Next Article in Journal
Economic, Low-Carbon Dispatch of Seasonal Park Integrated Energy System Based on Adjustable Cooling–Heating–Power Ratio
Previous Article in Journal
Development of Low-Emission Cooking Device Based on Catalytic Hydrogen Combustion Technology
Previous Article in Special Issue
Preparation and Electrochemical Performance of Zinc-Doped Copper Fluoride
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

H1.07Ti1.73O4-Derived Porous Plate-like TiO2 as High-Performance Bifunctional Anodes for Lithium- and Sodium-Ion Batteries

1
School of Marine Science and Technology, Northwestern Polytechnical University, Xi’an 710072, China
2
School of Materials Science and Engineering, Shaanxi University of Science and Technology, Xi’an 710021, China
*
Authors to whom correspondence should be addressed.
Energies 2025, 18(19), 5077; https://doi.org/10.3390/en18195077
Submission received: 5 September 2025 / Revised: 18 September 2025 / Accepted: 22 September 2025 / Published: 24 September 2025

Abstract

Porous plate-like anatase TiO2 particles were synthesized through a direct calcination approach using layered titanate H1.07Ti1.73O4 as a precursor. By controlling the calcination temperature (400 °C, 500 °C, and 600 °C), the morphology and electrochemical properties of the TiO2 samples were effectively tuned. When evaluated as anodes for lithium-ion batteries (LIBs), the porous TiO2 materials demonstrated markedly improved rate performance compared to commercial nano-TiO2 (n-TiO2). Specifically, at a high current density of 5.0 A/g, p-TiO2-500 and p-TiO2-600 delivered discharge capacities of 70.5 mAh/g and 87.5 mAh/g, respectively, far exceeding the 27.7 mAh/g achieved by n-TiO2. The corresponding capacity retention rates at this rate were 30.1% for p-TiO2-500, 41.2% for p-TiO2-600, and only 16.4% for n-TiO2. The enhancement in kinetics is ascribed to the unique porous plate-like architecture, which promotes efficient ion transport and introduces significant pseudocapacitive contributions. When applied as anodes for sodium-ion batteries (SIBs), p-TiO2-600 exhibited the most promising performance. This study underscores the potential of porous plate-like TiO2 as a high-performance bifunctional anode material suitable for both LIBs and SIBs.

1. Introduction

Carbon materials are promising anodes for LIBs due to their tunable properties, low cost, and abundance. However, graphite suffers from low theoretical capacity, while disordered carbons exhibit poor initial Coulombic efficiency (ICE), driving the search for higher-performance alternatives [1,2]. The development of anode materials that exhibit high-rate capability and enhanced safety is imperative for driving progress in the energy sector. Success in this endeavor is paramount for optimizing the multifaceted performance metrics of energy storage devices, namely energy density, cycle life, cost-effectiveness, and safety [3,4,5,6]. TiO2 has become a widely studied anode material for lithium/sodium-ion batteries due to its intrinsic safety brought by high working voltage (~1.75 V for Li, ~0.7 V for Na), simple synthesis method, and low cost [7,8]. As a semiconductor, TiO2 has a wide bandgap (3.2 eV), which leads to extremely low intrinsic electron/ion conductivity, severely restricting ion diffusion kinetics and electrochemical performance, resulting in poor rate performance and cycling stability [9]. In order to solve this problem, various strategies have been proposed to improve the lithium-ion diffusion pathway of TiO2 to increase its energy density [10,11]. Common methods include designing composite materials, morphology control, surface modification engineering, metal element doping, etc. [12]. Composite fabrication through the application of a conductive coating on TiO2 nanoparticles has been shown to facilitate rapid ion diffusion [13,14]. Notwithstanding this advantage, the high surface energy and subsequent agglomeration of these nanomaterials adversely impact their rate performance and long-cycle stability [15].
Constructing porous anatase TiO2 is also a significant strategy for enhancing the properties of TiO2 [16,17]. It is reported that the preparation of plate-like TiO2 is an effective method for enhancing the kinetics of electrochemical reactions [18,19]. Huang. H et al. [20] coated TiO2 with carbon nanotubes (CNTs), which improved its electrical conductivity. Jipeng Wang et al. [21] used carbon spheres as templates and synthesized TiO2 with a hollow-sphere morphology structure by the sol–gel method. The excellent morphology effectively improved its lithium storage performance. In addition, some metals, such as germanium (Ge) and nickel (Ni), have higher capacity and ion diffusion rates than carbon-based ones. Kim et al. prepared TiO2/Ge nanocomposites through the sol–gel process, which provided better lithium storage capacity [22]; Xu et al. [23] synthesized highly ordered 3D Ni-TiO2 core–shell nanoarrays (NTNAs), which demonstrated excellent electrochemical performance as the anode of sodium-ion batteries. Additionally, numerous studies have reported the potential applications of porous TiO2 in the field of lithium-ion and sodium-ion batteries [24,25,26,27]. However, the above-mentioned methods are cumbersome to operate and costly. There are still considerable challenges if these preparation methods are put into large-scale production.
In this work, porous plate-like anatase TiO2 was successfully fabricated, and its morphology was finely tuned by controlling the reaction temperature to enhance the electrical conductivity and cyclability of TiO2. The porous plate-like anatase TiO2 offers the following advantages. (1) The porous plate-like structure enhances dynamic performance. (2) The porous structure induces pseudocapacitive lithium/sodium storage behavior, thereby improving the electrochemical performance. This study systematically investigated the structure and electrochemical properties of porous plate-like anatase TiO2 electrodes. The influence of temperature on their morphology and electrochemical performance was thoroughly discussed. This also provides a new strategy for titanium-based oxide electrodes.

2. Experiment

2.1. Preparation of Porous Plate-like TiO2

A homogeneous stirring of 5.1 g KOH, 0.6 g LiOH-H2O, 6.9 g TiO2, and 25 mL of deionized water yields layered K0.81Li0.27Ti1.73O4·H2O (KLTO) under hydrothermal reaction at 250 °C for 24 h. Then, KLTO was protonated in nitric acid solution for 12 h to obtain the layered titanate H1.07Ti1.73O4·H2O (HTO) precursor; the preparation process is shown in Figure S1. Subsequently, the layered titanate HTO was heat-treated at 400 °C, 500 °C, and 600 °C for 3 h to obtain three porous plate-like TiO2 samples (Figure 1).

