” architecture. The synergistic coupling between the highly conductive VG framework and the N-C coating significantly enhanced the electronic/ionic conductivity and cycling stability of electrodes. Compared to pristine TNO, the N-C@TNO composite exhibited markedly reduced polarization and charge transfer resistance, accelerated Li+ diffusion kinetics, and predominantly pseudocapacitive charge storage behavior. Consequently, the N-C@TNO electrode delivered exceptionally high-rate capability and ultralong cycle life.
DECLARATIONS
Authors’ contributions
Investigation, data collection and analysis, methodology, writing - original draft: Li, J.
Methodology, data collection and analysis, writing - review and editing: Tang, C.
Data collection and analysis, resources, methodology: Li, C.; Zhang, T.
Investigation, methodology: Liang, X.; Sheng, Y.
Supervision, writing - review and editing, funding acquisition: Xia, X.; Zhang, Y.; Liu, J.
Availability of data and materials
The data are available upon request from the corresponding authors.
Financial support and sponsorship
This work was supported by the Hunan Provincial Natural Science Foundation of China (No. 2023JJ30499), the Guangdong Provincial Natural Science Foundation of China (No. 2022A1515010198), the National Natural Science Foundation of China (No. 22379092), the Science and Technology Department of Zhejiang Province (No. 2023C01231), the China Postdoctoral Science Foundation (No. 2024M750347), the Sichuan Natural Science Foundation (No. 2024NSFSC0951), and the Key Laboratory of Engineering Dielectrics and Its Application, Harbin University of Science and Technology, Ministry of Education (No. KMF 202303).
Conflicts of interest
All authors declared that there are no conflicts of interest.
Ethical approval and consent to participate
Not applicable.
Consent for publication
Not applicable.
Copyright
© The Author(s) 2025.
Supplementary Materials
REFERENCES
1. Duan, J.; Tang, X.; Dai, H.; et al. Building safe lithium-ion batteries for electric vehicles: a review. Electrochem. Energ. Rev. 2020, 3, 1-42.
2. Wu, J.; Zheng, Y.; Zhang, P.; et al. A rocking-chair rechargeable seawater battery. Research 2024, 7, 0461.
3. Tang, S.; Sun, F.; Wang, H.; et al. Revealing the degradation mechanism of lithium-ion batteries for electric aircraft. Adv. Mater. 2025, 37, 2502363.
4. Yang, Y.; Zhao, J. Wadsley-roth crystallographic shear structure niobium-based oxides: promising anode materials for high-safety lithium-ion batteries. Adv. Sci. 2021, 8, 2004855.
5. Geng, C.; Zhang, P.; Wu, J. M.; Qin, J.; Wen, W. Lattice expanded titania as an excellent anode for an aqueous zinc-ion battery enabled by a highly reversible H+-promoted Zn2+ intercalation. ACS. Nano. 2024, 18, 33119-30.
6. Chen, X.; Liu, J.; Li, W.; et al. Permselective covalent organic framework membrane as self-extinguishing separator for high-safety lithium-ion battery. Angew. Chem. Int. Ed. 2025, 64, e202512591.
7. Yuan, K.; Lin, Y.; Li, X.; et al. High-safety anode materials for advanced lithium-ion batteries. Energy. Environ. Mater. 2024, 7, e12759.
8. Zhang, X.; Sun, J.; Cheng, Z.; Wu, M.; Guo, Z.; Zhang, H. Design, perspective, and challenge of niobium-based anode materials for high-energy alkali metal-ion batteries. Adv. Funct. Materials. 2024, 34, 2405392.
9. Griffith, K. J.; Seymour, I. D.; Hope, M. A.; et al. Ionic and electronic conduction in TiNb2O7. J. Am. Chem. Soc. 2019, 141, 16706-25.
10. Zhang, Y.; Kang, C.; Zhao, W.; et al. Crystallographic engineering to reduce diffusion barrier for enhanced intercalation pseudocapacitance of TiNb2O7 in fast-charging batteries. Energy. Storage. Mater. 2022, 47, 178-86.
11. Huang, J.; Yang, Q.; Yang, W.; et al. Voltage regulation via covalent bond strength to increase energy density for safe fast-charging lithium-ion batteries. Adv. Funct. Mater. 2025, 35, 2422105.
12. Sieffert, J. M.; Lang, C. J.; Bazylevych, S.; Jia, S.; Mccalla, E. The Nb-Ti-W-O system as safe high-power anodes for Li-ion batteries. J. Mater. Chem. A. 2024, 12, 1429-37.
