Next Article in Journal
Electrostatic Field Modification Enhances the Electrocatalytic Oxygen Evolution Reaction Stability of CoFe2O4 Catalysts
Previous Article in Journal
Interposer-Based ESD Protection: A Potential Solution for μ-Packaging Reliability of 3D Chips
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rapid Synthesis of Fast-Charging TiNb2O7 for Lithium-Ion Storage via Ultrafast Carbothermal Shock

1
State Key Laboratory of ASIC & System, School of Microelectronics, Fudan University, Shanghai 200433, China
2
Key Laboratory of Multifunctional Nanomaterials and Smart Systems, Division of Advanced Materials, Suzhou Institute of Nano-Tech and Nano-Bionics, Chinese Academy of Sciences, Suzhou 215123, China
3
College of Textiles and Clothing, Yancheng Institute of Technology, Yancheng 224051, China
4
School of Microelectronics, Northwestern Polytechnical University, No. 1 Dongxiang Road, Chang’an District, Xi’an 710129, China
*
Authors to whom correspondence should be addressed.
Micromachines 2025, 16(5), 490; https://doi.org/10.3390/mi16050490
Submission received: 18 February 2025 / Revised: 8 April 2025 / Accepted: 16 April 2025 / Published: 22 April 2025
(This article belongs to the Section C:Chemistry)

Abstract

:
The development of fast-charging lithium-ion batteries urgently requires high-performance anode materials. In this paper, through an ultrafast carbothermal shock (CTS) strategy, titanium niobium oxide (TiNb2O7, TNO) with an optimized structure was successfully synthesized within 30 s. By regulating the synthesis temperature to 1200 °C, the TNO-1200 material was obtained. Its lattice parameters (a-axis: 17.6869 Å) and unit-cell volume (796.83 Å3) were significantly expanded compared to the standard structure (a-axis: 17.51 Å, volume ~790 Å3), which widened the lithium-ion migration channels. Rietveld refinement and atomic position analysis indicated that the partial overlap of Ti/Nb atoms and the cooperative displacement of oxygen atoms induced by CTS reduced the lithium-ion diffusion energy barrier. Meanwhile, the cation disorder suppressed the polarization effect. Electrochemical tests showed that after 3000 cycles at a current density of 10 C, the specific capacity of TNO-1200 reached 125 mAh/g, with a capacity retention rate of 98%. EDS mapping confirmed the uniform distribution of elements and the absence of impurity phases. This study provides an efficient synthesis strategy and theoretical basis for the design of high-performance fast-charging battery materials through atomic-scale structural engineering.

1. Introduction

With the rapid development of fields such as electric vehicles (EVs) and portable electronic devices, the demand for high-performance lithium-ion batteries is increasing [1,2]. Fast charging performance, as a key performance indicator of lithium-ion batteries, is significant for enhancing user experience and expanding the range of battery applications [3]. Traditional graphite anode materials, due to their relatively low theoretical specific capacity (372 mAh g−1) and safety issues such as lithium dendrite growth during fast charging, struggle to meet the demands for future high energy density and high power density batteries [4]. Therefore, the development of novel anode materials with high specific capacity, excellent rate performance, and good safety has become a focus of current research [5,6,7,8,9].
Among various novel anode materials, titanium niobium oxide (TiNb2O7, TNO) has attracted considerable attention due to its unique crystal structure and electrochemical properties [10,11,12]. TNO possesses three accessible redox couples (Ti4+/Ti3+, Nb5+/Nb4+, and Nb4+/Nb3+), allowing it to accommodate five lithium ions within the electrochemical stability window of organic liquid electrolytes (TiNb2O7 + 5Li+ + 5e → Li5TiNb2O7), with a theoretical specific capacity as high as 388 mAh g−1, surpassing that of traditional graphite anodes [13]. Moreover, the potentials of these redox couples are sufficiently high to effectively prevent electrolyte decomposition and lithium dendrite nucleation under high-rate charge and discharge conditions, thereby enhancing battery safety under harsh conditions [14,15,16,17,18,19]. However, TNO also presents some inherent limitations that restrict its practical applications. One of the main issues is its low electronic conductivity (<10−9 S cm−1), which leads to poor rate performance [20,21,22,23,24]. To enhance the electronic conductivity and ionic diffusion rate of TNO, researchers have explored various methods, such as carbon coating, reducing particle size, metal ion doping, and oxygen defect engineering [25,26,27,28]. Among these strategies, oxygen defect engineering introduces oxygen vacancies into the bulk material, inducing lattice distortion that enhances electronic conductivity and charge transfer kinetics, thus effectively improving the energy storage characteristics of the electrode material [29,30]. Nevertheless, existing methods for generating oxygen defects in metal oxides often involve chemical or electrochemical reducing agents or require vacuum sintering conditions [31]. These approaches are not only complex but also costly, making them less favorable for large-scale production [32,33,34]. In addition to the issue of low electronic conductivity, the synthesis methods of TNO also have significant drawbacks in terms of time and energy consumption. Traditional synthesis methods for TNO typically involve solid-state reaction processes that require long sintering times at high temperatures (ranging from several hours to tens of hours) [35]. This not only increases production costs but also reduces manufacturing efficiency. Therefore, developing a rapid and efficient synthesis method for TNO holds substantial practical significance.
In this work, we propose a carbothermal shock (CTS) strategy for the rapid synthesis of fast-charging TiNb2O7-1200, achievable within 30 s. Based on the above strategy, the synthesized fast-charging TiNb2O7 enables the adjustment of the occupancy of Ti and Nb atoms within the TiNb2O7 crystal framework, thereby expanding the transport pathways for Li ions in the TiNb2O7 crystal structure. Consequently, due to the modification of the lithium-ion transport channels in TiNb2O7, the material achieves a specific capacity of 125 mAh/g at a current density of 10 C under a voltage window of 0.01–3 V after 3000 cycles, with a capacity retention rate close to 98% post-formation. More importantly, through the refinement of the structural data of the TiNb2O7 samples, it was observed that the spatial occupancy of Ti and Nb atoms within the crystal framework significantly differs from that in the classic TiNb2O7 ilmenite structure. This change is crucial for the exceptional structural stability of TiNb2O7 under high current density conditions, providing a viable strategy to address the current limitations in the electronic conductivity of TiNb2O7.

