Teng Suia,
Tao Wanga,
Qin Feng*ab,
Changlai Yuan
b,
He Qi
c,
Nengneng Luo
*a,
Xiyong Chen
a,
Zhenyong Cena and
Jiwei Zhai
*d
aState Key Laboratory of Featured Metal Materials and Life-Cycle Safety for Composite Structures, School of Resources, Environment and Materials, Guangxi University, Nanning 530004, China. E-mail: fengqin307@163.com; luonn1234@163.com
bGuangxi Key Laboratory of Information Materials, School of Material Science and Engineering, Guilin University of Electronic Technology, Guilin 541004, PR China
cBeijing Advanced Innovation Center for Materials Genome Engineering, Department of Physical Chemistry, University of Science and Technology Beijing, Beijing 100083, P. R. China
dKey Laboratory of Advanced Civil Engineering Materials of Ministry of Education, Functional Materials Research Laboratory, School of Materials Science & Engineering, Tongji University, No. 4800 Caoan Highway, Shanghai 201804, China. E-mail: apzhai@tongji.edu.cn
First published on 19th August 2025
Perovskite oxides have emerged as predominant materials in energy storage capacitor research. The development of lead-free dielectric capacitors featuring innovative architectures, high energy storage density, and superior high-voltage endurance could substantially advance this field. In this study, we synthesized La2Ti2O7 ceramics with compositional simplicity that exhibit promising industrial applicability. Their exceptional electric field resistance originates from three key characteristics: a wide bandgap (∼3.81 eV), uniform distribution of fine-grained dense microstructures, and stable titanium valence states. Notably, under an ultrahigh electric field of 1100 kV cm−1, the ceramics demonstrate minimal residual polarization (1.31 μC cm−2) while achieving maximum polarization intensity (∼19.98 μC cm−2) comparable to conventional strong ferroelectrics. Furthermore, these ceramics exhibit outstanding energy storage density (8.83 J cm−3), rapid discharge kinetics, and exceptional cyclic stability. This investigation pioneers the systematic evaluation of La2Ti2O7 ceramics for energy storage applications. The comprehensive results substantiate the material's exceptional potential for high-field energy storage systems.
The effective energy storage density Wrec, the total energy storage density Wtot, and the energy storage efficiency η of energy storage ceramics can be calculated using the following equations.9,10
![]() | (1) |
![]() | (2) |
![]() | (3) |
The intrinsic breakdown field strength (Eb) of dielectric materials is fundamentally governed by their bandgap energy (Eg). A larger Eg establishes a higher potential barrier for electron transitions between valence and conduction bands, thereby enhancing Eb. Three additional microstructural factors critically influence Eb: reduced dielectric permittivity, refined grain dimensions, and highly dense microstructures with uniform phase distribution. Conventional perovskite ceramics exhibit intrinsically limited Eb values due to their characteristically narrow bandgaps and elevated dielectric constants, as extensively documented in prior studies.11–13 This fundamental limitation has driven the exploration of alternative ferroelectric systems combining wide bandgaps with low-loss characteristics, including tungsten bronze-type architectures and bismuth-layered structured materials.14,15
Our comprehensive literature investigation identified La2Ti2O7 (LTO) as a compositionally simple ferroelectric material exhibiting low dielectric loss (tanδ) and wide bandgap energy (Eg).16,17 The substantial Eg (∼3.81 eV) suggests intrinsic potential for enhanced breakdown strength (Eb), enabling superior voltage tolerance and broader operational applicability. A review of the literature reveals that the structure of LTO is a layered perovskite structure (Fig. 1(a)), with the cells arranged in a 2 × 2 × 2 stacked configuration. The crystal structure comprises corner-sharing TiO6 octahedra coordinated with La3+ cations. Each structural layer contains four interconnected TiO6 units bridged by apical oxygen atoms (red spheres), with La cations occupying distinct crystallographic sites: interlayer positions (green spheres) between perovskite slabs and interstitial sites (white spheres) within octahedral frameworks. This stratified configuration enhances dielectric insulation through restricted interlayer charge transport, thereby achieving exceptional Eb. At the same time, the layered topology allows for a high intrinsic polarization of LTO ceramics, where a high intrinsic polarization strength (Ps) is the basis for obtaining a high Pm and a strong electric field is necessary to overcome various energy barriers, including those for 90° domain wall shifts, to achieve a Pm close to Ps. Furthermore, the inherent TiO6 octahedral rotational distortion explains the ferroelectric behavior observed in LTO ceramics.