2.2. Characterization

The crystal structure and phase composition of the samples were characterized by X-ray diffraction (XRD, TD3500, Dandong Tongda Technology Co., Ltd., Dandong, China. The morphology, size, and elemental distribution of the electrodes before and after cycling were examined using a field emission scanning electron microscope (SEM, FEI Apreo S, The FEI Company of the United States, Hillsboro, OR, USA) equipped with energy-dispersive X-ray spectroscopy (EDS). Transmission electron microscopy (TEM, FEI Tecnai G2 F20 S-TWIN, The FEI Company of the United States, Hillsboro, OR, USA) was performed at an acceleration voltage of 200 kV to further analyze the microstructure of the powder samples. The specific surface area was determined through N2 adsorption–desorption measurements (ASAP 2460, Micromeritics, American Mc Instrument Company, Sebastian, FL, USA) and evaluated using the Brunauer–Emmett–Teller (BET) method. Phase transition behavior and thermal stability were investigated by thermogravimetric analysis (TGA, NETZSCH STA 409 PC, Germany NETZSCH, Selb, Germany).

2.3. Preparation of the Electrode

The p-TiO2 powder was blended with acetylene black and polyvinylidene fluoride (PVDF) binder in a weight ratio of 7:2:1 using N-methyl-2-pyrrolidone (NMP, Comio Chemical Reagents Co., Ltd., Tianjin, China) as the solvent to form a homogeneous slurry. The resulting slurry was coated onto copper foil and dried in a vacuum oven at 70 °C for 48 h. The dried electrode film was then punched into 10 mm diameter disks. CR2032-type coin half-cells were assembled in an argon-filled glove box (LABSTAR, Braun Inert Gas Company of Germany, Munich, Germany), using a microporous polypropylene membrane as the separator and lithium metal foil as the counter/reference electrode. The electrolyte consisted of 1 M LiPF6 dissolved in a mixture of ethylene carbonate (EC), ethyl methyl carbonate (EMC), and dimethyl carbonate (DMC) in a volume ratio of 1:1:1.

2.4. Electrical Measurement

Galvanostatic charge–discharge measurements were conducted within a voltage window of 0.01–3.0 V (vs. Li/Li+) at various current densities ranging from 0.1 to 5.0 A/g using a LAND battery test system (Wuhan Blue Power Electronics Co., Ltd., Wuhan, China). Cyclic voltammetry (CV) and pseudocapacitance analysis were carried out on a CHI660E electrochemical workstation (Shanghai Chenhua Instrument Co., Ltd., Shanghai, China) over the same voltage range, with scan rates varying from 0.1 to 1.0 mV·s−1. Electrochemical impedance spectroscopy (EIS) was also performed using the same CHI660E workstation.

3. Results

The XRD curves of the p-TiO2 prepared at different reaction temperatures are shown in Figure 2a. The samples were directly heat-treated from layered titanate HTO and were able to obtain pure phase anatase TiO2 (p-TiO2-400, p-TiO2-500, and p-TiO2-600) after heat treatments at 400 °C, 500 °C, and 600 °C. During the heat treatment of HTO powder, firstly, with the gradual increase in the muffle furnace temperature, the interlayer H3O+ and H+ began to be detached from the interlayer, accompanied by the weakening of the peak intensity of the characteristic peaks of the laminar structure near 2θ = 9.3, and shifted to a small angle, which signified that the laminar structure of lamellar titanate began to collapse, and the characteristic features of the laminar structure completely disappeared at 200 °C., and the formation of anatase TiO2 and rutile TiO2 began to appear. As the temperature continued to rise, the content of anatase TiO2 gradually increased, and when the temperature was higher than 400 °C, HTO was completely transformed into anatase TiO2. p-TiO2-400, p-TiO2-500, and p-TiO2-600 showed an increase in the crystallinity of samples with increasing temperature. The commercial nano TiO2 (n-TiO2) was also examined by XRD, and the characteristic peaks of the four groups of samples corresponded well to the (101), (0 0 4), (20 0), and (211) crystalline surfaces of anatase TiO2, which were consistent with the standard card Anatase TiO2 Jade 21-1272 [28], and all of them were anatase TiO2. The crystal structure of anatase TiO2 is shown in Figure 2b, where each titanium atom forms an octahedral coordination (TiO6) with six oxygen atoms, which has four formulaic units, where each TiO6 octahedron is connected to eight octahedra through four common vertices and four edges, thus constituting the tetragonal anatase crystal structure of I41/amd.
We have carried out the morphological analysis of three porous plate-like TiO2 and commercial nano TiO2, and the SEM images are shown in Figure 3. The layered titanate precursor is a smooth plate-like structure with an overall particle size of about 1–2 μm (Figure 3a); the three porous plate-like TiO2 obtained after heat treatment at different temperatures as a titanium source are shown in Figure 3b–d. It can be seen that p-TiO2-400 obtained at 400 °C maintains the plate-like structure, the overall particle size does not change significantly, and the porous structure does not appear on the plate surface. p-TiO2-500 obtained by heat treatment at 500 °C has an overall particle size of about 1.5 μm, and a large number of uniformly distributed pores on the plate surface. p-TiO2-600, obtained by heat treatment at 600 °C, has the largest number of pores and the largest pore diameter on the particle surface, and has the largest number of holes and the largest pore size; the plate-like structure appears to collapse and fracture under the influence of temperature, and the edges of the plate-like particles are slightly smooth. Figure 3e shows commercial nano-TiO2 (n-TiO2), and the powder is 50–80 nm nanorods with obvious agglomeration.
Figure 4 shows the TEM images of the four samples, from which it can be seen that the TiO2 prepared with HTO as precursor maintains the plate-like morphology after heat treatment at different temperatures. With the increase in temperature, the micropores increase, and the particle size increases. n-TiO2 is 50~80 nm nanoparticles with an obvious agglomeration phenomenon. The EDS analysis graph of p-TiO2-500 is shown in Figure S2 in the Supporting Information.
Subsequently, we tested the specific surface area of four groups of samples, which were 9.7 cm2/g, 40.6 cm2/g, 42.6 cm2/g, and 13.3 cm2/g, respectively, as shown in Figure S3. The results showed that the specific surface area of TiO2 with two porous structures, p-TiO2-500 and p-TiO2-600, was higher than that of n-TiO2. p-TiO2-400 samples had a larger particle size and did not form a porous morphology, so the specific surface area was the lowest. The specific surface area of the samples affects their electrochemical performance to a certain extent, and a larger specific surface area is favorable for adsorbing more ions during the electrochemical reaction, which makes the reaction occur more fully and improves the performance of the cells.
Then, we analyzed the formation process of nanostructures of two samples, p-TiO2-500 and p-TiO2-600. The precursor HTO layer is intercalated with H3O+, H+, and water molecules. The structure is a two-dimensional nanosheet layer composed of [TiO6] octahedra. As the temperature rises, the physically adsorbed water and some bound water between the layers first desorb. Subsequently, the H3O+ and H+ ions between the layers began to detach. From the XRD graph, it can be observed that the characteristic peak intensity of the HTO-layered structure weakens and shifts to a smaller angle, and eventually disappears completely. It indicates that the layer spacing changes due to the loss of water molecules. Between 200 and 400 degrees Celsius, the -OH and H+ in the precursor begin to condense and then escape, leaving oxygen vacancies and unsaturated bonds in the [TiO6] structure. Subsequently, Ti and O atoms undergo migration and rearrangement, forming Ti-O-Ti covalent bonds and undergoing topological transformation. At this point, the macroscopic plate-like morphology is retained, but the crystal structure of the surrounding area has changed. When the temperature exceeds 400 °C, HTO completely transforms into rutile-type TiO2. When the temperature is above 500 °C, the rapid escape of water vapor and the internal stress generated by structural reorganization are released by forming uniformly distributed pores. The structure undergoes local collapse, fracture, and reconstruction, forming a porous plate-like structure. When the temperature reaches 600 °C, atomic migration is too rapid and sintering takes the dominant role. Some plate-like structures collapse and fracture excessively, as shown in Figure 5.