2. Experiment

2.1. Ultrafast Carbothermal Shock Rapid Synthesis of TiNb2O7

Using analytical grade TiO2 (purity 99.8%, Sigma-Aldrich, St. Louis, MO, USA) and Nb2O5 (purity 99.8%, CBMM, Araxá, Brazil) as raw materials, precise weighing is performed according to a 1:1 molar ratio. The weighed raw materials are placed in a high-energy ball mill (SPEX 8000D Mixer/Mill, Metuchen, NJ, USA) and milled for 1 h to ensure thorough mixing. The mixed raw materials are then cold-pressed to form “precursor pellets”. A homemade Joule heating device is used to set up an ultrafast high-temperature annealing system (UHS) for synthesizing TNO. The specific steps are as follows: the prepared precursor pellets are placed between two pieces of graphite foil, which serve as heating elements, and are positioned in the UHS. By precisely controlling the current, the precursor pellets are rapidly heated to 1100, 1200, and 1400 °C under an argon protective atmosphere and maintained at this high temperature for 30 s, followed by rapid cooling. This process is repeated twice, with a 3 min interval each time to avoid thermal damage to the heating chamber, electrical connections, and graphite foil. After annealing, the resulting dense TNO-1100, TNO-1200, and TNO-1400 pellets are placed into a mortar and ground to a fine powder using a pestle. Then, the fine particles are placed in a planetary ball mill (Across International, PQ-N04, Livingston, NJ, USA) with isopropanol as the solvent and milled for 8 h at a speed of 300 RPM to reduce particle size. After wet milling, the samples are dried in an oven at 80 °C for 12 h, yielding the TNO-1100, TNO-1200, and TNO-1400 anode materials, respectively.

2.2. Material Characterization

To investigate the geometric morphology and crystalline structure, all TiNb2O7 samples were characterized using a scanning electron microscope (SEM, Zeiss Zigma FESEM, Oberkochen, Germany). Detailed testing conditions are as follows: samples were mounted on conductive tape, and a 5 kV accelerating voltage was applied to minimize beam damage to the samples and generate high-quality images. The crystal structure of the samples was further studied by the X-ray diffraction (XRD: Rigaku Smartlab 9000W, Tokyo, Japan). The XRD measurements were performed using Cu Kα radiation (λ = 0.154056 nm) with a scanning range of 2θ from 5° to 90°, a scanning rate of 5°/min, and a step size of 0.02°, based on typical high-resolution configurations for precise phase analysis of materials like TiNb2O7. This set up balances data quality and efficiency, ensuring comprehensive coverage of diffraction peaks while maintaining sufficient resolution for crystallographic characterization.

2.3. Electrochemical Test

The electrode was prepared by mixing TNO (TNO-1100, TNO-1200, and TNO-1400), Super P, and PVDF with a weight ratio of 7:2:1. With the addition of the NMP solvent, the formed slurry was coated onto Cu foil and dried at 80 ◦C in a vacuum oven for 12 h. Then, the electrode with a uniform mass loading of 1.0 mg/cm2 was formed on the copper foil. Subsequently, the electrode was cut into a disk of 14 mm in diameter. The electrochemical performance of the electrode we prepared was measured via CR2025 coin-type cells. The lithium foil and the polypropylene membrane (Celgard, Charlotte, NC, USA) served as counter electrode and separator, respectively. The electrolyte for the Li-ion half-cell was self-built with a composition of 1.0 M LiPF6 in EC-DEC (1:1, v%). All reagents and materials were used directly without further purification. Galvanostatic charge/discharge was conducted on a NEWARE battery testing system (CT-4008) within a voltage window of 0.01–3.0 V at room temperature. C-rate represents the ratio between the charge/discharge current and the rated capacity of the battery, with the unit “C”. The formula is as follows:
C   rate = Current   density   mAg 1 Practical   capacity   mAhg 1
Theoretical capacity: According to the literature, TiNb2O7 can intercalate 5 Li+ ions (reaction: TiNb2O7 + 5Li+ + 5e → Li5TiNb2O7), yielding a theoretical capacity of 388 mAh/g. Based on this, a rate of 10 C would correspond to 3880 mA/g.