![]() | ||
Fig. 1 (a) Crystal structure of LTO ceramics. (b–g) Schematic diagrams of energy storage characteristics of LTO ceramics. |
Contemporary research on La2Ti2O7 (LTO) has predominantly targeted luminescence enhancement via rare-earth ion doping (Pr3+/Tb3+).18 Intriguingly, LTO has emerged as a promising room-temperature multiferroic material, demonstrating concurrent ferroelectric and weak ferromagnetic ordering.19 Dielectric property optimization studies have focused on Bi3+ substitution strategies to enhance permittivity while suppressing dielectric loss (tanδ).20 Paradoxically, LTO's photocatalytic applications are constrained by its wide bandgap (Eg ≈ 3.81 eV), limiting optical activation to UV wavelengths. This inherent limitation has driven intensive bandgap engineering efforts through nitrogen doping and composite formation to extend visible-light responsiveness and improve photocatalytic efficiency.21 Crucially for energy storage applications, LTO's characteristically large Eg traditionally viewed as a photocatalytic disadvantage provides unique dielectric advantages. This fundamental material property enables the development of ceramics with exceptional intrinsic breakdown strength (Eb), forming the cornerstone of our innovative approach.
To the best of our knowledge, no systematic investigations have been reported on La2Ti2O7 (LTO) ceramics as novel lead-free dielectrics for high-field energy storage applications. While preliminary studies have incidentally observed energy storage characteristics in LTO-based flexible freestanding thin films,22 these findings remain complementary rather than contradictory to our core scientific objectives. This pioneering investigation establishes three critical breakthroughs: (1) experimental verification of LTO's exceptional breakdown strength (Eb > 1100 kV cm−1) enabled by its wide bandgap (Eg ≈ 3.81 eV); (2) development of phase-pure LTO ceramics via cost-effective solid-state reaction, achieving optimized microstructure through sintering protocol refinement (1250 °C/4 h); (3) as shown in Fig. 1, demonstration of record-high energy storage performance (Wrec = 8.83 J cm−3) with minimal polarization loss (Pm = 19.98 μC cm−2). And it exceeds the performance of some other classical systems. Microstructural analysis reveals some synergistic enhancement mechanisms (Fig. 1(e)–(g)): submicron-grained microstructure with uniform density distribution minimizes electric field concentration; stable Ti4+ valence state and wide Eg collectively suppress charge carrier migration. These combined attributes yield Eb values 2–3 times superior to conventional simple-composition ferroelectrics (BNT/BT/BF/NN systems), positioning LTO as a transformative candidate for next-generation high-voltage capacitor technologies.
Room-temperature Raman spectroscopy (100–900 cm−1) of LTO ceramics was systematically acquired (Fig. 2(c)). The spectrum exhibits broadened and overlapping phonon modes characteristic of complex crystalline environments. Thirteen distinct Raman-active phonon modes were resolved at 115 cm−1, 155 cm−1, 184 cm−1, 241 cm−1, 346 cm−1, 405 cm−1, 430 cm−1, 451 cm−1, 520 cm−1, 542 cm−1, 560 cm−1, 610 cm−1, and 796 cm−1, corresponding to the monoclinic P21 symmetry. Three distinct vibrational regimes were identified: low-frequency modes (<200 cm−1): dominated by interlayer La3+ cation displacements (115, 155, 184 cm−1); intermediate modes (200–400 cm−1): attributed to La–O bond bending vibrations (241, 346, 405 cm−1); high-frequency modes (>400 cm−1): originating from symmetric stretching vibrations within the TiO6 octahedral framework along O–Ti–O chains (430–796 cm−1).
This vibrational hierarchy reflects the layered perovskite architecture, where low-energy modes arise from interlayer interactions while high-energy modes correlate with intra-octahedral distortions.