3.1. Lithium-Ion Battery Performance Testing and Analysis

The multiplicity performance of the lithium-ion half-cells was first tested (Figure 6a). p-TiO2-500 and p-TiO2-600 both exhibited superior multiplicity performance to n-TiO2, while p-TiO2-400 had a slightly lower capacity than n-TiO2 at small current density, but gradually outperformed n-TiO2 as the current was increased. Overall, the multiplicative performance of all three groups of porous plate TiO2 is better than that of n-TiO2, and p-TiO2-500 obtains the most excellent multiplicative performance due to its optimal morphology. The details of the charging/discharging capacities of all the samples at 0.1 A/g, 0.2 A/g, 0.5 A/g, 1.0 A/g, 2.0 A/g, and 5.0 A/g recovered to 0.1 A/g current density and can be seen as shown in Table 1
The cycling performance at a current density of 0.1 A/g was subsequently tested, as shown in Figure 6b. The three porous plate TiO2 exhibited higher electrochemical performance than n-TiO2. Among them, p-TiO2-500 has the best performance, maintaining a reversible capacity of 350 mAh/g after 100 charge/discharge cyclings. p-TiO2-600, p-TiO2-400, and n-TiO2 can maintain a reversible capacity of 263 mAh/g, 207.5 mAh/g, and 169 mAh/g after 100 cyclings, respectively.
The p-TiO2-500, p-TiO2-600, and n-TiO2 were tested for 5000 charge/discharge cyclings at a high current density of 1.0 A/g, as shown in Figure 6c. The results show that the p-TiO2-500 and p-TiO2-600 samples can still maintain a reversible capacity of 112.3 mAh/g and 111.8 mAh/g, which are both higher than that of n-TiO2 (61.6 mAh/g), with a performance enhancement of 82.30% and 81.49%, respectively.
Figure 7 shows the capacity/voltage curves and CV curves for the n-TiO2, p-TiO2-500, and p-TiO2-600 samples. The capacity/voltage curves of the three groups of samples at 0.1 A/g current density for the first three laps, the fifth lap, and the tenth lap are shown in Figure 7a,c,e. The first discharges were 493.1 mAh/g, 612.6 mAh/g, and 518.8 mAh/g, and the Coulomb efficiencies were 41.53%, 38.96%, and 39.86%, respectively. The capacity decay of n-TiO2 is significantly more pronounced, as it can only sustain a capacity of 164.1 mAh/g by the 10th cycle. In contrast, p-TiO2-500 and p-TiO2-600 are able to maintain capacities of 205.3 mAh/g and 202.6 mAh/g at the same point, respectively, exhibiting no significant signs of capacity decay; their curves nearly overlap. All three samples exhibit an embedded lithium platform near 1.7 V, which belongs to the redox reaction of Ti4+/Ti3+ brought about by the intercalation reaction of anatase TiO2 with lithium ions at this working potential [29]. The electrochemical reaction equation is shown in Equation (1) [30]:
TiO2 + xLi + +xe→LixTiO2
Figure 7b,d, and f illustrate the cyclic voltammetry (CV) curves of n-TiO2, p-TiO2-500, and p-TiO2-600 at a scan rate of 0.1 mV/s, respectively. The characteristic peaks observed at 1.7 V and 2.0 V correspond to the reversible conversion between Ti4+ and Ti3+, which aligns with the lithium intercalation/de-intercalation plateau evident in the capacity/voltage profiles. The enhanced intensity of these characteristic peaks for both porous TiO2 variants indicates a more efficient intercalation reaction during charge and discharge cyclings; consequently, p-TiO2-500 and p-TiO2-600 exhibit superior reversible capacities.
It has been shown that when the electrode material is porous, pseudo-capacitive Li+ storage behavior occurs during charging and discharging. Voltametric cycling tests were conducted on four distinct TiO2 species at varying scan rates, as illustrated in Figure 8. The correlation between current and scan rate was analyzed to confirm the presence of pseudocapacitive behavior during both charging and discharging processes. For a redox reaction, the peak current obeys the following power law [31]:
i = avb
In the formula, i, v, and a represent current (mA), scanning rate (mV/s), and an arbitrary constant, respectively, while b can determine whether there is pseudocapacitance behavior in the lithium storage process. As shown in Figure 8d, the logarithm of both sides of the equation gives a linear relationship between log (i) and log (v), and this linear equation has a slope of b value. According to the pseudocapacitance judgment theory, if b = 0.5, it means that only the intercalation reaction exists in the battery. If b = 1, it means that only pseudocapacitance lithium storage behavior exists. The pseudocapacitance was calculated as a function of current and scanning rate, and the b-values of the three TiO2 samples were 0.79, 0.63, and 0.67, respectively, indicating that all three samples have a lithium storage mechanism that is controlled by both the intercalation reaction and the pseudocapacitance during the charge/discharge process.
In addition, we calculated the pseudocapacitance ratios of the three TiO2 according to Equation (3) [32]:
i(V) = k1v + k2v1/2
k1v and k2v1/2 correspond to the pseudocapacitance and the current contribution of the nested process, respectively. The lithium storage behavior of pseudocapacitance brings great capacity contribution to the composite anode electrode materials at different scanning rates, and CV curves at different scanning rates are shown in Figure 9.
We have performed analytical tests related to the kinetics of electrochemical reactions; electrochemical impedance spectroscopy (EIS) was performed at a frequency in the range of 0.01 Hz~1000 kHz for three groups of samples and fitted the data to an equivalent circuit diagram as shown in the inset (Figure 10a, all the samples have been subjected to 100 charging and discharging cyclings at a current density of 0.1 A/g). It can be seen that p-TiO2-500 has the lowest charge transfer resistance, p-TiO2-600 has a slightly higher charge transfer resistance, and n-TiO2 has the largest charge transfer resistance. Rs represents the internal resistance of the cell, and the diagonal line indicates the Warburg resistance (W), which is an important parameter for calculating the lithium-ion diffusion rate (DLi+), which can be calculated by Equation (4) [33]. The details of the impedance and the lithium-ion diffusion coefficient for each part of the three sets of samples are shown in Table 2. As calculated, the two porous TiO2 not only exhibit lower charge transfer impedance than n-TiO2, but also possess a higher lithium-ion diffusion rate.
D L i + = R 2 T 2 2 A 2 n 4 F 4 C 2 σ 2
In the formula, R denotes a constant gas (8.314 J K−1 mol−1) and T denotes the ambient temperature (298.15 K), A denotes the surface of the electrode (0.785 cm2), and n denotes the number of electrons in the electronic transfer reactions. The Faraday constant (96,486 C/mol) is F, σ [34] is the slope of Z′, and with respect to ω1/2 (Figure 7b), and C is the concentration of lithium-ion phase (mol/cm3). The capacity/voltage curves and dQ/dV curves of the three sets of samples are shown in Figure 10c,d. The two porous TiO2 exhibit a more pronounced intercalation reaction plateau at 1.7 V in the discharge curves than n-TiO2, as well as stronger intercalation reaction-embedded lithium peaks in the dQ/dV curves. p-TiO2-500 and p-TiO2-600 exhibit significantly improved electrochemical performance due to the higher degree of intercalation reaction.

