Robust assembly of TiO2 quantum dots onto Ti3C2Tx for excellent lithium storage capability

Xinlin Zhang Cheng Tang Haitao Li Jie Sun Aijun Du Minghong Wu Haijiao Zhang

Citation:  Xinlin Zhang, Cheng Tang, Haitao Li, Jie Sun, Aijun Du, Minghong Wu, Haijiao Zhang. Robust assembly of TiO2 quantum dots onto Ti3C2Tx for excellent lithium storage capability[J]. Chinese Chemical Letters, 2025, 36(6): 110088. doi: 10.1016/j.cclet.2024.110088 shu

Robust assembly of TiO2 quantum dots onto Ti3C2Tx for excellent lithium storage capability

English

  • To date, lithium-ion batteries (LIBs) stand out as one of the most dependable and practical devices owing to their high energy/power density and eco-friendliness [1]. However, the graphite anode utilized by LIBs suffers from low capacity, which brings new challenges in meeting the increasing demand for high-performance energy storage in rapidly evolving electronic devices [2]. Therefore, it is necessary to further develop high-performance LIBs anode materials with high capacity and long cycling life.

    Of these anode materials, titanium dioxide (TiO2) has been regarded as a promising anode material for LIBs, owing to its good structural stability and small volume changes (-5%) [3], in contrast to other metal oxides such as Nb2O5 [4], and MoO3 [5]. These features can effectually improve the cycling stability of the TiO2 electrode and prolong its operational lifespan. However, its inherent low conductivity and slow electrochemical reaction kinetics remain obstacles for achieving the high lithium storage performance. To address these issues, various approaches have been developed. For example, nanostructured engineering of TiO2 materials is one of the most effective strategies. To date, nanosheet [6,7], nanowire [3], and nanotube [8] structured TiO2 have been successfully prepared, demonstrating an improved electrochemical performance [9]. On the other hand, the conductivity of TiO2 can be markedly enhanced by combining it with some two-dimensional (2D) conductive matrix. Specifically, Ti3C2Tx MXene exhibits distinct advantages in lithium-ion storage owing to exceptional conductivity and unique 2D structure [10,11]. Consequently, incorporating TiO2 nanoparticles onto Ti3C2Tx holds significant promise for boosting their lithium storage performance. For instance, Shakoor et al. [12] loaded TiO2 particles onto the Ti3C2Tx surface, resulting in a specific discharge capacity of 200 mAh/g at 0.1 C. Additionally, Li et al. [13] prepared the Ti3C2@TiO2 MXene hybrid and exhibited a capacity of 302 mAh/g after 500 cycles at 200 mA/g. Nevertheless, the obtained TiO2 particles generally have large size and unstable combination between TiO2 and Ti3C2Tx, leading to the dissatisfactory electrochemical performance. As a unique nanostructure, zero-dimensional (0D) quantum dots (QDs) have garnered growing attention in LIBs due to their maximized surface area exposure and minimized charge diffusion distance [14,15]. More importantly, QDs can effectively buffer large volume expansion during ion insertion/extraction. Besides, the SiO2 with excellent rigidity has been demonstrated to be an effective stabilizer for the composite materials in previous study [16]. Therefore, exploring the facile method for robust assembly of TiO2 QDs onto Ti3C2Tx nanosheets by the incorporation of SiO2 particles is anticipated to significantly enhance the lithium storage performance of the composite, especially its cycling stability.

    In this work, we design a new 0D/2D heterostructure (namely, A-TiO2/Ti3C2Tx) through a scalable hydrothermal process induced by (3-aminopropyl)triethoxysilane (APTES). The results indicate that the SiO2 nanoparticles produced by the hydrolysis of APTES contribute to a strong coupling effect between TiO2 QDs and Ti3C2Tx nanosheets. Such a unique configuration endows the A-TiO2/Ti3C2Tx anode with an extremely high capacity and long-term cycling life for lithium storage. Furthermore, in-situ electrochemical impedance spectroscopy (EIS) indicates that the remarkable structural stability of A-TiO2/Ti3C2Tx plays a crucial role in enabling the electrochemical performance. In addition, the theoretical calculation indicates that the in-situ N-doping derived from APTES into the composite also enhances the adsorption of Li ions.