13. Zhou, Y.; Ding, Y.; Chen, Y.; et al. Thermal degradation of lithium-ion battery cathodes: a machine learning prediction of stability and safety. Energy. Mater. 2025, 5, 500077.
14. Zhang, Y.; Wang, Y.; Zhao, W.; et al. Delocalized electronic engineering of TiNb2O7 enables low temperature capability for high-areal-capacity lithium-ion batteries. Nat. Commun. 2024, 15, 6299.
15. Yu, G.; Huang, J.; Bai, X.; et al. Engineering of cerium modified TiNb2O7 nanoparticles for low-temperature lithium-ion battery. Small 2024, 20, 2308858.
16. Wei, C.; Zhang, Y.; Tian, Y.; et al. Design of safe, long-cycling and high-energy lithium metal anodes in all working conditions: progress, challenges and perspectives. Energy. Storage. Mater. 2021, 38, 157-89.
17. Gao, F.; Yang, S.; Zhang, Z.; et al. Synergistic internal and external modification of TiNb2O7 through ion doping and interfacial engineering for high-performance lithium-ion batteries. Carbon 2025, 238, 120217.
18. Mao, B.; Guo, D.; Qin, J.; Meng, T.; Wang, X.; Cao, M. Solubility-parameter-guided solvent selection to initiate ostwald ripening for interior space-tunable structures with architecture-dependent electrochemical performance. Angew. Chem. Int. Ed. 2018, 57, 446-50.
19. Jiang, M.; Hu, Y.; Mao, B.; et al. Strain-regulated Gibbs free energy enables reversible redox chemistry of chalcogenides for sodium ion batteries. Nat. Commun. 2022, 13, 5588.
20. Huang, L.; Huang, Y.; Wang, J.; et al. Synergistic dual-carbon networks bridged Mn-doped TiNb2O7 anode for fast-charging lithium-ion batteries. ACS. Appl. Mater. Interfaces. 2025, 17, 53659-69.
21. Li, S. Q.; Wang, Z.; Zheng, X.; et al. Biomimetic sandwich-structured tubular ion pump arrays for lithium metal batteries. J. Am. Chem. Soc. 2025, 147, 25883-95.
22. Sun, X.; Qin, C.; Zhao, B.; et al. A cation and anion dual-doping strategy in novel Li-rich Mn-based cathode materials for high-performance Li metal batteries. Energy. Storage. Mater. 2024, 70, 103559.
23. Tian, K.; Wang, Z.; Di, H.; et al. Superimposed effect of La doping and structural engineering to achieve oxygen-deficient TiNb2O7 for ultrafast Li-Ion storage. ACS. Appl. Mater. Interfaces. 2022, 14, 10478-88.
24. Huang, S.; Qin, J.; Yu, X.; et al. Material hardening caused by lattice distortion enables good cycling stability of entropy-increased energy-storage materials. Adv. Energy. Mater. 2025, 15, 2502286.
25. Wu, H.; Wei, L.; Li, W.; et al. Highly conductive carbon/carbon composites as advanced multifunctional anode materials for structural lithium-ion batteries. Adv. Funct. Mater. 2024, 34, 2403729.
26. Wu, Y.; Liu, D.; Qu, D.; et al. Porous oxygen-deficient TiNb2O7 spheres wrapped by MXene as high-rate and durable anodes for liquid and all-solid-state lithium-ion batteries. Chem. Eng. J. 2022, 438, 135328.
27. Yang, Y.; Yue, Y.; Wang, L.; et al. Facile synthesis of mesoporous TiNb2O7/C microspheres as long-life and high-power anodes for lithium-ion batteries. Int. J. Hydrogen. Energy. 2020, 45, 12583-92.
28. Ho, V.; Huynh, T. N.; Pham, T. H.; et al. Dry basal plane graphene wrappings on spherical nickel-rich oxide layered particles for lithium-ion batteries. J. Energy. Chem. 2025, 104, 10-9.
29. Folorunso, O.; Sadiku, R.; Hamam, Y.; Kupolati, W. Sustainable metal oxides and their composites for lithium-ion batteries and multifunctional applications. J. Mater. Sci. 2025, 60, 12347-84.
30. Guo, W.; Liu, S.; Guan, X.; Zhang, X.; Liu, X.; Luo, J. Mixed ion and electron-conducting scaffolds for high-rate lithium metal anodes. Adv. Energy. Mater. 2019, 9, 1900193.