3.2. Sodium-Ion Battery Performance Testing and Analysis

The electrochemical performance of p-TiO2-500, p-TiO2-600, and n-TiO2 sodium-ion batteries is shown in Figure 11. The porous structured TiO2 all demonstrated a performance enhancement to varying degrees compared to n-TiO2, with the p-TiO2-600 sample showing the most significant performance enhancement.
We tested the sodium-ion CV curves and capacity/voltage curves for three groups of samples. The CV curves of the three sets of samples for the first three revolutions at a scan rate of 0.1 mV/s show (Figure 12a,c,e) that all three sets of samples exhibit a de-embedding peak of sodium ions near 0.7 V, where an intercalation reaction occurs at this potential, corresponding to the reversible transition between Ti3+ and Ti4+ that occurs in Equation (5) [35]. The capacity/voltage profiles of the three groups of samples for the first three laps at a current density of 0.1 A/g, as well as the tenth lap, are shown in Figure 12b,d,f. The first-lap discharges of all the samples exhibit a large irreversible capacity loss:
TiO2 + xNa+ + xe→NaxTiO2
In order to determine the embedded sodium platform more accurately, we plotted the capacity/voltage curves of the three groups of samples after 100-turn charge–discharge cyclings at 0.1 A/g current density and differentiated the corresponding discharge curves (embedded sodium stage) in terms of dQ/dV, as shown in Figure 13. The results show that both porous TiO2 exhibit higher charge/discharge capacities than n-TiO2, with a downward-sloping line near 1.0–0 V in the embedded sodium stage for all of them. By differentiating the three discharge curves in terms of dQ/dV (Figure 13b), an embedded peak near 0.7 V is seen for all three groups of samples, which is consistent with the CV curve results. Compared with the corresponding dQ/dV curves tested for Li-ion batteries, the characteristic peak intensities of the dQ/dV curves for porous TiO2 sodium-ion batteries are slightly higher than those for n-TiO2, with the highest characteristic peak intensity for p-TiO2-600 (the dQ/dV characteristic peak intensities for p-TiO2-500 and p-TiO2-600 Li-ion batteries, on the other hand, are significantly higher than those for n-TiO2, and the characteristic peak intensities for p-TiO2-500 have the highest characteristic peak intensity, Figure 10d). The difference between the two cells may be due to the lower intercalation reaction of the porous structure of the sodium-ion radius ambassador and the larger pore size of the p-TiO2-600 surface micropores, which resulted in the highest electrochemical performance in the sodium-ion batteries.
Subsequently, we investigated the electrochemical kinetic properties of sodium ions across the three sample groups, as illustrated in Figure 14. The equivalent circuit fitting of the data indicates that the porous plate-like p-TiO2-500 exhibits the lowest charge-transfer resistance, followed by p-TiO2-600 with a slightly higher resistance, while n-TiO2 demonstrates the highest resistance. These findings are consistent with those obtained from lithium-ion battery EIS tests. Additionally, we calculated the sodium ion diffusion rate (DNa+) for all three sample groups in accordance with Equation (4).
The ion-diffusion relation coefficient Weber factor σ was obtained by fitting the linear relationship between Z′ and ω1/2, and then the sodium-ion diffusion coefficient calculation was carried out. The details of impedance, as well as sodium-ion diffusion coefficients for each part of the three sets of samples, are shown in Table 3.