3. Results and Discussion

The morphologies were observed using scanning electron microscopy (SEM) at both low and high magnifications, with corresponding results presented in Figure 1. From the results shown in Figure 1a,b, it is evident that the particles formed from TNO-1100 after ball milling exhibit relatively uniform diameters and smoother surfaces. This indicates that the strength of TiNb2O7 synthesized at 1100 °C is relatively weak, making it easier to achieve uniform particle sizes under high-energy ball milling impacts. In contrast, when the temperature is raised to 1200 °C and 1400 °C, the results in Figure 1c–f reveal a significant increase in particle diameter under the same ball milling conditions. This implies that the strength of the resulting TiNb2O7 gradually increases with temperature. This observation also explains why larger particles exhibit the presence of finer particles on their surfaces. Furthermore, it can be noted that the particles obtained from TNO-1200, synthesized at 1200 °C, demonstrate a diameter distribution that aligns more closely with a normal distribution after ball milling, which is advantageous for enhancing the packing density of electrode materials. This was further confirmed in subsequent lithium storage performance tests.
As can be seen from Figure 2, the X-ray diffraction (XRD) patterns of TNO-1100, TNO-1200, and TNO-1400 synthesized at 1100 °C, 1200 °C, and 1400 °C, respectively, are shown. By comparing with the standard card (JCPDS: 77-1374), it can be found that the peak positions are in good agreement with it, demonstrating that TNO-1100, TNO-1200, and TNO-1400 all belong to the layered structure of ReO₃. At 1100 °C (blue line), the XRD pattern exhibits relatively weak diffraction peaks, indicating a lower degree of crystallinity. The intensity of the peaks is moderate, and the peak width is relatively broad, suggesting that the crystal grains may be small or the crystallization process is not fully completed. When the temperature increases to 1200 °C (red line), the diffraction peaks become sharper and more intense compared to those at 1100 °C. This implies an improvement in the crystallinity of the TiNb2O7 electrode material. The enhanced peak intensity and sharpness suggest that larger and more perfect crystal grains are formed during the higher-temperature treatment. At 1400 °C (green line), the XRD pattern shows the most intense and sharpest diffraction peaks among the three temperatures. The well-defined and high-intensity peaks further confirm the high crystallinity of the sample at this temperature. Additionally, some new peaks may appear, or the relative intensities of existing peaks may change, which could be related to phase transformations or structural changes in the TiNb2O7 electrode material at such a high temperature. These differences in XRD patterns at different temperatures provide valuable insights into the thermal-induced structural evolution and crystallization behavior of the TiNb2O7 electrode material. However, the structural differences resulting from different synthesis temperatures can significantly alter the lithium storage performance of TiNb2O7. This conclusion has been verified in the evaluation of the lithium storage performance of TNO-1100, TNO-1200, and TNO-1400, and the results are presented in Figure 3.
To examine the lithium storage performance of the synthesized TiNb2O7 during long-term cycling under a high current density, the test was conducted at a current density of 10 C within a voltage window of 0.01–3.0V. After 3000 cycles, it was found that TNO-1200 achieved a specific capacity of 125 mAh/g. This value is higher than the specific capacities of the TNO-1100 and TNO-1400 electrodes. Moreover, the capacity retention rate of TNO-1200 after 3000 cycles was 98%.
To elucidate the underlying mechanisms responsible for the exceptional fast-charging performance of TNO-1200, a systematic investigation of its elemental distribution and microstructure was conducted using area scan energy-dispersive X-ray spectroscopy (EDS) (Figure 4b–d) and line scan EDS (Figure 4e). The elemental mapping images of Ti, Nb, and O (Figure 4b–d) reveal highly homogeneous spatial distributions across the TNO-1200 surface, with no apparent segregation or localized enrichment observed (scale bar: 2.5 μm). This uniformity indicates successful construction of a homogeneous crystal structure during synthesis, accompanied by precise control of elemental stoichiometry. The coherent arrangement of Ti, Nb, and O not only preserves the integrity of lithium-ion insertion/extraction channels but also effectively mitigates localized polarization effects during charge/discharge processes, thereby significantly enhancing lithium-ion diffusion kinetics. Further examination of intraparticle compositional homogeneity was performed through cross-sectional line scanning (Figure 4e). The concentration profiles demonstrate gradual variations for Ti, Nb, and O along the particle cross-section (scan range: 0.1–1.6 μm), with oxygen content stabilized within 0.8–1.0 and a constant Ti/Nb atomic ratio. These findings confirm global structural homogeneity, fundamentally eliminating stress concentration or volume expansion caused by compositional gradients, thereby ensuring structural stability under high-rate cycling conditions. Notably, the uniform coexistence of Ti4+ and Nb5+ suggests a unique synergistic mechanism—the Ti4+ ions establish a stable crystalline framework, while the high oxidation state of Nb5+ substantially enhances electronic conductivity, collectively optimizing rate capability. Additionally, the absence of impurity phases in EDS analysis confirms superior surface cleanliness, which facilitates reduced electrode/electrolyte interfacial resistance and accelerated charge transfer kinetics, ultimately supporting stable cycling performance at elevated current densities. While EDS mapping was conducted solely on TNO-1200 due to its exceptional performance, XRD patterns (Figure 2) and Rietveld refinement (Figure 5) confirm phase purity for all samples, suggesting homogeneous element distribution. The performance disparities among TNO-1100, TNO-1200, and TNO-1400 are attributed to differences in crystallinity (Figure 2) and lattice expansion (Table 1) rather than compositional inhomogeneity. This comprehensive structural characterization provides critical insights into the intrinsic advantages of TNO-1200 as a high-performance anode material for fast-charging lithium-ion batteries.
The crystallographic characteristics of TiNb2O7 (TNO-1100, TNO-1200, and TNO-1400) synthesized via the CTS were systematically investigated through Rietveld refinement of XRD data (Figure 5) and the TNO-1200 detailed atomic position analysis (Table 1). By comparing the refined structural parameters and atomic coordinates with the standard monoclinic TiNb2O7 phase (JCPDS: 77-1374, space group: C2/m), we elucidated how synthesis-induced structural modifications enhance lithium-ion migration kinetics [36]. The Rietveld-refined XRD pattern (Figure 5) exhibits excellent agreement between experimental and calculated profiles, supported by low reliability factors (Rₚ = 5.08, RWP = 8.30, χ2 = 5.09). Minor deviations in lattice parameters (a = 17.6869 Å, b = 3.8030 Å, c = 11.8976 Å, β = 95.32233°, V = 796.830 Å3) from the JCPDS standard suggest subtle lattice distortions induced by the rapid Joule flash process [37]. Atomic coordinate analysis (Table 2) reveals critical insights: Ti and Nb atoms occupy distinct Wyckoff positions compared to the conventional structure. For instance, Nb2 and Ti2 share identical coordinates (0.18483, 0.0000, 0.18442), while Nb3 and Ti3 are displaced to 0.07844, 0.0000, −0.55762. This partial overlap and positional displacement contrast with the strict cation ordering in standard TiNb2O7, indicating a tailored atomic arrangement that reduces site-specific energy barriers for Li+ diffusion [38]. The adjusted atomic configuration directly optimizes lithium-ion migration pathways. The slight expansion in the a-axis (17.6869 Å vs. 17.51 Å) and increased unit cell volume (V = 796.83 Å3 vs. ~790 Å3) enlarge interstitial spaces, reducing steric hindrance for Li+ transport. Concurrently, coordinated displacements of oxygen atoms (O1–O11), particularly O10 at 0.50000, 0.0000, 0.50000, stabilize Li+ hopping trajectories through a symmetric framework. Furthermore, the overlapping Ti/Nb positions introduce controlled cation disorder, disrupting long-range electrostatic interactions and lowering activation energy for Li+ migration—a feature absent in the standard structure, underscoring the unique advantage of CTS in atomic-scale engineering. These structural modifications address key limitations of conventional TiNb2O7 anodes. Widened ionic channels and disordered cation arrangements enable faster Li+ diffusion, as inferred from reduced polarization in electrochemical measurements. Simultaneously, homogeneous Ti4+/Nb5+ distribution mitigates localized stress during high-rate cycling, enhancing structural stability and preventing capacity fade. The CTS method thus effectively engineers TiNb2O7’s atomic structure to optimize ionic and electronic transport properties. Validated by XRD refinement and atomic position analysis, these refinements highlight the material’s potential as a high-performance anode for fast-charging lithium-ion batteries. This work aligns with advances in defect engineering for battery materials, offering a blueprint for rational design of fast-charging electrode architectures.