    Fig. 1a illustrates the synthesis process of A-TiO2/Ti3C2Tx. Initially, Ti3C2Tx nanosheets were synthesized according to our recent work [10]. Then, positive APTES was adsorbed onto the negative Ti3C2Tx surface via electrostatic interactions between them. As confirmed by the zeta potential (Fig. S1 in Supporting information), the potential changes from −23 mV for pure Ti3C2Tx to 10 mV after the APTES modification. Along with the hydrolysis of APTES, an ultrathin SiO2 layer was deposited onto the Ti3C2Tx surface, which was further validated by the XRD pattern (Fig. S2 in Supporting information). The broad peak around 23° is ascribed to the SiO2 characteristic [17]. Moreover, the TEM images also clearly observe the presence of SiO2 layer (Fig. S3a in Supporting information) compared to pristine Ti3C2Tx (Fig. S3b in Supporting information). Subsequently, in a mixed ethanol-glycerol solvent, Ti4+ underwent gradual hydrolysis and subsequent hydrothermal treatment, forming a uniform TiO2 QDs structure onto Ti3C2Tx (Fig. S4 in Supporting information). Ultimately, after heat treatment at 600 ℃, the A-TiO2/Ti3C2Tx composite with a well-defined morphology is obtained.

    Figure 1

    Figure 1.  (a) The fabrication process of A-TiO2/Ti3C2Tx. (b, c) SEM images, (d) TEM image, (e, f) HRTEM images (insert is the SAED pattern), (g) particle size distribution profile, and (h) STEM image and corresponding elemental mappings of A-TiO2/Ti3C2Tx.

    Scanning electron microscopy (SEM) and transmission electron microscopy (TEM) technique were used to observe the morphology and structure of A-TiO2/Ti3C2Tx. Seen from SEM images (Figs. 1b and c), the composite maintains the lamellar structure of Ti3C2Tx, and a lot of TiO2 nanoparticles are uniformly anchored onto the Ti3C2Tx surface. The TEM image further confirms the formation of this composite (Fig. 1d). In contrast, the TiO2/Ti3C2Tx sample exhibits a uniform morphology before heat treatment (Fig. S5a in Supporting information). However, after calcination, its whole structure is obviously damaged (Fig. S5b in Supporting information), indicating a poor structural stability. This result indicates that the formed SiO2 layer serves as a "glue" for tightly coupling the TiO2 QDs with Ti3C2Tx. That further enhances the structural stability of A-TiO2/Ti3C2Tx, suggesting a superior cycling stability for lithium storage. Figs. 1e and f show the HRTEM images of A-TiO2/Ti3C2Tx. Clearly, the TiO2 nanoparticles are highly dispersed into Ti3C2Tx. Fig. 1f displays the lattice spacing of 0.352 nm, aligning with the anatase (100) crystal face of TiO2 [18]. The particle size distribution in Fig. 1g indicates that the TiO2 QDs size predominantly fell within the range of 1.5–3.5 nm. Additionally, STEM image and energy dispersive spectroscopy (EDS) mapping manifest the uniform distribution of Ti, O, Si, C, and N elements in the composite (Fig. 1h).

    Fig. 2a presents the X-ray diffraction (XRD) patterns of two composites. A-TiO2/Ti3C2Tx shows the broad peak around 23°, corresponding to the characteristic peaks of SiO2 [17], while peaks at 25.3° and 62.7° are assigned to the crystal faces (100) and (204) of anatase TiO2, indicating the formation of TiO2 onto the Ti3C2Tx. Differently, no characteristic peaks of SiO2 can be observed in TiO2/Ti3C2Tx, in accordance with the experimental observations [19]. In Raman spectrum (Fig. 2b), the characteristic peaks at 200 cm−1 derived from Ti3C2Tx can be visible in three samples. Moreover, the characteristic peaks of TiO2 at 394, 513 and 635 cm−1 are well-indexed in A-TiO2/Ti3C2Tx and TiO2/Ti3C2Tx [12]. Besides, the chemical bonds of two composites were further analyzed via FT-IR spectra (Fig. 2c). A-TiO2/Ti3C2Tx and TiO2/Ti3C2Tx both displays the absorption vibration peaks C—F (1097 cm−1) and Ti—O (620 cm−1) belonging to Ti3C2Tx. However, the absorption vibration peaks of C—N (1050 cm−1) and Si—O (809 cm−1) bonds are observed in A-TiO2/Ti3C2Tx [20], suggesting the existence of SiO2 originated from the hydrolysis of APTES, which is crucial for the formation of TiO2 QDs. Additionally, the appearance of C—N (1670 cm−1) and Si—O—Si (459 cm−1) bonds on the SiO2/Ti3C2Tx surface also confirms the point (Fig. S6 in Supporting information) [21]. Based on the N2 sorption isotherms in Fig. 2d and Fig. S7a (Supporting information), the BET specific surface areas are about 110.1 m2/g and 100.2 m2/g for A-TiO2/Ti3C2Tx and TiO2/Ti3C2Tx, respectively. The larger surface areas of A-TiO2/Ti3C2Tx can provide the more active sites for lithium storage, suggesting a better performance. Furtherly, two composites both illustrate the mesoporous characteristics (inset of Fig. 2d and Fig. S7b in Supporting information).