4. Conclusions

In this study, porous-platelet anatase TiO2 was prepared using layered titanate as the titanium source and platelet-template heat treatment. Porous platelet TiO2 exhibits higher electrochemical performance than n-TiO2 in both lithium/sodium-ion batteries, with the most significant performance enhancement from the p-TiO2-500 lithium-ion battery and the best electrochemical performance from the p-TiO2-600 sodium-ion battery. The porous structure improves the kinetic properties and is accompanied by a large number of pseudocapacitive behaviors that work together and thus can bring a significant performance enhancement to them. The tests related to lithium/sodium-ion batteries confirm that this porous plate-like TiO2 is a promising and versatile anode material for batteries.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/en18195077/s1, Figure S1: Flow chart of preparation of layered titanate H1.07Ti1.73O4·H2O (HTO) precursor; Figure S2: Corresponding elemental mapping images of p-TiO2-500; Figure S3: BET analysis of p-TiO2-400 (a), p-TiO2-500 (b), p-TiO2-600 (c) and n-TiO2 (d).

Author Contributions

Y.S.: Writing—original draft, visualization, methodology, formal analysis, data curation, and conceptualization. Y.L.: Methodology and investigation. J.L.: Methodology and investigation. S.L. (Siyuan Liu): Methodology and investigation. S.L. (Silun Luo): Methodology and investigation. S.Z.: Methodology and investigation. X.K.: Methodology and validation. Q.H.: Investigation. C.L.: Project administration, methodology, and investigation. All authors have read and agreed to the published version of the manuscript.

Funding

Funded by the National Key Research and Development Program (2020YFB1313200, 2022YFC2805200).