4. Conclusions

In this study, fast-charging TiNb2O7 (TNO) anode materials were successfully synthesized via an ultrafast carbothermal shock (CTS) strategy, precisely controlling atomic-scale structural changes. The optimized TNO-1200 showed excellent electrochemical performance, with a 125 mAh/g specific capacity at 10 C over 3000 cycles and 98% capacity retention. Rietveld refinement and atomic position analysis revealed that the rapid Joule flash process led to lattice expansion (a-axis: 17.6869 Å vs. 17.51 Å) and unit cell volume increase (V = 796.83 Å3 vs. ~790 Å3), widening Li+ migration channels. Coordinated oxygen displacements and controlled cation disorder (overlapping Ti/Nb sites) reduced Li+ diffusion activation energy and enhanced conductivity, mitigating polarization and stress, thus ensuring high-rate cycling stability. The CTS method was highly efficient, synthesizing phase-pure TNO in 30 s, avoiding long-term, high-temperature treatments of conventional methods. EDS mapping confirmed homogeneous element distribution and no impurity phases, highlighting its scalability. This work emphasizes atomic-scale engineering’s importance for fast-charging battery materials. Future research should correlate structural parameters with electrochemical metrics to establish structure-property relationships. The CTS strategy offers a platform for next-generation electrode design, connecting rapid synthesis with high-performance energy storage.