    Figure 2

    Figure 2.  (a) XRD patterns, (b) Raman spectra, and (c) FT-IR spectra of A-TiO2/Ti3C2Tx and TiO2/Ti3C2Tx. (d) N2 sorption isotherm and corresponding pore size distribution curve, and high-resolution XPS spectra of A-TiO2/Ti3C2Tx for (e) Ti 2p, (f) O 1s, (g) C 1s, (h) Si 2p and (i) N 1s.

    X-ray photoelectron spectrometer (XPS) technique was applied to analyze the chemical states of A-TiO2/Ti3C2Tx. The survey XPS spectrum confirms the existence of Ti, Si, O, C, and N (Fig. S8a in Supporting information), consistent with the above EDS analysis. And the N atomic content is 1.29% (Fig. S8b in Supporting information), which mainly comes from the containg-NH2 APTES. Fig. 2e shows the high-resolution spectrum of Ti 2p, the peaks at 459.33 and 465.28 eV are attributed to Ti4+, while the doublet peaks at 456.49 and 464.29 eV are associated with the Ti-C bond. Meanwhile, the Ti-O bond at 459.05 eV is also observed. This further verifies the strong coupling of TiO2 QDs with Ti3C2Tx. In the O 1s region, Si—O, Ti—O—C and Ti—O—Ti bonds can be well identified at the binding energy of 530.5, 531.09 and 532.41 eV (Fig. 2f) [6]. Two peaks at 284.72 eV and 286.08 eV in the C 1s spectrum (Fig. 2g) are ascribed to C—C (284.72 eV) and C-Ti/C—N/C—O [22]. The Si 2p spectrum in Fig. 2h displays distinct Si4+ (103 eV) and Si-O (102.19 eV) peaks [21], corresponding to the spectral analysis of O 1s. The N 1s high-resolution spectrum in Fig. 2i shows the presence of pyridinic N (398.3 eV), pyrrolic N (399.8 eV), and graphitic N (401 eV) [23], which can effectively enhance the adsorption of Li ions by heterostructure, thereby promoting the lithium storage capacity of the composite electrode.

    To evaluate the lithium storage performance of two composites, a half cell was assembled using lithium metal as the counter electrode. Fig. 3a illustrates the cyclic voltammetry (CV) curves for the first three cycles of the A-TiO2/Ti3C2Tx electrode at a scan rate of 0.1 mV/s. In the first cycle, a cathodic peak appears around 0.75 V, and an anodic peak is observed near 0.95 V, indicating the Li+ insertion/extraction in the composite [24]. Additionally, an irreversible cathode peak at around 1.75 V and an anode peak at approximately 2.02 V are seen, attributing to the electrochemical activities of TiO2. The irreversible cathodic peak is related to the formation of a solid electrolyte interface (SEI) [25]. In the following two cycles, the irreversible peaks are significantly reduced, and the curves almost completely overlap. Furthermore, in comparison to TiO2/Ti3C2Tx (Fig. S9a in Supporting information), the CV curve of the A-TiO2/Ti3C2Tx electrode exhibits a more rectangular shape, indicating a superior reversibility during cycling process [26].

    Figure 3

    Figure 3.  (a) CV curves of initial three cycles at a scan rate of 0.1 mV/s, and (b) charge-discharge profiles of the A-TiO2/Ti3C2Tx electrode at 100 mA/g. (c) Rate capabilities, (d) cycling performances at 100 mA/g, and (e) comparison of cycle performances between the A-TiO2/Ti3C2Tx electrode and other TiO2-based anodes for LIBs. (f) Long cycling performances of A-TiO2/Ti3C2Tx and TiO2/Ti3C2Tx electrodes at a high current density of 1000 mA/g.