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Li, S.; Wang, K.; Zhang, G.; Li, S.; Xu, Y.; Zhang, X.; Zhang, X.; Sun, X.; Ma, Y.; Zheng, S. Fast Charging Anode Materials for Lithium-Ion Batteries: Current Status and Perspectives. Adv. Funct. Mater. 2022, 32, 2200796. [Google Scholar] [CrossRef]
  2. Wang, R.; Wang, L.; Liu, R.; Li, X.; Wu, Y.; Ran, F. “Fast-Charging” Anode Materials for Lithium-Ion Batteries from Perspective of Ion Diffusion in Crystal Structure. ACS Nano 2024, 18, 2611–2648. [Google Scholar] [CrossRef]
  3. Yang, Z.-Z.; Zhang, C.-Y.; Ou, Y.-Q.; Su, Z.-K.; Zhao, Y.; Cong, H.-J.; Ai, X.-P.; Qian, J.-F. Amorphous Sb/C composite with isotropic expansion property as an ultra-stable and high-rate anode for lithium-ion batteries. Rare Met. 2024, 43, 2039–2052. [Google Scholar] [CrossRef]
  4. Siti Norhasanah, S.; Hwa-Young, Y.; Jeevan, K.; Ye-Chong, M.; Gangasagar Sharma, G.; Seung-Ju, Y.; Yong Ju, K.; Sejung, K.; Bong-Hyun, J.; Won-Yeop, R. Machine Learning-Assisted Fabrication of PCBM-Perovskite Solar Cells with Nanopatterned TiO2 Layer. Energy Environ. Mater. 2024, 7, 219–226. [Google Scholar] [CrossRef]
  5. Yu, L.; Yu, X.Y.; Lou, X.W. The Design and Synthesis of Hollow Micro-/Nanostructures: Present and Future Trends. Adv. Mater. 2018, 30, 1800939. [Google Scholar] [CrossRef] [PubMed]
  6. Wang, G.; Lei, H.; Liu, Z.; Yuan, Z.; Li, L.; Zhan, Z.; Wang, X. An oxygen-deficient Ge/GeO2/C anode for lithium ion batteries with enhanced reversible energy storage performance. J. Power Sources 2025, 632, 236319. [Google Scholar] [CrossRef]
  7. Hyeon, C.-W.; Kim, B.; Kim, C.W.; Li, L.; Chung, C.-Y.; Chun, S.-E. Thin-shelled hollow mesoporous TiO2 spheres with less tortuosity as fast-charging anode. Compos. Part B 2024, 286, 111760. [Google Scholar] [CrossRef]
  8. Wang, G.; Gao, W.; Zhan, Z.; Li, Z. Defect-engineered TiO2 nanocrystals for enhanced lithium-ion battery storage performance. Appl. Surf. Sci. 2022, 598, 153869. [Google Scholar] [CrossRef]
  9. Hsu, H.-J.; Li, C.-C. TiO2-based microsphere with large pores to improve the electrochemical performance of Li-ion anodes. Ceram. Int. 2021, 47, 12038–12046. [Google Scholar] [CrossRef]
  10. Liu, C.; Zhao, P.; Lu, K.; Zhang, C.; Xia, X.; Lei, W.; Guo, Q.; Hao, Q.; Han, S. Facilitating Highly Reversible Li-Ion Storage of MoSe2-TiO2-MXene via Double Heterostructures. Adv. Funct. Mater. 2024, 34, 2401392. [Google Scholar] [CrossRef]
  11. Ke, J.; Li, M.; Hu, A.; Gao, P.; Liu, J.; Chen, S.; Xiao, P.; Xu, C. Tailored Design Ti4+ Coordination via Coplanar Carboxyl and Hydroxyl Groups Toward High Purity TiO2(B) with Ultrafast Li+ Storage. Adv. Funct. Mater. 2024, 34, 2315218. [Google Scholar] [CrossRef]
  12. Su, Z.; Liu, J.; Li, M.; Zhu, Y.; Qian, S.; Weng, M.; Zheng, J.; Zhong, Y.; Pan, F.; Zhang, S. Defect Engineering in Titanium-Based Oxides for Electrochemical Energy Storage Devices. Electrochem. Energy Rev. 2020, 3, 286–343. [Google Scholar] [CrossRef]
  13. Yao, T.; Wang, H.; Ji, X.; Zhang, Q.; Shi, J.W.; Han, X.; Cheng, Y.; Wang, D.; Meng, L. Introducing Hybrid Defects of Silicon Doping and Oxygen Vacancies into MOF-Derived TiO2–X@Carbon Nanotablets Toward High-Performance Sodium-Ion Storage. Small 2023, 19, 2302831. [Google Scholar] [CrossRef]
  14. Zhang, D.; Xu, H. Nickel modified TiO2/C nanodisks with defective and near-amorphous structure for high-performance sodium-ion batteries. Battery Energy 2024, 3, 20230032. [Google Scholar] [CrossRef]
  15. Chen, J.; Fu, Y.; Sun, F.; Hu, Z.; Wang, X.; Zhang, T.; Zhang, F.; Wu, X.; Chen, H.; Cheng, G.; et al. Oxygen vacancies and phase tuning of self-supported black TiO2-X nanotube arrays for enhanced sodium storage. Chem. Eng. J. 2020, 400, 125784. [Google Scholar] [CrossRef]
  16. Xu, X.; Zhou, T.; Sun, Q.; Lu, Y. Advanced keratin-coated polypropylene composite current collectors: Design for in-situ, integrated flexible temperature sensing in lithium-ion batteries. Chem. Eng. J. 2025, 519, 165276. [Google Scholar] [CrossRef]
  17. Kong, X.; Su, Y.; Xing, C.; Cheng, W.; Huang, J.; Zhang, L.; Ouyang, H.; Feng, Q. Facile synthesis of porous TiO2/SnO2 nanocomposite as lithium ion battery anode with enhanced cycling stability via nanoconfinement effect. Chin. Chem. Lett. 2024, 35, 109428. [Google Scholar] [CrossRef]
  18. Lu, S.; Shang, Y.; Zheng, W.; Huang, Y.; Wang, R.; Zeng, W.; Zhan, H.; Yang, Y.; Mei, J. TiO2(B) nanosheets modified Li4Ti5O12 microsphere anode for high-rate lithium-ion batteries. Nanotechnology 2022, 33, 245404. [Google Scholar] [CrossRef]
  19. Ge, H.; Cui, L.; Zhang, B.; Ma, T.-Y.; Song, X.-M. Ag quantum dots promoted Li4Ti5O12/TiO2 nanosheets with ultrahigh reversible capacity and super rate performance for power lithium-ion batteries. J. Mater. Chem. A 2016, 4, 16886–16895. [Google Scholar] [CrossRef]
  20. Huang, H.; Zhang, W.K.; Gan, X.P.; Wang, C.; Zhang, L. Electrochemical investigation of TiO2/carbon nanotubes nanocomposite as anode materials for lithium-ion batteries. Mater. Lett. 2007, 61, 296–299. [Google Scholar] [CrossRef]
  21. Wang, J.; Bai, Y.; Zhang, W.F.; Wu, M.; Yin, J. Preparation and electrochemical properties of TiO2 hollow spheres as an anode material for lithium-ion batteries. J. Power Sources 2009, 191, 614–618. [Google Scholar] [CrossRef]
  22. Kim, H.; Kim, M.C.; Choi, S.; Moon, S.H.; Kim, Y.S.; Park, K.W. Facile one-pot synthesis of Ge/TiO2 nanocomposite structures with improved electrochemical performance. Nanoscale 2019, 11, 17415–17424. [Google Scholar] [CrossRef]
  23. Xu, Y.; Zhou, M.; Wen, L.; Wang, C.; Zhao, H.; Mi, Y.; Liang, L.; Lei, Y.; Fu, Q.; Wu, M. Highly Ordered Three-Dimensional Ni-TiO2 Nanoarrays as Sodium Ion Battery Anodes. Chem. Mater. 2015, 27, 4274–4280. [Google Scholar] [CrossRef]
  24. Li, Y.; Sun, S.; Tang, J.; Han, S.; Yang, Y.; Xing, J.; Zong, L.; Wang, L.; Li, B. Volume expansion restriction by TiO2 structural unit in silicon anodes with yolk-shell structure for lithium-ion batteries. Nano Res. 2025, 18, 94907474. [Google Scholar] [CrossRef]
  25. Wang, J.; Cheng, C.; Zhang, J.; Yang, J.; Wang, Z.; Li, L.; Bai, W.; Richard, Y.K.K. Mechanically robust and fireproof separator for safer lithium-ion batteries with enhanced thermal stability. J. Power Sources 2025, 641, 236905. [Google Scholar] [CrossRef]
  26. Helaley, A.; Liang, X. Solid-state sodium-ion batteries with composite polymer electrolytes and ALD-modified Na0.7MnO2 cathodes. Chem. Eng. J. 2025, 514, 163173. [Google Scholar] [CrossRef]
  27. Li, C.; Pan, Q.; Jiang, W.-J.; Liu, Z.-Y.; Chen, L. Titanium dioxide-based materials for alkali metal-ion batteries: Safety and development. J. Power Sources 2025, 634, 236492. [Google Scholar] [CrossRef]
  28. Zhang, W.; Yao, F.; Al Samarai, M.; Feng, Q. Mesoporous anatase TiO2 mesocrystal for high-performance photocatalysis and lithium-ion batteries. Nanoscale 2025, 17, 9418–9426. [Google Scholar] [CrossRef]
  29. Liu, X.; Zhou, J.; Zhu, G.; Li, J.; Zhang, H. Phosphorus-doped amorphous TiO2/C interface enables hierarchical SEI formation on micron-sized SiO anodes for ultra-stable lithium-ion batteries. J. Mater. Chem. A 2025, 13, 19429–19439. [Google Scholar] [CrossRef]
  30. Cheng, W.; Feng, Q.; Guo, Z.; Chen, G.; Wang, Y.; Yin, L.; Li, J.; Kong, X. Electrochemical reaction mechanism of porous Zn2Ti3O8 as a high-performance pseudocapacitive anode for Li-ion batteries. Chin. Chem. Lett. 2022, 33, 4776–4780. [Google Scholar] [CrossRef]
  31. Luebke, M.; Shin, J.; Marchand, P.; Brett, D.; Shearing, P.; Liu, Z.; Darr, J.A. Highly pseudocapacitive Nb-doped TiO2 high power anodes for lithium-ion batteries. J. Mater. Chem. A 2015, 3, 22908–22914. [Google Scholar] [CrossRef]
  32. Cui, L.; Hou, Y.; Sun, S.; Wang, M.; Wang, Y.; Li, G.; Yu, K. Nanoflower-shaped multivariate heterogeneous composites with multiple redox sites for high-performance lithium-ion batteries. J. Energy Storage 2025, 132, 117903. [Google Scholar] [CrossRef]