Author Contributions

Conceptualization, F.Y.; Methodology, Y.Z.; Software, Y.Z.; Formal analysis, Y.Z.; Investigation, X.H. (Xianyu Hu); Resources, X.F.; Data curation, X.H. (Xianyu Hu), Y.Z. and F.Y.; Writing—original draft, X.H. (Xianyu Hu); Writing—review & editing, X.H. (Xiaosai Hu), X.F. and F.Y.; Visualization, X.H. (Xianyu Hu); Supervision, X.H. (Xiaosai Hu), X.F. and F.Y.; Project administration, X.H. (Xiaosai Hu), X.F. and F.Y.; Funding acquisition, X.H. (Xiaosai Hu), X.F. and F.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (52402068), Natural Science Foundation of Jiangsu Province (Grant No. BK20191192 and BK20240478), the Guiding Program of China Textile Industry Federation (Grant No. 2022035), Jiangsu Independent Innovation Fund Project of Agricultural Science and Technology (CX (21)3163), National Natural Science Foundation of China (Grant No. 51705113, Grant No. 21971113, Grant No. 22175094), the Jiangsu Funding Program for Excellent Postdoctoral Talent (No. 2023ZB167, No. 2023ZB695, No. 2023ZB065, No. 2023ZB293), the China Postdoctoral Science Foundation (No. 2023M742557, No. 2023M742561), and the Project supported by the Fundamental Research Funds for the Central Universities of Ministry of Education of China (No. D5000240188).

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Acknowledgments

The authors express their gratitude for the technical support provided by Nano-X from the Suzhou Institute of Nano-Tech and Nano-Bionics, Chinese Academy of Sciences (SINANO).

Conflicts of Interest

The authors declare no competing interest.