31. Sun, H.; Mei, L.; Liang, J.; et al. Three-dimensional holey-graphene/niobia composite architectures for ultrahigh-rate energy storage. Science 2017, 356, 599-604.
32. Xu, W.; Xu, Y.; Grzimek, V.; et al. Insights into the kinetics-morphology relationship of 1-, 2-, and 3D TiNb2O7 anodes for Li-ion storage. Nano. Res. 2024, 17, 2770-80.
33. Du, H.; Wang, Y.; Kang, Y.; et al. Side reactions/changes in lithium-ion batteries: mechanisms and strategies for creating safer and better batteries. Adv. Mater. 2024, 36, 2401482.
34. Wen, W.; Geng, C.; Li, X.; et al. A membrane-free rechargeable seawater battery unlocked by lattice engineering. Adv. Mater. 2024, 36, 2312343.
35. Chen, G.; Liu, X.; Liu, Z.; et al. Novel “sandwich” configuration with ALD-coating layers on electrode/electrolyte interfaces for durable all-solid-state lithium metal batteries with high-voltage cathodes. Energy. Mater. 2025, 5, 500064.
36. Park, S. K.; Kim, S.; He, R.; et al. Synergistic lithium alloying and plating in 3D Cu/CNT/Sn electrodes for stable lithium metal batteries. Small 2025, 21, e2501292.
37. Zhang, S.; Xiao, S.; Li, D.; et al. Commercial carbon cloth: an emerging substrate for practical lithium metal batteries. Energy. Storage. Mater. 2022, 48, 172-90.
38. Yang, Y.; Li, Y.; Liu, K.; et al. Hierarchical porous TiNb2O7@N-doped carbon microspheres as superior anode materials for lithium ion storage. Int. J. Hydrogen. Energy. 2021, 46, 3425-36.
39. Wang, J.; Paul, T.; Chandan, P.; et al. Rate performance enhancement in lithium-ion batteries using TiNb2-xAlxO7 anodes with self-generated protective layers. Chem. Eng. J. 2025, 504, 158464.
40. Yao, Z.; Xia, X.; Zhang, S.; et al. Oxygen defect boosted N-doped Ti2Nb10O29 anchored on core-branch carbon skeleton for both high-rate liquid & solid-state lithium ion batteries. Energy. Storage. Mater. 2020, 25, 555-62.
41. Lou, S.; Cheng, X.; Zhao, Y.; et al. Superior performance of ordered macroporous TiNb2O7 anodes for lithium ion batteries: understanding from the structural and pseudocapacitive insights on achieving high rate capability. Nano. Energy. 2017, 34, 15-25.
42. Oh, Y. J.; Park, J. H.; Park, J. S.; et al. Fast-chargeable N-doped multi-oriented graphitic carbon as a Li-intercalation compound. Energy. Storage. Mater. 2022, 44, 416-24.
43. Chung, W.; Bang, J. H. Carbon-doped TiNb2O7 suppresses amorphization-induced capacity fading. ACS. Appl. Mater. Interfaces. 2022, 14, 19365-75.
44. Cui, P.; Zhang, P.; Chen, X.; et al. Oxygen defect and Cl--doped modulated TiNb2O7 compound with high rate performance in lithium-ion batteries. ACS. Appl. Mater. Interfaces. 2023, 15, 43745-55.
45. Shen, S.; Deng, S.; Zhong, Y.; et al. Binder-free carbon fiber/TiNb2O7 composite electrode as superior high-rate anode for lithium ions batteries. Chin. Chem. Lett. 2017, 28, 2219-22.
46. Zhong, X.; Huang, T.; Liang, J.; et al. Porous TiNb2O7@N-C as anode materials for lithium-ion batteries with ultrahigh-rate performance. J. Phys. Chem. C. 2021, 125, 23960-7.
Cite This Article
How to Cite
Download Citation
Export Citation File:
Type of Import
Tips on Downloading Citation
Citation Manager File Format
Type of Import
Direct Import: When the Direct Import option is selected (the default state), a dialogue box will give you the option to Save or Open the downloaded citation data. Choosing Open will either launch your citation manager or give you a choice of applications with which to use the metadata. The Save option saves the file locally for later use.
Indirect Import: When the Indirect Import option is selected, the metadata is displayed and may be copied and pasted as needed.
About This Article
Copyright
Data & Comments
Data














Comments
Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at [email protected].