  33. Song, Z.; Zhou, Q.; Zeng, J.; Zhang, W.; Zhuang, S.; Luo, H.; Lu, M.; Li, X. Lithium-philic organic polymer@mixed-phase TiO2 core-shell nanospheres for high-rate and long-cyclic performance in liquid/solid-state lithium-ion batteries. J. Power Sources 2025, 626, 235782. [Google Scholar] [CrossRef]
  34. Du, X.Y.; He, W.; Zhang, X.D.; Yue, Y.Z.; Liu, H.; Zhang, X.G.; Min, D.D.; Ge, X.X.; Du, Y. Enhancing the electrochemical performance of lithium ion batteries using mesoporous Li3V2(PO4)(3)/C microspheres. J. Mater. Chem. 2012, 22, 5960–5969. [Google Scholar] [CrossRef]
  35. Yan, Z.; Fan, S.; Zou, X.; Jiang, Q.; Tang, D.; Lan, K.; Li, J.; Lin, Y.; Huang, Z.; Peng, D.-L.; et al. Scalable synthesis of anatase TiO2 nano-in-micro spheres for hybrid sodium-ion capacitors. Chem. Eng. J. 2025, 519, 165100. [Google Scholar] [CrossRef]
Figure 1. Preparation flow chart and formation process of microstructures of porous plate-like TiO2 (p-TiO2).
Figure 1. Preparation flow chart and formation process of microstructures of porous plate-like TiO2 (p-TiO2).
Energies 18 05077 g001
Figure 2. (a) XRD patterns of porous plate TiO2 and Commercial nano TiO2. (b) Crystal structure diagram of anatase TiO2.
Figure 2. (a) XRD patterns of porous plate TiO2 and Commercial nano TiO2. (b) Crystal structure diagram of anatase TiO2.
Energies 18 05077 g002
Figure 3. (a) SEM photos of layered HTO titanate precursor; (bd) three porous plate-like TiO2 samples obtained at different temperatures; (e) commercial nano-TiO2 reference sample.
Figure 3. (a) SEM photos of layered HTO titanate precursor; (bd) three porous plate-like TiO2 samples obtained at different temperatures; (e) commercial nano-TiO2 reference sample.
Energies 18 05077 g003
Figure 4. (a) The TEM images of p-TiO2-400, (b) p-TiO2-500, (c) p-TiO2-600, (d) and n-TiO2.
Figure 4. (a) The TEM images of p-TiO2-400, (b) p-TiO2-500, (c) p-TiO2-600, (d) and n-TiO2.
Energies 18 05077 g004
Figure 5. Schematic diagram of microstructure formations for porous plate-like TiO2.
Figure 5. Schematic diagram of microstructure formations for porous plate-like TiO2.
Energies 18 05077 g005
Figure 6. (a) Rate performance test diagram and (b) cyclic performance diagram at current density of 0.1 A/g of p-TiO2-400, p-TiO2-500, p-TiO2-600, and n-TiO2 reference samples. (c) Long-cycle performance diagram of p-TiO2-500, p-TiO2-600, and n-TiO2 reference samples at a large current density of 1.0 A/g.
Figure 6. (a) Rate performance test diagram and (b) cyclic performance diagram at current density of 0.1 A/g of p-TiO2-400, p-TiO2-500, p-TiO2-600, and n-TiO2 reference samples. (c) Long-cycle performance diagram of p-TiO2-500, p-TiO2-600, and n-TiO2 reference samples at a large current density of 1.0 A/g.
Energies 18 05077 g006
Figure 7. The capacity/voltage curves of commercial nano TiO2 reference samples, p-TiO2-500, and p-TiO2-600 samples in the first three, fifth, and tenth turns at a current density of 0.1 A/g (a,c,e). The first three voltammetry cycle curves of commercial nano TiO2 reference samples, p-TiO2-500, and p-TiO2-600 samples at a sweep rate of 0.1 mV/s (b,d,f).
Figure 7. The capacity/voltage curves of commercial nano TiO2 reference samples, p-TiO2-500, and p-TiO2-600 samples in the first three, fifth, and tenth turns at a current density of 0.1 A/g (a,c,e). The first three voltammetry cycle curves of commercial nano TiO2 reference samples, p-TiO2-500, and p-TiO2-600 samples at a sweep rate of 0.1 mV/s (b,d,f).
Energies 18 05077 g007
Figure 8. (a) Voltammetry cycle curves of n-TiO2 reference sample; (b) p-TiO2-500 and (c) p-TiO2-600 at different sweep speeds; and (d) their calculation of b value.
Figure 8. (a) Voltammetry cycle curves of n-TiO2 reference sample; (b) p-TiO2-500 and (c) p-TiO2-600 at different sweep speeds; and (d) their calculation of b value.
Energies 18 05077 g008
Figure 9. The pseudocapacitance contribution ratio of n-TiO2, p-TiO2-500, and p-TiO2-600 at scanning rates of 0.1 mV/s, 0.2 mV/s, 0.5 mV/s, 0.8 mV/s, and 1.0 mV/s (n-TiO2: (ae), p-TiO2-500: (fj), p-TiO2-600: (ko)).
Figure 9. The pseudocapacitance contribution ratio of n-TiO2, p-TiO2-500, and p-TiO2-600 at scanning rates of 0.1 mV/s, 0.2 mV/s, 0.5 mV/s, 0.8 mV/s, and 1.0 mV/s (n-TiO2: (ae), p-TiO2-500: (fj), p-TiO2-600: (ko)).
Energies 18 05077 g009
Figure 10. (a) Electrochemical impedance spectra of n-TiO2, p-TiO2-500, and p-TiO2-600; (b) linear relationship diagram of Z′ and ω1/2; (c) capacity/voltage curve diagram; and (d) dQ/dV curve analysis diagram.
Figure 10. (a) Electrochemical impedance spectra of n-TiO2, p-TiO2-500, and p-TiO2-600; (b) linear relationship diagram of Z′ and ω1/2; (c) capacity/voltage curve diagram; and (d) dQ/dV curve analysis diagram.
Energies 18 05077 g010
Figure 11. Sodium-ion battery rate performance test (a) and cycle performance test (b) of p-TiO2-500, p-TiO2-600, and n-TiO2 samples.
Figure 11. Sodium-ion battery rate performance test (a) and cycle performance test (b) of p-TiO2-500, p-TiO2-600, and n-TiO2 samples.
Energies 18 05077 g011
Figure 12. The voltammogram curves of the n-TiO2, p-TiO2-500 and p-TiO2-600 samples during the first three cycles at a scanning rate of 0.1 mV/s (a,c,e), as well as the capacity/voltage curves at an current density of 0.1 A/g during the first three cycles and the 10th cycle (b,d,f).
Figure 12. The voltammogram curves of the n-TiO2, p-TiO2-500 and p-TiO2-600 samples during the first three cycles at a scanning rate of 0.1 mV/s (a,c,e), as well as the capacity/voltage curves at an current density of 0.1 A/g during the first three cycles and the 10th cycle (b,d,f).
Energies 18 05077 g012
Figure 13. Sodium-ion battery capacity/voltage curve (a) and corresponding dQ/dV curve (b) of n-TiO2, p-TiO2-500, and p-TiO2-600 samples after stabilization.
Figure 13. Sodium-ion battery capacity/voltage curve (a) and corresponding dQ/dV curve (b) of n-TiO2, p-TiO2-500, and p-TiO2-600 samples after stabilization.
Energies 18 05077 g013
Figure 14. (a) Sodium-ion battery electrochemical impedance spectra of n-TiO2, p-TiO2-500, and p-TiO2-600. (b) linear relationship diagram of Z′ and ω1/2.
Figure 14. (a) Sodium-ion battery electrochemical impedance spectra of n-TiO2, p-TiO2-500, and p-TiO2-600. (b) linear relationship diagram of Z′ and ω1/2.
Energies 18 05077 g014
Table 1. Comparison table of the rate performance of four TiO2 samples.
Table 1. Comparison table of the rate performance of four TiO2 samples.
MaterialDischarge/ChargeThe Capacity (mAh/g) at Current Density (A/g) of:
0.10.20.51.02.05.00.1
p-TiO2-400Discharge152.3122.192.376.966.754.2159.4
Charge154.2122.192.376.966.754.2159.4
p-TiO2-500Discharge228.8196.3153.9127.5101.670.5226.9
Charge234.5201.2155.7127.6101.670.5228.6
p-TiO2-600Discharge208.6178.1149.0126.4109.587.5217.8
Charge212.3178.3149.3126.4109.587.5218.6
n-TiO2Discharge166.6142.390.265.647.627.7162.0
Charge169.2143.991.665.847.627.7162.2
Table 2. Comparison table of impedance of each part for n-TiO2, p-TiO2-500, and p-TiO2-600, and their lithium-ion diffusion coefficients.
Table 2. Comparison table of impedance of each part for n-TiO2, p-TiO2-500, and p-TiO2-600, and their lithium-ion diffusion coefficients.
SampleRs (Ω)Rct (Ω)Z′ (Ω)DLi+ (cm2/s)
n-TiO212.33160.19172.521.04 × 10−16
p-TiO2-50011.9890.51102.495.99 × 10−16
p-TiO2-60012.01120.76132.774.93 × 10−16
Table 3. Comparison table of impedance of each part for p-TiO2-500, p-TiO2-600, and n-TiO2 sodium-ion battery and their sodium-ion diffusion coefficients.
Table 3. Comparison table of impedance of each part for p-TiO2-500, p-TiO2-600, and n-TiO2 sodium-ion battery and their sodium-ion diffusion coefficients.
SampleRs (Ω)Rct (Ω)Z′ (Ω)DNa+ (cm2/s)
n-TiO29.659358.8368.4594.88 × 10−19
p-TiO2-5008.853143.1151.9531.09 × 10−18
p-TiO2-6008.752235.9244.6521.60 × 10−18
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