References

  1. Fellner, J.P.; Loeber, G.J.; Sandhu, S.S. Testing of lithium-ion 18650 cells and characterizing/predicting cell performance. J. Power Sources 1999, 81, 867–871. [Google Scholar] [CrossRef]
  2. Kida, Y.; Kinoshita, A.; Yanagida, K.; Funahashi, A.; Nohma, T.; Yonezu, I. A study on the cycle performance of lithium secondary batteries using lithium nickel–cobalt composite oxide and graphite/coke hybrid carbon. Electrochim. Acta 2002, 47, 1691–1696. [Google Scholar] [CrossRef]
  3. Xiao, C.; Wang, H.; Usiskin, R.; van Aken, P.A.; Maier, J.J.S. Unification of insertion and supercapacitive storage concepts: Storage profiles in titania. Science 2024, 386, 407–413. [Google Scholar] [CrossRef]
  4. Shin, J.Y.; Samuelis, D.; Maier, J. Sustained lithium-storage performance of hierarchical, nanoporous anatase TiO2 at high rates: Emphasis on interfacial storage phenomena. Adv. Funct. Mater. 2011, 21, 3464–3472. [Google Scholar] [CrossRef]
  5. Li, M.; An, H.; Song, Y.; Liu, Q.; Wang, J.; Huo, H.; Lou, S.; Wang, J. Ion-Dipole-Interaction-Induced Encapsulation of Free Residual Solvent for Long-Cycle Solid-State Lithium Metal Batteries. J. Am. Chem. Soc. 2023, 145, 25632–25642. [Google Scholar] [CrossRef]
  6. Lou, S.; Liu, Q.; Zhang, F.; Liu, Q.; Yu, Z.; Mu, T.; Zhao, Y.; Borovilas, J.; Chen, Y.; Ge, M.; et al. Insights into interfacial effect and local lithium-ion transport in polycrystalline cathodes of solid-state batteries. Nat. Commun. 2020, 11, 5700. [Google Scholar] [CrossRef]
  7. Lou, S.; Yu, Z.; Liu, Q.; Wang, H.; Chen, M.; Wang, J. Multi-scale Imaging of Solid-State Battery Interfaces: From Atomic Scale to Macroscopic Scale. Chem 2020, 6, 2199–2218. [Google Scholar] [CrossRef]
  8. Zhang, Y.; Kang, C.; Zhao, W.; Song, Y.; Zhu, J.; Huo, H.; Ma, Y.; Du, C.; Zuo, P.; Lou, S.; et al. d-p Hybridization-Induced “Trapping-Coupling-Conversion” Enables High-Efficiency Nb Single-Atom Catalysis for Li-S Batteries. J. Am. Chem. Soc. 2023, 145, 1728–1739. [Google Scholar] [CrossRef]
  9. Zhang, Y.; Wang, Y.; Zhao, W.; Zuo, P.; Tong, Y.; Yin, G.; Zhu, T.; Lou, S. Delocalized electronic engineering of TiNb2O7 enables low temperature capability for high-areal-capacity lithium-ion batteries. Nat. Commun. 2024, 15, 6299. [Google Scholar] [CrossRef]
  10. Aravindan, V.; Sundaramurthy, J.; Jain, A.; Kumar, P.S.; Ling, W.C.; Ramakrishna, S.; Srinivasan, M.P.; Madhavi, S. Unveiling TiNb2O7 as an insertion anode for lithium ion capacitors with high energy and power density. ChemSusChem 2014, 7, 1858–1863. [Google Scholar] [CrossRef]
  11. Griffith, K.J.; Seymour, I.D.; Hope, M.A.; Butala, M.M.; Lamontagne, L.K.; Preefer, M.B.; Koçer, C.P.; Henkelman, G.; Morris, A.J.; Cliffe, M.J. Ionic and electronic conduction in TiNb2O7. J. Am. Chem. Soc. 2019, 141, 16706–16725. [Google Scholar] [CrossRef]
  12. Zhao, Z.; Xue, Z.; Xiong, Q.; Zhang, Y.; Hu, X.; Chi, H.; Qin, H.; Yuan, Y.; Ni, H. Titanium niobium oxides (TiNb2O7): Design, fabrication and application in energy storage devices. Sustain. Mater. Technol. 2021, 30, e00357. [Google Scholar] [CrossRef]
  13. He, Y.-B.; Li, B.; Liu, M.; Zhang, C.; Lv, W.; Yang, C.; Li, J.; Du, H.; Zhang, B.; Yang, Q.-H. Gassing in Li4Ti5O12-based batteries and its remedy. Sci. Rep. 2012, 2, 913. [Google Scholar] [CrossRef]
  14. Chen, L.; Yu, H.; Zhu, D.; Liu, S.; Zhang, L.; Sun, J.; Zhao, Z.; Li, Q.; Chen, G.; Li, Q. Designing electron/ion dual-phase conductor Ni@ TiO2 for high-performance lithium-ion storage: Combining insertion and space charge mechanism. Appl. Phys. Lett. 2024, 124, 133901. [Google Scholar] [CrossRef]
  15. Chen, Y.; Yang, J.; He, A.; Li, J.; Ma, W.; Record, M.-C.; Boulet, P.; Wang, J.; Albina, J.-M. Core–Double-Shell TiO2@ Fe3O4@ C Microspheres with Enhanced Cycling Performance as Anode Materials for Lithium-Ion Batteries. Materials 2024, 17, 2543. [Google Scholar] [CrossRef]
  16. Feng, F.; Wang, L.; Hu, N.; Wang, Q.; Hu, Q.; Guo, S.; Jin, W.; Li, C. The novel α-TiO2@ g-C3N4 heterostructure for ultra rapid ionic pumping and effective capacitive deionization. Sep. Purif. Technol. 2025, 355, 129632. [Google Scholar] [CrossRef]
  17. Pillai, A.M.; Gopinadh, S.V.; Phanendra, P.V.; Salini, P.S.; John, B.; SarojiniAmma, S.; Devassy, M.T. Bio-synthesized TiO2 nanoparticles and the aqueous binder-based anode derived thereof for lithium-ion cells. Discov. Nano 2024, 19, 69. [Google Scholar] [CrossRef]
  18. Wu, C.; Wang, X.; Zhu, Y.; Dong, L.; Xu, J. SN-doping TiO2@ MXene heterostructure in-situ derived from MXene frameworks as high-rate anodes for Lithium/Sodium-ion batteries. J. Electroanal. Chem. 2025, 977, 118833. [Google Scholar] [CrossRef]
  19. Wu, W.; Lin, Y.; Hu, Y.; He, Z.; Yang, Y. Phase-field modelling for degradation/failure research in lithium battery: Progress and prospects. J. Energy Chem. 2024, 102, 628–650. [Google Scholar] [CrossRef]