The temperature-dependent dielectric properties of LTO ceramics were systematically investigated across wide frequency (1 kHz to 1 MHz) and temperature ranges (25–300 °C), as shown in Fig. S2. Both dielectric permittivity (εr) and loss tangent (tanδ) demonstrate exceptional thermal stability at low temperatures (<150 °C), exhibiting minimal temperature/frequency dependence with εr = 104–130 and tan
δ < 0.05. This ultrastable dielectric behavior maintaining ultralow loss coupled with moderate permittivity persists throughout the operational temperature range, fulfilling critical requirements for high-field energy storage applications. The material's unique combination of temperature-insensitive dielectric response and ultralow dissipation positions it as an ideal dielectric for capacitive energy storage systems.23
To elucidate the microstructure–property relationships in LTO ceramics, the samples were subjected to transmission electron microscopy (TEM) analysis. The grain boundaries appear well-defined and densely packed (Fig. 2(d)). A closer examination of the grain boundaries is presented (Fig. 2(e)), revealing the presence of atomic-level bonding, indicating a high level of density. Furthermore, the observation of Moire fringes (Fig. 2(f)) suggests an interference pattern resulting from the overlap of two different lattice orientations, further confirming the strong and tight bonding at the grain boundaries. Additionally, this lattice interaction leads to local polar lattice distortion, facilitating the formation of ferroelectric domains. The LTO samples have good crystalline properties, indicating lattice spacing and crystal orientation (Fig. 2(g)). The selective area electron diffraction (SAED) images obtained along various crystallographic axes (Fig. 2(h)–(k)) confirm the monoclinic and layered structure of LTO, as previously deduced from X-ray diffraction analysis. Furthermore, the distinctive morphology of ferroelectric domains is depicted (Fig. 2(l)–(o)). It is notable that LTO exhibits relatively large long-range ferroelectric stripe domains, alongside 90° striped domains. In general, 90° ferroelectric domains are less active and less likely to produce high polarization, but at high electric fields, 90° ferroelectric domains still have the potential to produce high Pm. The localized magnification (Fig. 2(n)) highlights the presence of domain walls. These unique orientations of ferroelectric domains significantly contribute to the ferroelectric response of LTO, consequently yielding a high Pm and low Pr under an external electric field. This attribute is beneficial for achieving high energy storage properties.
Systematic microstructural evolution of LTO ceramics was investigated through SEM analysis combined with thermal etching simulations across sintering temperatures (1100–1300 °C), as shown in Fig. S3. The samples sintered at 1150 °C showed submicron grains with high porosity, which seriously affected the dielectric reliability. In contrast, optimal sintering at 1250 °C yields fully dense microstructures featuring uniform grain growth, sharp grain boundaries, and negligible porosity. This defect-engineered microstructure enables record-breaking breakdown field strength (Eb = 1100 kV cm−1) through effective mitigation of electric field concentration hotspots.
To establish microstructure–property correlations in LTO ceramics, multiphysics finite element analysis (FEA) was employed to simulate the evolution of electric trees, dielectric constants, potential, and local electric field distribution of LTO samples sintered at different temperatures. The outcomes demonstrate that this method effectively integrates considerations of grain, grain boundary, and intrinsic dielectric constants to realistically simulate the experimental findings (Fig. 3 and S4, SI).
In addition, the grain size and dielectric constant of LTO ceramics at different sintering temperatures were recorded, which showed that the permittivity (εr) of the grains was obviously greater than that of the grain boundaries, indicating the resistance of the grain boundaries to the growth of electric trees (Table S2, SI). Further examination (Fig. 3(a)–(c)) illustrates that over time, electric tree growth predominantly occurs within the grains, where the electric field is lower, facilitating controlled growth and mitigating potential disruptions. Conversely, the grain boundaries serve as focal points for high electric field concentration, filtering the electric potential and impeding electric tree expansion. A comparative analysis with samples sintered at other temperatures (Fig. S4, SI) underscores that samples sintered at 1250 °C exhibit slow and restrained electric tree growth, characterized by well-defined, compact grains that rely on dense grain boundaries to inhibit the electric tree growth. Moreover, these samples manifest fewer regions of high potential and high electric field at the grain boundaries, resulting in a more uniform potential and electric field distribution, ultimately enhancing the breakdown field (Eb) of LTO ceramics and augmenting their energy storage performance.