Su, Y.; Li, J.; Liu, S.; Luo, S.; Li, Y.; Zhang, S.; Kong, X.; Huang, Q.; Lu, C. H1.07Ti1.73O4-Derived Porous Plate-like TiO2 as High-Performance Bifunctional Anodes for Lithium- and Sodium-Ion Batteries. Energies 2025, 18, 5077. https://doi.org/10.3390/en18195077

AMA Style

Su Y, Li J, Liu S, Luo S, Li Y, Zhang S, Kong X, Huang Q, Lu C. H1.07Ti1.73O4-Derived Porous Plate-like TiO2 as High-Performance Bifunctional Anodes for Lithium- and Sodium-Ion Batteries. Energies. 2025; 18(19):5077. https://doi.org/10.3390/en18195077

Chicago/Turabian Style

Su, Yabei, Juchen Li, Siyuan Liu, Silun Luo, Yuhan Li, Shaowei Zhang, Xingang Kong, Qiaogao Huang, and Chengyi Lu. 2025. "H1.07Ti1.73O4-Derived Porous Plate-like TiO2 as High-Performance Bifunctional Anodes for Lithium- and Sodium-Ion Batteries" Energies 18, no. 19: 5077. https://doi.org/10.3390/en18195077

APA Style

Su, Y., Li, J., Liu, S., Luo, S., Li, Y., Zhang, S., Kong, X., Huang, Q., & Lu, C. (2025). H1.07Ti1.73O4-Derived Porous Plate-like TiO2 as High-Performance Bifunctional Anodes for Lithium- and Sodium-Ion Batteries. Energies, 18(19), 5077. https://doi.org/10.3390/en18195077

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