    The galvanostatic charge/discharge (GDC) curves for the first three cycles of A-TiO2/Ti3C2Tx and TiO2/Ti3C2Tx electrodes are presented in Fig. 3b and Fig. S9b (Supporting information). In the first cycle, the discharge and charge capacities of the A-TiO2/Ti3C2Tx electrode are 820.9 mAh/g and 376.9 mAh/g, respectively, while the TiO2/Ti3C2Tx electrode exhibits a lower initial discharge capacity of only 662.4 mAh/g. Moreover, in subsequent cycles, the A-TiO2/Ti3C2Tx electrode remains a higher capacity, demonstrating an enhanced lithium-storage performance.

    The rate performance of two electrodes was shown in Fig. 3c. At the current densities of 0.05, 0.1, 0.2, 0.5, 1.0, and 2.0 A/g, the A-TiO2/Ti3C2Tx electrode delivers the specific capacities of 415.2, 375.8, 348.1, 315.5, 278.9, and 224.3 mAh/g, respectively. Furtherly, when the current density is returned to 0.1 A/g, the capacity increases to 420.4 mAh/g. In contrast, the TiO2/Ti3C2Tx electrode has lower capacities of 370.9, 332.7, 298.3, 213.2, 123.6, and 54.1 mAh/g at the same current densities. Upon returning to 0.1 A/g, only 308 mAh/g of the capacity can be maintained. These results indicate that such a unique heterostructure can promote the lithium ion transport in the electrochemical reaction.

    Fig. 3d illustrates the cycle performances of A-TiO2/Ti3C2Tx and TiO2/Ti3C2Tx electrodes at a current density of 0.1 A/g. Clearly, the A-TiO2/Ti3C2Tx electrode has a better cycle performance than that TiO2/Ti3C2Tx electrode. More importantly, it maintains a high specific capacity of 424.7 mAh/g, corresponding to 126% of the theoretical capacity of TiO2 (335 mAh/g) after 400 cycles. In comparison, the TiO2/Ti3C2Tx electrode only sustains a capacity of 176 mAh/g mAh/g under the same conditions. The outstanding cycling performance of A-TiO2/Ti3C2Tx can be related to its superior structural stability facilitated by the presence of SiO2 during cycling. Moreover, as listed in Fig. 3e and Table S1 (Supporting information), the newly-developed A-TiO2/Ti3C2Tx electrode in our work is comparable to those TiO2-based anodes reported [3,27-34].

    Impressively, the A-TiO2/Ti3C2Tx anode demonstrates an exceptional long cycling performance. As shown in Fig. 3f, after 2500 cycles, it maintains a specific capacity of 170 mAh/g at a high current density of 1000 mA/g, achieving an average coulombic efficiency of nearly 99.99%. The rate of capacity loss per cycle was a mere 0.009%. In stark contrast, the TiO2/Ti3C2Tx electrode preserves a capacity of only 100 mAh/g after 2500 cycles at the same current density. This contrast underscores the enhanced durability and performance of the A-TiO2/Ti3C2Tx electrode, attributed to the strong coupling effect of TiO2 QDs and Ti3C2Tx nanosheets.

    To further demonstrate the N-doping advantages, the contact angles of two electrodes in the electrolytes were tested (Fig. S10 in Supporting information). Obviously, the wettability of the A-TiO2/Ti3C2Tx electrode is superior to that of the TiO2/Ti3C2Tx electrode, suggesting that A-TiO2/Ti3C2Tx has more fully contact with the electrolyte during the electrochemical reaction. This enhanced wettability can be ascribed to the nitrogen introduction. Moreover, the TEM images after cycles of two electrodes (Fig. S11 in Supporting information) further demonstrate that the existence of thin SiO2 layer in the heterostructure can ensure the stability of A-TiO2/Ti3C2Tx during the long-term cycles, so as to obtain the superior long cycle life.

    A series of CV tests were carried out at various scan rates to investigate the transfer behavior of two electrodes. As illustrated in Fig. 4a and Fig. S12a (Supporting information), their curves display comparable profiles and trends. Furthermore, the presence of redox peaks points to rapid electrochemical reactions occurring within the materials. The relationship between the peak current and the scanning rate (ν) determine the material's capacitance and diffusion contributions [35], and the two obey the following relationship:

    (1)

    (2)

    where a and b are constants, and the fitted value of b () can indicate whether the electrode dynamic process is primarily governed by capacitance control (b-1.0) or diffusion control (b-0.5) [36]. In Fig. 4b, the fitting b values for the cathode and anode peaks of A-TiO2/Ti3C2Tx are 0.90 and 0.84, respectively, suggesting a capacitance-controlled electrochemical process. Conversely, for TiO2/Ti3C2Tx electrode, the fitted b values for the cathode and anode peaks are 0.61 and 0.56, respectively, indicating a diffusion-dominated behavior (Fig. S12b in Supporting information). The contribution of capacitance and diffusion control to capacity is analyzed by using the following formula [37]:

    (3)

    Figure 4

    Figure 4.  (a) CV curves at various scan rates, (b) the calculated b values at six redox peaks, (c) surface capacitive contribution at 1.0 mV/s, and (d) the percentage of capacitive contribution of the A-TiO2/Ti3C2Tx electrode at diverse scan rates. (e) GITT curves of A-TiO2/Ti3C2Tx and TiO2/Ti3C2Tx electrodes and their (f) Li+ diffusion coefficients. In-situ EIS diagram of (g) discharge and (h) charge of A-TiO2/Ti3C2Tx, and calculated Rct resistance values for (i) discharge and (j) charge.

    where k1ν and k2ν represent non-Faraday reactions and Faraday reactions, respectively. As depicted in Fig. 4c and Fig. S13a (Supporting information), at a scan rate of 1.0 mV/s, the capacitance contribution ratio is found to be 74.8% for A-TiO2/Ti3C2Tx and 66.2% for TiO2/Ti3C2Tx, respectively. Moreover, the capacitance contribution values were computed for multiple sweep speeds simultaneously. As illustrated in Fig. 4d and Fig. S13b (Supporting information), the A-TiO2/Ti3C2Tx electrode exhibits a higher capacitance contribution rate across all scanning rates than TiO2/Ti3C2Tx. That indicates its better transport dynamics, thereby improving the rate performance of A-TiO2/Ti3C2Tx.

    Fig. 4e presents the galvanostatic intermittent titration technique (GITT) curves of A-TiO2/Ti3C2Tx and TiO2/Ti3C2Tx electrodes. The Li+ diffusion coefficients of two electrodes were calculated based on Fick's second law and the following equation [38]:

    (4)

    As shown in Fig. 4f, the DLi+ values of two electrodes are relatively similar [39]. However, the overall DLi+ levels of the A-TiO2/Ti3C2Tx electrode is higher than those of TiO2/Ti3C2Tx. These results further confirm that the A-TiO2/Ti3C2Tx electrode exhibits an enhanced Li+ diffusion ability, thus effectively improving the energy storage performance.

    In-situ electrochemical impedance spectroscopy (EIS) was also conducted to analysis the impedance changes of A-TiO2/Ti3C2Tx electrodes during the first cycle. The data was fitted with a suitable circuit, yielding the 2D Nyquist diagram, as depicted in Figs. 4g and h (where the line graph represents the fitted data). With the discharge to 1.7–1.0 V, the SEI film gradually forms, Nyquist curve in the intermediate frequency region is semicircular, indicating the resistance value of the SEI film [40]. The semicircle of the Rct resistance region at this point also tends to stabilize (Fig. 4g). In the final process, two semicircular resistors are stabilized as RSEI and Rct, indicating the complete formation of the SEI film. In the subsequent charging process (Fig. 4h), the RSEI resistance tends to stabilize, and only the Rct resistance region gradually increases until it appears as a straight line. This change indicates the formation of the stable SEI film on the material's surface [41]. Additionally, based on the fitting results, the changes in Rct during the charge and discharge processes were plotted (Figs. 4i and j), in line with the changes observed in the Nyquist curve and confirmed the above changes in Rct.

    Then, we calculate the adsorption energy and surface charge density difference to show the enhancement of the Li adsorption and surface electron redistribution in the TiO2/Ti3C2N2 heterostructure compared to TiO2/Ti3C2, Ti3C2N2 and TiO2 surface. Two adsorption sites at the TiO2 (Ⅰ for Ti-top and Ⅱ for O-top) and Ti3C2 surface (Ⅰ for C-hollow, Ⅱ for Ti-hollow) were choose, respectively, to identity the more stable one for Li adsorption. As presented in Fig. 5a, the Li adsorption energy on the TiO2/Ti3C2N2 reaches −6.73 eV at the C-hollow site (Ⅰ), significantly lower than those on TiO2/Ti3C2 (−1.06 eV), Ti3C2N2 (−4.04 eV) and TiO2 surface (−3.21 eV), manifesting the enhanced Li adsorption stability in the TiO2/Ti3C2N2. Besides, the charge density difference shows that the electrons preferred to deplete from the Li-ion and accumulate at the surface of the substrate, suggesting the electron transfer from the Li to the Ti3C2 or TiO2 surface. The Bader analysis exhibits the 0.38e, 0.07e, 0.35e and 0.26e electrons transfer from the Li to the TiO2/Ti3C2N2, TiO2/Ti3C2, Ti3C2N2 and TiO2 surface, respectively (Fig. 5b). Herein, the TiO2/Ti3C2N2 heterostructure attracts more electrons from the Li-ion, rendering it a stronger interaction on Li adsorption. Therefore, the enhanced adsorption of Li ions makes the TiO2/Ti3C2N2 heterostructure a promising electrode material for LIBs.

    Figure 5

    Figure 5.  (a) Lithium adsorption energies (Eads) at different sites (Ⅰ and Ⅱ) for TiO2/Ti3C2N2, TiO2/Ti3C2, Ti3C2N2 and TiO2 surface, respectively. (b) Corresponding charge density difference (iso-value of 0.002 e/Å3) for each surface at their most stable site. (c) Schematic diagram of lithium storage merits of A-TiO2/Ti3C2Tx.

    Based on the above results, the outstanding lithium storage performance of the A-TiO2/Ti3C2Tx electrode is mainly due to the following reasons (Fig. 5c): Firstly, the construction of 0D/2D heterostructure shortens the ion transport path, thus promoting the rapid transfer dynamics during cycles. Secondly, Ti3C2Tx nanosheets as the substrate establishes an excellent conductive network, effectively enhancing the conductivity of the whole electrode, and ultrasmall TiO2 QDs provide the exposure of more active sites for lithium-ion storage, thereby boosting the reversible capacity. Furthermore, the N atoms derived from APTES enhance the adsorption of lithium ions, leading to increased lithium storage. More importantly, the SiO2 layer ensures the tight attachment of TiO2 QDs onto the Ti3C2Tx surface, significantly enhancing the structural integrity of the overall composite during the long cycling.

    In summary, a new type of 0D TiO2 quantum dots/2D Ti3C2Tx composite with exceptional lithium storage performance has been constructed using the APTES-induced assembly route. The results indicate that the SiO2 layer formed from the APTES hydrolysis significantly improves the structural stability of the whole electrode. The prepared A-TiO2/Ti3C2Tx anode maintains a high capacity of 425.4 mAh/g after 400 cycles at 100 mA/g, surpassing the theoretical capacity of original TiO2 by 126%. Furthermore, it has an impressive long-cycling performance, retaining a capacity of 170 mAh/g after 2500 cycles at 1000 mA/g. Through electrochemical kinetics analysis combining with DFT calculation, the excellent lithium storage performance of A-TiO2/Ti3C2Tx is attributed to its unique heterostructure and in-situ N doping derived from APTES, which not only reduces the Li+ adsorption energy, but also gives the fast charge transfer dynamics. The strategy proposed in our work offers a promising approach towards high-performance nanocomposites with enhanced structural stability.

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

    Xinlin Zhang: Investigation, Methodology, Writing – original draft. Cheng Tang: Software. Haitao Li: Validation. Jie Sun: Data curation. Aijun Du: Validation. Minghong Wu: Funding acquisition. Haijiao Zhang: Conceptualization, Project administration, Supervision, Writing – review & editing.

    We appreciate the support from the Natural Science Foundation of Shanghai (No. 23ZR1423800), and Key Laboratory of Advanced Energy Materials Chemistry (Ministry of Education), Nankai University.

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.110088.


    1. [1]

      X. Xiao, Z.J. Zhang, W.T. Yu, et al., Appl. Energy 328 (2022) 120186. doi: 10.1016/j.apenergy.2022.120186

    2. [2]

      F.Y. He, C. Tang, Y.D. Liu, et al., J. Mater. Sci. Technol. 100 (2022) 101–109. doi: 10.1016/j.jmst.2021.05.048

    3. [3]

      L. Luo, K. Liang, Z. Khanam, et al., Small 20 (2024) 2307103. doi: 10.1002/smll.202307103

    4. [4]

      C.F. Zhang, M. Beidaghi, M. Naguib, et al., Chem. Mater. 28 (2016) 3937–3943. doi: 10.1021/acs.chemmater.6b01244

    5. [5]

      J. Jang, H. s. Kim, S. Moon, et al., Nano Lett. 22 (2022) 761–767. doi: 10.1021/acs.nanolett.1c04290

    6. [6]

      M.H. Wu, L.B. Gu, Q. Wang, et al., ChemNanoMat 4 (2018) 387–393. doi: 10.1002/cnma.201700369

    7. [7]

      C.F. Zhang, S.J. Kim, M. Ghidiu, et al., Adv. Funct. Mater. 26 (2016) 4143–4151. doi: 10.1002/adfm.201600682

    8. [8]

      I.Z. Nejad, N. Zolfaghari, M.M. Mashhadi, et al., Sustain. Energy Technol. Assess. 47 (2021) 101438.

    9. [9]

      W.L. Cai, C. Yan, Y.X. Yao, et al., Small Struct. 1 (2020) 2000010. doi: 10.1002/sstr.202000010

    10. [10]

      H.T. Li, F.T. Lv, X. Fang, et al., Carbon Neutral. 2 (2023) 678–688. doi: 10.1002/cnl2.93

    11. [11]

      Y.P. Wang, P.C. Yuan, Z.Y. Xu, et al., Chin. Chem. Lett. 35 (2024) 108776. doi: 10.1016/j.cclet.2023.108776

    12. [12]

      H.A. Tariq, U. Nisar, J.J. Abraham, et al., Appl. Surf. Sci. 583 (2022) 152441. doi: 10.1016/j.apsusc.2022.152441

    13. [13]

      L. Li, G.X. Jiang, C.H. An, et al., Nanoscale 12 (2020) 10369–10379. doi: 10.1039/d0nr01222j

    14. [14]

      M.H. Wu, Y.P. Gao, Y. Hu, et al., Chin. Chem. Lett. 31 (2020) 897–902. doi: 10.1016/j.cclet.2019.07.039

    15. [15]

      H. Zhang, L. Yang, P.G. Zhang, et al., Adv. Mater. 33 (2021) 2008447. doi: 10.1002/adma.202008447

    16. [16]

      L. Petrizza, D. Genovese, G. Valenti, et al., Electroanalysis 28 (2016) 2777–2784. doi: 10.1002/elan.201600262

    17. [17]

      X.L. Liu, Y.X. Chen, H.B. Liu, et al., J. Mater. Sci. Technol. 33 (2017) 239–245. doi: 10.1016/j.jmst.2016.07.021

    18. [18]

      Y. Zhang, Z.H. Bao, R. Wang, et al., Nanoscale 16 (2024) 1751–1757. doi: 10.1039/d3nr05723b

    19. [19]

      J.P. Yang, Y.X. Wang, W. Li, et al., Adv. Mater. 29 (2017) 1700523. doi: 10.1002/adma.201700523

    20. [20]

      N. Talreja, M. Ashfaq, D. Chauhan, et al., Mater. Chem. Phys. 294 (2023) 127029. doi: 10.1016/j.matchemphys.2022.127029

    21. [21]

      Q. Liu, Y.X. Ji, X.M. Yin, et al., Energy Storage Mater. 46 (2022) 384–393. doi: 10.1016/j.ensm.2021.12.017

    22. [22]

      Y.C. Zhang, Y. Tian, Z.R. Wang, et al., Chem. Eng. J. 458 (2023) 141388. doi: 10.1016/j.cej.2023.141388

    23. [23]

      N. Huang, C. Tang, H. Jiang, et al., Nano Res. 17 (2023) 2619–2627.

    24. [24]

      C. Yang, Y. Liu, X. Sun, et al., Electrochim. Acta 271 (2018) 165–172. doi: 10.1016/j.electacta.2018.03.118

    25. [25]

      X.F. Lou, Y.Y. Zhang, L. Zhao, et al., Sci. Rep. 13 (2023) 18560. doi: 10.1038/s41598-023-45763-3

    26. [26]

      Y. Li, S. Wang, D.N. Lei, et al., J. Mater. Chem. A 5 (2017) 12236–12242. doi: 10.1039/C7TA02361H

    27. [27]

      H.L. Wang, P.C. Zhang, P.F. Huang, J. Chem. Technol. Biotechnol. 97 (2022) 2815–2819. doi: 10.1002/jctb.7150

    28. [28]

      W. Devina, D. Nam, J. Hwang, et al., Electrochim. Acta 321 (2019) 134639. doi: 10.1016/j.electacta.2019.134639

    29. [29]

      Y.J. Zheng, B.J. Liu, P. Cao, et al., J. Mater. Sci. Technol. 35 (2019) 667–673. doi: 10.1016/j.jmst.2018.10.028

    30. [30]

      W.L. Xu, Y.L. Xu, T. Schultz, et al., ACS Appl. Mater. Interfaces 15 (2022) 795–805.

    31. [31]

      G. Hasegawa, T. Sato, K. Kanamori, et al., New J. Chem. 38 (2014) 1380–1384. doi: 10.1039/c3nj01332d

    32. [32]

      D. McNulty, E. Carroll, C. O'Dwyer, Adv. Energy Mater. 7 (2017) 1602291. doi: 10.1002/aenm.201602291

    33. [33]

      M.N. Hyder, B.M. Gallant, N.J. Shah, et al., Nano Lett. 13 (2013) 4610–4619. doi: 10.1021/nl401387s

    34. [34]

      M.N. Tahir, B. Oschmann, D. Buchholz, et al., Adv. Energy Mater. 6 (2016) 1501489. doi: 10.1002/aenm.201501489

    35. [35]

      F.Y. He, C. Tang, G.J. Zhu, et al., Sci. China Chem. 64 (2021) 964–973. doi: 10.1007/s11426-020-9942-9

    36. [36]

      D.P. Li, X.H. Ren, Q. Ai, et al., Adv. Energy Mater. 8 (2018) 1802386. doi: 10.1002/aenm.201802386

    37. [37]

      W.X. Zhao, X.Q. Ma, X.D. Wang, et al., J. Mater. Chem. A 11 (2023) 2431–2442. doi: 10.1039/d2ta08276d

    38. [38]

      L.B. Tang, P.Y. Li, R.D. Cui, et al., Small Methods 7 (2023) 2201387. doi: 10.1002/smtd.202201387

    39. [39]

      S.F. Ho, H.Y. Tuan, Chem. Eng. J. 452 (2023) 139199. doi: 10.1016/j.cej.2022.139199

    40. [40]

      D.G. Sun, C. Tang, H. Cheng, et al., J. Energy Chem. 72 (2022) 479–486. doi: 10.1016/j.jechem.2022.05.043

    41. [41]

      M.M. Abdelaal, C.C. Yang, T.F. Hung, J. Energy Storage 77 (2024) 109926. doi: 10.1016/j.est.2023.109926

  • Figure 1  (a) The fabrication process of A-TiO2/Ti3C2Tx. (b, c) SEM images, (d) TEM image, (e, f) HRTEM images (insert is the SAED pattern), (g) particle size distribution profile, and (h) STEM image and corresponding elemental mappings of A-TiO2/Ti3C2Tx.

    Figure 2  (a) XRD patterns, (b) Raman spectra, and (c) FT-IR spectra of A-TiO2/Ti3C2Tx and TiO2/Ti3C2Tx. (d) N2 sorption isotherm and corresponding pore size distribution curve, and high-resolution XPS spectra of A-TiO2/Ti3C2Tx for (e) Ti 2p, (f) O 1s, (g) C 1s, (h) Si 2p and (i) N 1s.

    Figure 3  (a) CV curves of initial three cycles at a scan rate of 0.1 mV/s, and (b) charge-discharge profiles of the A-TiO2/Ti3C2Tx electrode at 100 mA/g. (c) Rate capabilities, (d) cycling performances at 100 mA/g, and (e) comparison of cycle performances between the A-TiO2/Ti3C2Tx electrode and other TiO2-based anodes for LIBs. (f) Long cycling performances of A-TiO2/Ti3C2Tx and TiO2/Ti3C2Tx electrodes at a high current density of 1000 mA/g.

    Figure 4  (a) CV curves at various scan rates, (b) the calculated b values at six redox peaks, (c) surface capacitive contribution at 1.0 mV/s, and (d) the percentage of capacitive contribution of the A-TiO2/Ti3C2Tx electrode at diverse scan rates. (e) GITT curves of A-TiO2/Ti3C2Tx and TiO2/Ti3C2Tx electrodes and their (f) Li+ diffusion coefficients. In-situ EIS diagram of (g) discharge and (h) charge of A-TiO2/Ti3C2Tx, and calculated Rct resistance values for (i) discharge and (j) charge.

    Figure 5  (a) Lithium adsorption energies (Eads) at different sites (Ⅰ and Ⅱ) for TiO2/Ti3C2N2, TiO2/Ti3C2, Ti3C2N2 and TiO2 surface, respectively. (b) Corresponding charge density difference (iso-value of 0.002 e/Å3) for each surface at their most stable site. (c) Schematic diagram of lithium storage merits of A-TiO2/Ti3C2Tx.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  153
  • HTML全文浏览量:  2
文章相关
  • 发布日期:  2025-06-15
  • 收稿日期:  2024-05-21
  • 接受日期:  2024-06-04
  • 网络出版日期:  2024-06-04
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

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

/

返回文章