The ultraviolet-visible (UV-vis) absorption spectra and corresponding bandgap energy (Eg) values of LTO ceramics were analyzed, as depicted in Fig. 4(a) and (b). The Eg value is indicative of the energy barrier for electron transition from the valence band to the conduction band. A higher Eg signifies a greater energy requirement for such transitions, thus hindering electron mobility and enhancing the ceramic's breakdown electric field strength. Deng recently demonstrated an inverse correlation between dopant concentration and bandgap in Co/Nb-modified Bi0.5Na0.5TiO3 ceramics, accompanied by substantially elevated leakage currents (>103 times higher than undoped counterparts).24 The Eg values were calculated using the Tauc equation in conjunction with the UV absorption spectrum.16,24
(αhν)m = A(hν − Eg) | (4) |
To further elucidate the high Eg of LTO ceramics, the energy band structure was computed using a hybrid functional approach. The calculations employed spin-polarized density-functional theory (DFT) to ensure accuracy.26,27 The projected augmented wave method, based on plane wave basis sets, was implemented in the Vienna ab initio simulation package (VASP).28,29 Exchange–correlation potentials were determined using a combination of PBE and the generalized gradient approximation (GGA),30 while the DFT-D3 model of Grimme was applied with van der Waals correction.31 A cutoff energy of 600 eV was used, and a Γ-centered Monkhorst–Pack grid with dimensions 3 × 2 × 5 was selected to sample the Brillouin zone integrals.32 To maintain a maximum force below 0.01 eV Å−1 for each atom, a fully relaxed crystal structure was utilized. The energy convergence criterion was set at 10−6 eV. The energy band structure was calculated using the Heyd–Scuseria–Ernzerhof (HSE06) method.33 The energy band structure and density of states are depicted in Fig. 4(c). The calculated Eg of LTO was 4.44 eV, which is higher than the experimental value. This discrepancy is attributed to the high-precision generalized HSE06 method's tendency to occasionally overestimate the bandgap. Although there is a deviation from the actual measured value. However, this analysis also confirms the high bandgap of LTO. Nevertheless, this analysis confirms the high bandgap of LTO. Examination of the figure reveals that the valence band maximum (VBM) and conduction band minimum (CBM) are closely located at the same wavevector k. For direct semiconductors like LTO, the k offset is minimal, indicating that electrons in the valence band are unlikely to transition to the conduction band. This property enables LTO to withstand a higher breakdown field strength.
To evaluate the breakdown electric field (Eb) of LTO ceramics, the experimental datasets were analyzed by implementing the Weibull probability model (Fig. 4(g)). This statistical framework facilitated the derivation of Eb values, aligning with established methodologies for interpreting material failure thresholds under electrical stress. The computational procedure involved systematic parameter optimization to ensure congruence between theoretical distributions and empirical observations, as documented in the graphical representation provided:10,34
Xi = ln![]() | (5) |
Yi = ln(ln(1/1 − Pi)) | (6) |
Pi = i/(n − 1) | (7) |
Impedance analysis of LTO ceramics was carried out in order to study in depth the effect of resistance on the ceramic characteristics. Impedance diagrams were generated at high temperatures between 480 °C and 540 °C, with frequencies spanning from 100 Hz to 1 MHz (Fig. 4(d)). Within these diagrams, Z′ denotes the real part of the impedance, while Z′′ signifies the imaginary component. The Nyquist plot clearly illustrates that the impedance curve of LTO ceramics exhibits a substantial radius. This characteristic suggests significant resistance and robust insulating qualities, advantageous for functioning at elevated breakdown field strengths. The curves form approximately semicircular arcs, and the inset illustrates equivalent circuit diagrams fitted using ZView software, based on the EIS data, indicating that the high resistance of LTO is primarily due to the grains and grain boundaries of the substrate. Additionally, LTO exhibits small grain size with dense and well-defined grain boundaries, underscoring the significant role of grain boundaries and grains in achieving high breakdown field strength (Fig. 3, S3 and S4, SI). Grain boundaries offer substantial resistance to the propagation of electrical trees, with the most significant contribution arising from these boundaries. As temperature increases, the arc radius of LTO ceramics gradually decreases, and the curve's center shifts towards lower frequencies, indicating thermally-induced relaxation phenomena and increased dielectric loss. Moreover, the variation of Z′′ with frequency displays a single Debye peak, which broadens and shifts to higher frequencies as temperature rises (Fig. 4(e)). This behavior suggests that LTO ceramics exhibit thermally-activated relaxation phenomena, potentially associated with oxygen loss in the ceramic samples and an increased presence of oxygen vacancies. The resistance R of the grains and grain boundaries of LTO follows the thermally-activated Arrhenius law:35