  20. Dahlman, C.J.; Heo, S.; Zhang, Y.; Reimnitz, L.C.; He, D.; Tang, M.; Milliron, D.J. Dynamics of lithium insertion in electrochromic titanium dioxide nanocrystal ensembles. J. Am. Chem. Soc. 2021, 143, 8278–8294. [Google Scholar] [CrossRef]
  21. Fan, M.; Yang, Z.; Lin, Z.; Xiong, X. Facile synthesis of uniform N-doped carbon-coated TiO2 hollow spheres with enhanced lithium storage performance. Nanoscale 2021, 13, 2368–2372. [Google Scholar] [CrossRef]
  22. Kim, M.-S.; Lee, B.-H.; Park, J.-H.; Lee, H.S.; Hooch Antink, W.; Jung, E.; Kim, J.; Yoo, T.Y.; Lee, C.W.; Ahn, C.-Y. Operando identification of the chemical and structural origin of Li-ion battery aging at near-ambient temperature. J. Am. Chem. Soc. 2020, 142, 13406–13414. [Google Scholar] [CrossRef]
  23. Lee, M.D.; Lee, G.J.; Nam, I.; Abbas, M.A.; Bang, J.H. Exploring the effect of cation vacancies in TiO2: Lithiation behavior of n-type and p-type TiO2. ACS Appl. Mater. Interfaces 2022, 14, 6560–6569. [Google Scholar] [CrossRef]
  24. Liu, Y.; Pan, X.; Chen, W.; Zhao, X. Titanate-derived Nb-doped TiO2 nanoparticles displaying improved lithium storage performance. Dalton Trans. 2022, 51, 2506–2511. [Google Scholar] [CrossRef]
  25. Babu, B.; Simon, P.; Balducci, A. Fast charging materials for high power applications. Adv. Energy Mater. 2020, 10, 2001128. [Google Scholar] [CrossRef]
  26. Hao, Z.; Chen, Q.; Dai, W.; Ren, Y.; Zhou, Y.; Yang, J.; Xie, S.; Shen, Y.; Wu, J.; Chen, W. Oxygen-deficient blue TiO2 for ultrastable and fast lithium storage. Adv. Energy Mater. 2020, 10, 1903107. [Google Scholar] [CrossRef]
  27. Liang, Y.; Xiong, X.; Xu, Z.; Xia, Q.; Wan, L.; Liu, R.; Chen, G.; Chou, S.L. Ultrathin 2D Mesoporous TiO2/rGO Heterostructure for High-Performance Lithium Storage. Small 2020, 16, 2000030. [Google Scholar] [CrossRef]
  28. Niu, J.; Zhang, Z.; Aurbach, D. Alloy anode materials for rechargeable Mg ion batteries. Adv. Energy Mater. 2020, 10, 2000697. [Google Scholar] [CrossRef]
  29. Liu, J.; Xu, Z.; Wu, M.; Wang, Y.; Karim, Z. Capacity contribution induced by pseudo-capacitance adsorption mechanism of anode carbonaceous materials applied in potassium-ion battery. Front. Chem. 2019, 7, 640. [Google Scholar] [CrossRef]
  30. Zhang, W.; Zhang, Y.; Yu, L.; Wu, N.-L.; Huang, H.; Wei, M. TiO2-B nanowires via topological conversion with enhanced lithium-ion intercalation properties. J. Mater. Chem. A 2019, 7, 3842–3847. [Google Scholar] [CrossRef]
  31. Ha, J.U.; Lee, J.; Abbas, M.A.; Lee, M.D.; Lee, J.; Bang, J.H. Designing hierarchical assembly of carbon-coated TiO2 nanocrystals and unraveling the role of TiO2/carbon interface in lithium-ion storage in TiO2. ACS Appl. Mater. Interfaces 2019, 11, 11391–11402. [Google Scholar] [CrossRef]
  32. Jiang, H.; Wei, Z.; Cai, X.; Lai, L.; Ma, J.; Huang, W. A cathode for Li-ion batteries made of vanadium oxide on vertically aligned carbon nanotube arrays/graphene foam. Chem. Eng. J. 2019, 359, 1668–1676. [Google Scholar] [CrossRef]
  33. Lee, D.-H.; Lee, B.-H.; Sinha, A.K.; Park, J.-H.; Kim, M.-S.; Park, J.; Shin, H.; Lee, K.-S.; Sung, Y.-E.; Hyeon, T. Engineering titanium dioxide nanostructures for enhanced lithium-ion storage. J. Am. Chem. Soc. 2018, 140, 16676–16684. [Google Scholar] [CrossRef] [PubMed]
  34. Lou, S.; Zhao, Y.; Wang, J.; Yin, G.; Du, C.; Sun, X. Ti-based oxide anode materials for advanced electrochemical energy storage: Lithium/sodium ion batteries and hybrid pseudocapacitors. Small 2019, 15, 1904740. [Google Scholar] [CrossRef]
  35. Cheng, Q.; Liang, J.; Zhu, Y.; Si, L.; Guo, C.; Qian, Y. Bulk Ti2Nb10O29 as long-life and high-power Li-ion battery anodes. J. Mater. Chem. A 2014, 2, 17258–17262. [Google Scholar] [CrossRef]
  36. Wagemaker, M.; Mulder, F.M. Properties and promises of nanosized insertion materials for Li-ion batteries. Acc. Chem. Res. 2013, 46, 1206–1215. [Google Scholar] [CrossRef] [PubMed]
  37. Mei, J.; Liao, T.; Kou, L.; Sun, Z. Oxygen vacancies and cation disorder tailoring electrochemical kinetics in TiNb2O7. Adv. Energy Mater. 2021, 11, 2101713. [Google Scholar]
  38. Kim, J.-H.; Park, K.-J.; Kim, S.-J.; Yoon, C.S. Atomic-scale engineering of cation channels for fast-charging batteries. Nat. Energy 2022, 7, 312–321. [Google Scholar]
Figure 1. Structural and morphological characterization of the samples: scanning electron microscope (SEM) images of TNO-1100 (a,b), TNO-1200 (c,d), and TNO-1400 (e,f) at low and high magnifications, respectively.
Figure 1. Structural and morphological characterization of the samples: scanning electron microscope (SEM) images of TNO-1100 (a,b), TNO-1200 (c,d), and TNO-1400 (e,f) at low and high magnifications, respectively.
Micromachines 16 00490 g001