![]() | (8) |
The elemental composition and oxidation states of LTO ceramics were analyzed via temperature-dependent X-ray photoelectron spectroscopy (XPS), employing the C 1s peak for binding energy calibration. This methodological approach enabled precise characterization of valence configurations under varying thermal conditions, adhering to standardized protocols for surface-sensitive chemical analysis. The total elemental XPS spectra of LTO ceramics from 25 to 400 °C and the high-resolution XPS spectra of Ti 2p are shown (Fig. 4(h) and (i)). The sample is composed solely of three elements: La, Ti, and O, with no impurities present (C was included for calibration). This result is consistent with the XRD pattern (Fig. 2(a)), confirming the successful synthesis of LTO ceramics. The Ti 2p XPS spectra display well-defined symmetrical peaks for Ti 2p1/2 and Ti 2p3/2 at room temperature, with binding energies of 463.96 eV and 458.16 eV, respectively. The separation between Ti 2p1/2 and Ti 2p3/2 peaks is approximately 5.80 eV, indicating the presence of the Ti4+ state.36,37 As the temperature rises, two new peaks emerge at 456.01 eV and 461.06 eV at 200 °C, corresponding to the Ti3+ state.37 This change is attributed to the loss of oxygen atoms from the TiO6 octahedrons at elevated temperatures, leading to oxygen vacancies and a reduction in the valence state of some Ti ions. At 400 °C, further oxygen loss occurs, resulting in the appearance of a new Ti valence state. Two additional peaks at approximately 455.50 eV and 460.55 eV are observed, indicating the presence of the Ti2+ state.38,39 Additionally, the Ti 2p peak shifts towards higher binding energies with increasing temperature (Fig. 4(i)). This shift is due to oxygen loss in the LTO system, the formation of oxygen vacancies, and changes in the chemical states of Ti and O. It confirms the presence of Ti3+ and Ti2+ valence states and corroborates the changes observed in the impedance diagram. A summary of the changes in Ti element valence states with temperature is provided (Table S3, SI). The data indicate that the Ti element exhibits excellent temperature stability at low temperatures, maintaining its valence without excessive oxygen vacancy generation within a specific temperature range. This characteristic imparts the substrate with good insulating properties and high breakdown field strength.
The unipolar P–E characteristics of LTO ceramics were evaluated under various electric fields at 20 °C and 10 Hz (Fig. 5(a)), presenting a standard hysteresis loop diagram typical of ferroelectrics. Fig. 5(b) illustrates the corresponding Pm–Pr plot. LTO ceramics demonstrate a high breakdown field (Eb) of approximately 1100 kV cm−1, a high maximum polarization (Pm) of around 19.98 μC cm−2, and a low remnant polarization (Pr) of about 1.31 μC cm−2, indicating favorable energy storage properties. Under an electric field of 100 kV cm−1, the Pr of LTO ceramics is remarkably low at only 0.01 μC cm−2, while Pm measures 1.11 μC cm−2. As the applied electric field increases, Pm significantly rises to 19.98 μC cm−2, while Pr increases to 1.31 μC cm−2 but remains relatively low, resulting in decreased energy storage efficiency. The monopole I–E curve (Fig. 5(c)) under different electric fields at 20 °C and 10 Hz shows minor changes only when the applied electric field exceeds 500 kV cm−1. The P–E curve displays approximate linearity, influenced by the presence of small ferroelectric domains within the substrate. These domains can rapidly respond to the applied electric field, leading to quick domain flipping and a low Pr. However, during this period, polarization intensity growth slows. With further increase in the applied electric field, a prominent current peak gradually emerges, and the P–E loop begins to exhibit hysteresis, indicating the expansion of tiny ferroelectric domains under strong electric fields. The response to the electric field becomes somewhat hysteretic compared to earlier, resulting in an increase in Pr, substantial growth in polarization intensity, and a decrease in energy storage efficiency (η). However, the newly formed long-range ferroelectric domains are not yet fully developed, allowing Pr to remain relatively low. The energy storage of the ceramic shows a gradual increment with rising electric field strength, highlighting its strong dependence on high electric fields (Fig. S5(a), SI). Ultimately, at a high electric field of 1100 kV cm−1, the recoverable energy density (Wrec) reaches 8.83 J cm−3, while efficiency decreases from an initial 97% to 71%, indicating a reduction in efficiency but still maintaining satisfactory levels. Leakage current tests on LTO ceramics were also conducted (Fig. S5(b), SI). As the applied electric field increases, the leakage current of LTO ceramics only rises from 8.44 × 10−9 A cm−2 to 5.67 × 10−8 A cm−2. The extremely low leakage current density indicates excellent insulation performance, beneficial for achieving higher breakdown field strength.
The recently reported relationship between the recoverable energy density (Wrec) in energy storage ceramics and the number of cations in the system is illustrated in Fig. 5(d). It is evident that only two systems achieve Wrec values exceeding 10 J cm−3; however, their chemical compositions are exceedingly complex, and the preparation of such materials is both challenging and costly, making them unsuitable for large-scale industrial production. The electric field breakdown (Eb) values of various systems were calculated and compared, including those based on BF,40–43 BNT-based,44–47 BT-based,48–51 and NN-based52–55 (Fig. S5(c), SI). In contrast to the current scenario, where high energy storage properties require highly intricate chemical compositions, we have demonstrated that LTO ceramics can achieve both high Wrec and elevated Eb values using a simple chemical composition without the need for doping with other elements. Consequently, LTO ceramics exhibit significant potential for applications that demand simple chemical compositions to attain high values of Wrec and Eb.
Another advantage of dielectric capacitors is their exceptionally rapid charging and discharging rates, which result in significantly reduced charging and discharging times. The pulsed discharge characteristics of lithium titanate (LTO) ceramics were examined across multiple electric field intensities using a charge–discharge measurement system, as illustrated in Fig. 6(a). A fixed resistive load (RL) of 10260 Ω was integrated into the experimental configuration to analyze current behavior during non-oscillatory energy release processes. According to the equation (Wd = Wtol(RL + ESR)/RL), it is evident that when the applied load resistance is much greater than the equivalent sample resistance (ESR), it facilitates efficient energy storage aggregation within the load resistance. However, an excessively large load resistor may hinder the discharge rate. Therefore, a moderate resistance of 10
260 Ω was selected. Additionally, the variation curves of energy density (Wd) over time under different applied electric fields were plotted (Fig. 6(b)). The value of (Wd) can be obtained using eqn (9).24
![]() | (9) |
With the increase of the applied electric field from 100 kV cm−1 to 600 kV cm−1, it can be observed that t0.9 values stay below 1.5 μs, indicating an ultra-high instantaneous discharge rate of the sample (Fig. 6(b) and S6(a), SI). Concurrently with the augmentation of the externally applied electric field strength, the observed energy density (Wd) demonstrates a significant increase, progressing from an initial value of 0.05 J cm−3 to a peak measurement of 1.95 J cm−3. The measured Wd values under different applied electric fields align closely with the Wrec values calculated from the P–E loop at the identical electric field. The slight discrepancies arise due to the energy storage loss resulting from the equivalent resistance of the sample within the line, which is an unavoidable and negligible loss.
To investigate the stability of LTO ceramics' energy storage characteristics under conditions beyond room temperature, several properties of LTO were tested at different temperatures. The XRD curves at different temperatures (Fig. 6(c)) indicate that LTO exhibits high structural stability without any structural changes. Wrec, Wtot, η (Fig. 6(d)), and Pm, Pr (Fig. 6(e)) remain unaffected by temperature fluctuations.
Under an applied electric field intensity of 600 kV cm−1 and a cyclic frequency of 10 Hz, polarization–electric field hysteresis loops (P–E curves) were systematically evaluated across a broad spectrum of thermal conditions. It is noteworthy that the P–E curves of LTO ceramics within the range of 20–100 °C remain almost unchanged, indicating robust stability (Fig. 6(f)). Several properties of LTO at different frequencies were also tested. Wtot, Pm, and Pr exhibit a slight decrease with increasing frequency, while η increases from 74% to 84% with the rise in frequency. Wtot, Pm, Pr, and η all stabilize when reaching above 150 Hz (Fig. S6(b) and (c), SI), with Wrec maintaining high stability. The P–E curves over a wider frequency range were tested at 20 °C, and the curves exhibit minimal changes with increasing frequency (Fig. S6(d), SI). This demonstrates that LTO ceramics exhibit excellent temperature and frequency stability of their energy storage characteristics over a broader temperature and frequency range.
Supplementary information is available. See DOI: https://doi.org/10.1039/d5ta05234c.
This journal is © The Royal Society of Chemistry 2025 |