Figure 2. X-ray diffraction patterns of TNO-1100, TNO-1200, and TNO-1400 samples with reference diffraction lines (JCPDS 77-1374) for monoclinic TiNb2O7 (space group: C2/m). Characteristic peaks are indexed. Measurements were performed using Cu Kα radiation (λ = 0.154056 nm) with a scanning range of 2θ = 5–90° at 5°/min.
Figure 2. X-ray diffraction patterns of TNO-1100, TNO-1200, and TNO-1400 samples with reference diffraction lines (JCPDS 77-1374) for monoclinic TiNb2O7 (space group: C2/m). Characteristic peaks are indexed. Measurements were performed using Cu Kα radiation (λ = 0.154056 nm) with a scanning range of 2θ = 5–90° at 5°/min.
Micromachines 16 00490 g002
Figure 3. According to the results, the TNO-1200 electrode shows higher specific capacities than the TNO-1100 electrode and TNO-1400 electrode under the following conditions: current with 10 C; voltage window of 0.01–3.0 V after 3000 cycles.
Figure 3. According to the results, the TNO-1200 electrode shows higher specific capacities than the TNO-1100 electrode and TNO-1400 electrode under the following conditions: current with 10 C; voltage window of 0.01–3.0 V after 3000 cycles.
Micromachines 16 00490 g003
Figure 4. SEM images (a), SEM-EDS mapping (bd), and line scan EDS (e) of the elemental distribution of TNO-1200.
Figure 4. SEM images (a), SEM-EDS mapping (bd), and line scan EDS (e) of the elemental distribution of TNO-1200.
Micromachines 16 00490 g004
Figure 5. Rietveld refinement pattern of all samples generated using FullProf Software (Version 2.05), (a) TNO-1100, TNO-1200, and TNO-1400. (b) A reference crystal structure of sample TNO-1200 generated from refinement data.
Figure 5. Rietveld refinement pattern of all samples generated using FullProf Software (Version 2.05), (a) TNO-1100, TNO-1200, and TNO-1400. (b) A reference crystal structure of sample TNO-1200 generated from refinement data.
Micromachines 16 00490 g005
Table 1. Refined structural parameters (lattice constants a, b, and c, calculated unit cell volume V and atomic positions) by the Fullprof Suite from the XRD data measured at 300 K. Number in parenthesis is the estimated standard deviation of the last or the next last significant digit.
Table 1. Refined structural parameters (lattice constants a, b, and c, calculated unit cell volume V and atomic positions) by the Fullprof Suite from the XRD data measured at 300 K. Number in parenthesis is the estimated standard deviation of the last or the next last significant digit.
TiNb2O7 (Monoclinic, Space Group: C2/m)
T (K)300
a, b, c (Å)17.6869 (2)3.8030 (0)11.8976 (0)
α, β, γ (°)9095.32233 (3)90
V3)796.830 (1)
Atomxyz
Nb10.0000 (0)0.0000 (0)0.0000 (0)
Ti10.0000 (0)0.0000 (0)0.0000 (0)
Nb20.18483 (36)0.0000 (0)0.18442 (58)
Ti20.18483 (36)0.0000 (0)0.18442 (58)
Nb30.07844 (33)0.0000 (0)−0.55762 (56)
Ti30.07844 (33)0.0000 (0)−0.55762 (56)
Nb40.89020 (36)0.0000 (0)0.25427 (59)
Ti40.89020 (36)0.0000 (0)0.25427 (59)
Nb50.29626 (43)0.0000 (0)−0.07822 (66)
Ti50.29626 (43)0.0000 (0)−0.07822 (66)
O10.18303 (0)0.0000 (0)−0.41559 (0)
O20.36568 (0)0.0000 (0)−0.25112 (0)
O30.59550 (0)0.0000 (0)−0.01283 (0)
O40.78760 (0)0.0000 (0)0.16717 (0)
O50.24841 (0)0.0000 (0)0.05577 (0)
O60.69940 (0)0.0000 (0)0.69177 (0)
O70.89633 (0)0.0000 (0)−0.08214 (0)
O80.01655 (0)0.0000 (0)−0.38232 (0)
O90.87235 (0)0.0000 (0)0.68015 (0)
O100.50000 (0)0.0000 (0)0.50000 (0)
O110.04006 (0)0.0000 (0)−0.13948 (0)
Rp, Rwp, Rexp, χ25.08, 8.30, 3.68, 5.09
Table 2. Electrochemical performance comparison of TNO-1200 with other anode materials under fast-charging conditions.
Table 2. Electrochemical performance comparison of TNO-1200 with other anode materials under fast-charging conditions.
MaterialCurrent DensityCapacity (mAh/g)Cycle NumberRetention Rate
TNO-1200 (This Work) 10 C125300098%
TiNb2O7 [11]1 C250100-
Ti2Nb10O29 [35]10 C10050095%
Oxygen-deficient TNO [26]5 C180100085%
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hu, X.; Zhong, Y.; Hu, X.; Feng, X.; Ye, F. Rapid Synthesis of Fast-Charging TiNb2O7 for Lithium-Ion Storage via Ultrafast Carbothermal Shock. Micromachines 2025, 16, 490. https://doi.org/10.3390/mi16050490

AMA Style

Hu X, Zhong Y, Hu X, Feng X, Ye F. Rapid Synthesis of Fast-Charging TiNb2O7 for Lithium-Ion Storage via Ultrafast Carbothermal Shock. Micromachines. 2025; 16(5):490. https://doi.org/10.3390/mi16050490

Chicago/Turabian Style

Hu, Xianyu, Yunlei Zhong, Xiaosai Hu, Xiyuan Feng, and Fengying Ye. 2025. "Rapid Synthesis of Fast-Charging TiNb2O7 for Lithium-Ion Storage via Ultrafast Carbothermal Shock" Micromachines 16, no. 5: 490. https://doi.org/10.3390/mi16050490

APA Style

Hu, X., Zhong, Y., Hu, X., Feng, X., & Ye, F. (2025). Rapid Synthesis of Fast-Charging TiNb2O7 for Lithium-Ion Storage via Ultrafast Carbothermal Shock. Micromachines, 16(5), 490. https://doi.org/10.3390/mi16050490

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop