Hyungu Kanga,
Heesook Roha,
Jongseo Leea,
Sang-Hyeon Parka,
Joohyeon Parkb,
Heonjae Jeong
bc,
Hyun-Ki Yoona,
Tae-Young Ahn
*a and
Yusong Choi
*a
aDefense Materials and Energy Development Center, Agency for Defense Development, Yuseong P.O. Box 35, Daejeon, 34060, South Korea. E-mail: tyahn84@gmail.com; richpine87@gmail.com; Tel: +82-42-821-3520 Tel: +82-42-821-2457
bDepartment of Electronic Engineering, Gachon University, 1342 Seongnam-daero, Seongnam, Gyeonggi 13120, South Korea
cDepartment of Semiconductor Engineering, Gachon University, 1342 Seongnam-daero, Seongnam, Gyeonggi 13120, South Korea
First published on 1st August 2025
The paradigm shift from conventional molten-salt electrolytes to solid-state garnet-type Li7La3Zr2O12 (LLZO) electrolytes in thermal batteries represents a critical advancement in high-temperature energy-storage systems. This study evaluated Ta- and Ga-doped LLZO electrolytes for FeS2/Li–Si thermal batteries, focusing on their structural stability and electrochemical performance at 500 °C. While Ga-doped LLZO exhibited superior ionic conductivity at 25 °C, Ta-doped LLZO demonstrated exceptional high-temperature stability. Ta-doped LLZO cells achieved longer discharge durations and higher energy densities than Ga-doped LLZO cells, which is attributed to the retained cubic phase and minimised interfacial degradation. Conversely, Ga-doped LLZO exhibited cubic-to-tetragonal phase transitions, Ga precipitation, and formation of impurities such as Ga2O3 and Li–Ga alloys, leading to 54% loss of ionic conductivity post-discharge. These results contribute valuable insights for the optimisation of solid-state electrolytes in thermal battery systems, suggesting that conventional room-temperature performance metrics may not translate directly to elevated-temperature operations.
When replacing molten salts with solid electrolytes, the material selection must balance ionic conductivity, thermal stability, and interfacial compatibility. Sulfide-based electrolytes (e.g. Li10GeP2S12) are known for their ultrahigh ionic conductivity (>10 mS cm−1 at 25 °C); however, their poor thermal stability and reactivity with Li anodes limit their high-temperature applications.5–7 Polymer electrolytes, e.g. polyethylene oxide–polytetrafluoroethylene composites, offer mechanical flexibility and processability but suffer from decomposition above 200 °C.8,9 Oxide electrolytes, e.g. Li0.33LaTiO3 and Li1.3Al0.3Ti1.7(PO4)3, provide moderate conductivity and thermal resilience but face challenges such as reactivity with Li anodes.10,11
Garnet-type Li7La3Zr2O12 (LLZO) has emerged as a promising alternative as it offers competitive Li+ ionic conductivity, broad electrochemical stability (>5.5 V vs. Li/Li+), and structural robustness at high temperatures.12–16 Furthermore, in contrast to sulfides and polymers, the use of LLZO eliminates toxic byproducts, making it uniquely suited for thermal battery systems requiring both safety and high performance at extreme temperatures. Native LLZO exists in a tetragonal phase (space group I41/acd) at room temperature, characterised by ordered Li+ sublattices and poor ionic conductivity (∼10−6 S cm−1). At elevated temperatures (∼650 °C), it transitions to a cubic phase (space group Iad), where a disordered Li+ distribution enables significantly increases the conductivity (>10−4 S cm−1).15,17 This phase transition underscores the potential of the material in battery applications but necessitates doping strategies to stabilise the cubic phase.
Aliovalent doping (e.g. Ta5+, Ga3+) at Zr4+ or Li+ sites introduces Li+ vacancies and adjustments in Li+ transmission pathways with regulation of the lattice parameters, disrupting Li ordering and lowering the cubic-phase stabilisation temperature.12,18–24 Among these, Ga-doped LLZO (Ga-LLZO) has demonstrated superior efficacy, displaying room-temperature ionic conductivity of ∼2 × 10−3 S cm−1.25,26 Ga3+ is substituted directly onto Li+ sites, which creates two vacancies on adjacent Li+ sites and disorder in the crystal lattice to maintain charge balance.12,23,27 This creates additional Li+ ion channels such as the 96h → 96h migration path.28 Moreover, compared with other dopants, the diminished coulombic repulsion between Ga3+ and Li+ ions lowers the energy barrier for localised Li+ diffusion, promoting efficient percolation of Li+ across extended lattice networks.27
In this study, we systematically evaluated Ta-doped LLZO (Ta-LLZO) and Ga-LLZO electrolytes in FeS2/Li–Si thermal batteries, focusing on their structural evolution, ionic transport, and discharge performance at 500 °C. By integrating experimental and computational methods, we demonstrated how dopant chemistry governs phase stability and interfacial reactions. Our findings challenge conventional electrolyte selection criteria by demonstrating that room-temperature conductivity metrics are unable to accurately predict high-temperature performance, emphasising the need for operational-condition-specific material design.
Ta0.5 and Ga0.2 were selected as the dopant concentrations for LLZO according to previous reports indicating optimal ionic conductivity at these compositions. For Ta-LLZO, studies have demonstrated that Li6.5La3Zr1.5Ta0.5O12 (Ta0.5-LLZO) has high conductivity (>1 × 10−3 S cm−1) owing to its increased Li+ vacancy concentration and lattice stability.19,20 Similarly, Li6.4Ga0.2La3Zr2O12 (Ga0.2-LLZO) achieves ionic conductivities above 1 × 10−3 S cm−1 by balancing Li+ vacancy formation.22,23,29 Sharifi et al.21 achieved a high ionic conductivity of 5.85 × 10−3 S cm−1 at 20 °C using a sol–gel combustion process. In the present study, the consistent use of Ta0.5-LLZO and Ga0.2-LLZO across all experiments ensured a direct, literature-supported comparison of dopant effects, as these concentrations are widely recognised as the concentrations for achieving optimal ionic conductivity in their respective systems.
LiCl (Vitzrocell Co., Ltd) and KCl (DUKSAN HIGHCHEM CORP.) powders were separately vacuum-dried and then mixed in eutectic proportions. The LK salt mixture was subsequently melted under an Ar atmosphere and mixed with MgO (Sigma-Aldrich). The cathode material was prepared by mixing FeS2 powder (Sigma-Aldrich) with the LK salt/MgO mixture and Li2O (DUKSAN HIGHCHEM CORP). For the anode, the LK salt/MgO mixture was combined with an Li–Si alloy (Gelon Energy Corp). The electrolyte preparation began by mixing the LK salt mixture with doped LLZO powder (POSCO JK Solid Solution, Pohang, Korea) in a mass ratio of 25:
75, which was essential to improve formability and facilitate large-scale processing, as cold pressing of pure LLZO presents challenges in fabricating large-diameter pellets. The cathode, anode, and electrolyte mixture powder were placed in a 56 mm-diameter mould and uniaxially pressed to form a circular disk. These components were then assembled to form the complete thermal battery unit cell with current collectors (STS304) and used in subsequent electrochemical testing and performance evaluation.
Discharge tests were conducted under controlled conditions to evaluate the electrochemical performance of the fabricated thermal battery unit cells. To minimise the contact resistance between components, the cells were placed in a press and subjected to a constant applied force of 250 kgf throughout the testing process. Prior to testing, the batteries were heated to 500 °C and allowed to stabilise for at least 10 min to ensure uniform temperature and complete melting of the LK salt, increasing the ionic conductivity. To evaluate the internal resistance of the thermal battery, pulse discharge tests were conducted under two distinct current-density conditions: 0.1 Aavg cm−2 (low) and 0.5 Aavg cm−2 (high). For the low-current-density discharge, pulse sequences consisted of a 2.45 A constant current for 8 s, followed by 0 A for 1 s and then 4.9 A for 1 s. Similarly, 12.25 A (8 s), 0 A (1 s), and 24.5 A (1 s) cycles were applied for high-current-density pulse testing. The voltage profile was recorded using a potentiostat/galvanostat system equipped with high-temperature-compatible probes.
To examine the influence of the Li–Si alloy on doped LLZO at elevated temperatures, Li–Si pellets with identical dimensions were fabricated via uniaxial pressing. The experimental setup comprised a Li–Si pellet placed beneath both Ta-LLZO and Ga-LLZO pellets in a stacked configuration. This assembly was placed in an Ar-controlled glovebox containing a furnace, where the two stacks were heated at 500 °C for 5 h.
Impedance measurements were performed using a high-temperature ion conductivity measurement system (TPU-005N, TeraLeader) capable of simultaneous pressure application and thermal control. Pellets were placed in a ceramic insulator and subjected to a constant pressure of 150 kgf during testing. Electrochemical impedance spectra were acquired using a BioLogic SP-50e potentiostat over the frequency range of 100 Hz to 1 MHz with an alternating-current (AC) voltage amplitude of 100 mV. Measurements were peformed from 25 to 100 °C, and the acquired impedance data were fitted using ZFit software (BioLogic) to extract resistance values from Nyquist plots.
We modelled an interface between LLZO(100) and Li13Si4(011) using VASPKIT,34 where the interfacial area was 25.2 × 18.0 Å2 with a total thickness of 36.4 Å and vacuum length of 10.0 Å. The interfacial gap was separated by 3.0 Å, and the lattice mismatch was 5.5%.
For AIMD simulations, owing to high computational demands, a reduced plane-wave cutoff energy of 400 eV was used, and the Brillouin zone was sampled only at the gamma point. AIMD simulations were conducted for 1 ps with a timestep of 1 fs, employing a Nosé–Hoover thermostat and a canonical ensemble (NVT) at 773 K.
To investigate the tendency of dopant precipitation from the doped LLZO, we calculated the defect formation energy (DFE). The precipitation sites of the dopant (Ta or Ga) were selected according to the lowest DFE value. The DFE is given by35
DFE = Edefect − Epristine + μdefect, | (1) |
μdefect = μDFTdefect + Δμ(T), | (2) |
Δμ(T) = ΔH(T) − TΔS(T). | (3) |
The DFE represents the thermodynamic stability of Ta and Ga dopants, indicating their tendency either to remain incorporated within the system or to be released. Lower DFE values correspond to a higher likelihood of dopant annihilation (precipitation) from LLZO, whereas higher values reflect enhanced stability in LLZO. Negative DFE values suggest that dopant release is thermodynamically favourable and can spontaneously occur. To compare the favourable regions, the DFEs were calculated for both bulk and interface structures. To provide thermal perturbation on structural relaxation, structures were obtained from AIMD simulations at an elevated temperature of 773 K.
Fig. 1(b) shows the Nyquist plots for Ta- and Ga-LLZO pellets, obtained at 25 °C after uniaxial pressing. The impedance spectra typically show two semicircles in the high- and intermediate-frequency regions, followed by a long diagonal tail in the low-frequency domain. Each semicircle corresponds to a parallel combination of resistance and constant-phase elements in the equivalent circuit model. While studies have suggested that the two semicircles correspond to the bulk and grain-boundary components,26,38,39 other reports have indicated that these features in the Nyquist plot cannot be easily separated.40–43 Therefore, the first semicircle in the high-frequency region is assigned to the combined resistance of bulk and grain-boundary components, representing the total solid electrolyte resistance. The second semicircle is attributed to the interfacial resistance between the solid electrolyte pellet and electrodes.41–43 The low-frequency tail represents the blocking effect resulting from the use of Pt electrodes, where ions accumulate at the electrode surface and cannot pass through, leading to increased double-layer capacitance.40,43–45 Therefore, the experimental data were analysed using an equivalent circuit model of (Rbulk/GBQbulk/GB)(RinterfaceQinterface)Qblocking to calculate the solid electrolyte resistance.46–48 The total conductivity (σtotal) was calculated using the thickness (t) and area (A) of the LLZO pellet: σtotal = t/(ARbulk/GB). The room-temperature ionic conductivities calculated using the total solid electrolyte resistance were 2.32 × 10−4 and 7.14 × 10−4 S cm−1 for Ta- and Ga-LLZO, respectively, which are consistent with previously reported findings.23,27,28,49
Moreover, EIS measurements were conducted at 50, 75, and 100 °C, and the results are presented as an Arrhenius plot in Fig. 1(c). The calculated activation energies were 0.288 and 0.266 eV for Ta- and Ga-LLZO, respectively, with the values falling within the typical range (0.20–0.40 eV) that has been reported for doped LLZO materials.23,27,28,49 The lower activation energy of Ga-LLZO may reflect its superior ionic conductivity compared with that of Ta-LLZO, as a reduced activation energy signifies lower energy barriers for Li+ migration. The temperature-dependent Nyquist plots for each composition are shown in Fig. S1.† According to both previous studies and our investigation, the room-temperature ionic conductivity of doped LLZO follows the trend of Ga-LLZO > Ta-LLZO.23,25–27,49
Fig. 2(a) and (b) illustrate the discharge profiles at constant current densities of 0.1 and 0.5 Aavg cm−2, respectively. Following the initial discharge period, repeated pulse patterns were applied to evaluate the dynamic response characteristics of the cells. Both cells exhibited polarisation phenomena during these rapid current transitions, with the voltage unable to fully respond to the abrupt current changes. The delayed voltage recovery may have resulted from the inability of Li ions in the solid electrolyte to reach the electrode surface with sufficient speed to sustain the electrode reactions.50,51
The open-circuit voltage (OCV) recorded immediately before discharge was approximately 1.92 V for every discharge test. This aligns well with previous reports presenting similar values for thermal batteries employing FeS2 cathodes and Li–Si alloy anodes at a discharge temperature of 500 °C.52–55 During discharge, three distinct voltage plateaus were observed for both cells. This multi-plateau behaviour has been extensively documented in previous research on Li–Si electrochemistry during Li–Si alloy delithiation.56–60 The phase transitions in Li–Si alloys at discharge typically develop a series of reactions involving the formation of intermediate phases as follows:
Li22Si5 → Li13Si4 → Li7Si3 → Li12Si7. | (4) |
Note that Li22Si5 exhibits extreme moisture sensitivity, presenting significant handling challenges even in dry room conditions (relative humidity below 3%).56,57 Consequently, this investigation employed Li13Si4 as the initial material for anode fabrication. Owing to the smaller quantity of the anode material compared with the cathode material, the observed plateaus correspond to the sequential phase transformations of Li13Si4 → Li7Si3 → Li12Si7.
The Ga-LLZO electrolyte cell exhibited an earlier voltage decline than the Ta-LLZO cell under both discharge conditions, resulting in consistently lower voltage values across all discharge regions. Furthermore, when considering the 1.4 V cutoff voltage,61 the Ta-LLZO cell demonstrated 1.23 times longer discharge at a low current density and 1.04 times longer discharge at a high current density. This indicated superior discharge performance with 1.23- and 1.04-times higher unit cell energy density for Ta-LLZO compared with Ga-LLZO. More detailed numerical values are presented in Table 1. The extended plateaus observed for the Ta-LLZO cell suggest stable electrochemical reactions. These plateaus are maintained by minimising voltage drops, which depends on a continuous reaction between electrodes and Li+ ions. This, in turn, relies on rapid ion transport through the electrolyte to deliver sufficient Li+ ions to the electrode surface during discharge. From these results, we can infer that the use of Ta-LLZO leads to a lower overall cell resistance at 500 °C compared with that of Ga-LLZO.
Discharge condition | Dopant type | OCV at 0 s [V] | Discharge time [s] | Specific capacity [mA h g−1] | Volumetric energy density of the unit cell [W h L−1] | Mass energy density of the unit cell [W h kg−1] |
---|---|---|---|---|---|---|
0.1 Aavg cm−2 | Ta-LLZO | 1.92 | 2114 | 740.6 | 46.9 | 250 |
Ga-LLZO | 1.92 | 1713 | 600.2 | 37.8 | 201 | |
0.5 Aavg cm−2 | Ta-LLZO | 1.92 | 379 | 663.9 | 40.0 | 213 |
Ga-LLZO | 1.92 | 364 | 637.6 | 38.1 | 203 |
The internal resistance of the cell can be calculated using the voltage differentials, which can be partially related to the ionic conductivity of the electrolytes. The internal resistance (Rt) was calculated using the current value (I) and difference between the OCV (VOC) and closed-circuit voltage (VCC), according to Ohm's law: Rt = (VOC − VCC)/I. The internal resistance calculated over time, which is shown in Fig. 2(c) and (d), confirms that cells employing Ga-LLZO exhibit higher internal resistance under both discharge conditions. The divergence between the two cells becomes markedly pronounced beyond 2500 s in Fig. 2(c) and 500 s in Fig. 2(d). This conclusively demonstrates that when Ga-LLZO is utilised in thermal batteries operating at 500 °C, it delivers inferior performance to Ta-LLZO.
Notably, the performance gap between the two electrolytes narrowed under high-current-density conditions. This convergence suggests dopant-dependent electrolyte stability under dynamic operating conditions and implies opposing degradation pathways. The total discharge time was approximately 1 h under 0.1 Aavg cm−2, whereas under 0.5 Aavg cm−2, it marginally exceeded 10 min. This increase in current density reflects accelerated reaction kinetics and shortened exposure time to the battery operational environment, potentially mitigating the degradation processes that occur during prolonged high-temperature operation. Extended exposure time to other battery components, such as the FeS2 cathode, Li–Si anode, and molten LK salt, at 500 °C may progressively destabilise Ga-LLZO. Conversely, the superior structural or interfacial stability of Ta-LLZO alleviates such degradation but is limited by accelerated electrochemical reaction kinetics under high current densities. The reasons for this unexpected performance reversal may include differences in interface stability at high temperatures, which induce an undesirable side reaction or temperature-dependent phase transformations within the doped LLZO structures.
![]() | ||
Fig. 3 (a) Photographs of the samples before and after heating; (b) Arrhenius plots of Ta- and Ga-doped LLZO in contact with the Li–Si alloy after being heated to 500 °C. |
Fig. 3(b) shows the Arrhenius plots of the ionic conductivity measurements for both types of LLZO pellets before and after heating with Li–Si alloy pellets. Interestingly, while the activation energies remained relatively constant for both materials after the heat-treatment process, significant changes in absolute conductivity values were observed. Ta-LLZO exhibited a moderate increase in ionic conductivity (2.32 × 10−4 to 2.86 × 10−4 S cm−1 at room temperature) following contact with the Li–Si alloy. This is attributed to the partial infiltration of Li-based alloys into the pellet structure, as evidenced by the localised blackening, which resulted from the reduction of LLZO.38,39,41,62 This Li penetration significantly reduces the distance that Li ions must traverse through the electrolyte, thereby improving overall ionic transport. Excessive Li infiltration can lead to short circuiting, compromising the functionality of the electrolyte. Conversely, Ga-LLZO demonstrated a substantial decrease in ionic conductivity (7.14 × 10−4 to 3.29 × 10−4 S cm−1 at room temperature) post-contact. Furthermore, Ta-LLZO exhibited increasingly superior ionic conductivity to Ga-LLZO as the temperature increased, as demonstrated by their conductivities at 373 K: 1.54 10−3 S cm−1 and 2.40 10−3 S cm−1, respectively. These findings align with the discharge test results shown in Fig. 2, where Ta-LLZO exhibits superior electrical performance to Ga-LLZO. Although Ga-LLZO may have also benefited from shortened Li+-ion pathways, performance-degrading factors (e.g. structural instability and side reactions) may have played a more dominant role, given the complex interplay of factors governing ionic conductivity.
Fig. 4(c) and (d) show the cross-sectional SEM images and EDS point analysis at 50000× magnification. In the Ga-LLZO electrolyte disks, Ga-rich regions with 47.41 at% (Spot 1) were detected after completion of the full-cell assembly with LK salt and subsequent discharge testing. A more detailed elemental composition of Spots 1 and 2, obtained from the EDS point analysis, is presented in Table 2.
Element (at%) | O | Zr | Cl | La | Ga |
---|---|---|---|---|---|
Spot 1 | 40.41 | 3.60 | 4.66 | 3.92 | 47.41 |
Spot 2 | 78.05 | 6.34 | 4.92 | 9.40 | 1.29 |
As the Ga-rich region was observed only after the discharge test, Ga precipitation is presumed to have occurred under high-temperature discharge conditions, probably owing to interfacial side reactions with other components of the thermal battery. The precipitated Ga is expected to exist primarily in the form of LiGaO2, Ga2O3, and the Li–Ga alloy upon interaction with a Li source.38,39,41,62,64 Although the present system utilised a Li–Si alloy rather than pure Li metal as the anode, Li–Si alloys exhibit sufficient reactivity to provide Li+ ions and Li metal, which may induce similar degradation mechanisms. According to the findings of Windmüller et al. and Kim et al., Ga-based oxides (e.g. Ga2O3) present at the interface can also react with Li+ ions under discharge conditions, forming Li2O and Li–Ga alloys.62,63 Because oxides such as Ga2O3 and Li2O typically possess low ionic conductivity, their formation at the interface can negatively impact the performance of the electrolyte. Moreover, the volume expansion accompanying oxide formation may lead to the development of microcracks in the solid electrolyte, further degrading the connectivity between particles and overall ionic conductivity.38,39 This hypothesis could also account for the post-discharge observation that the Ga-LLZO disk fractured readily under minimal mechanical impact.
![]() | ||
Fig. 5 DFE for Ga- and Ta-doped LLZO at 0 and 773 K. (a) LLZO bulk and (b) LLZO–Li13Si4 interfacial configuration. |
In contrast, Ta-LLZO maintained its structural integrity, confirming its enhanced interfacial stability. The superior interfacial stability of Ta-LLZO compared with that of Ga-LLZO stems from the self-limiting nature of the “oxygen-deficient interphase” (ODI)-layer formation during reaction with Li sources. In Ta-LLZO, the ODI layer remains confined to an ultrathin thickness as Ta dopants exhibit minimal segregation to the interface, suppressing Zr4+ reduction.65 The limited oxygen depletion may stabilize the bulk cubic phase, and suppress the phase transition to the tetragonal phase observed in Ga-LLZO. This enhanced electrochemical stability under operational conditions makes Ta-LLZO more suitable for practical applications in solid-state Li batteries where interface stability is paramount.
Prior to investigating the interaction between doped LLZO and Li–Si alloy, the thermal stability of the Li–Si alloy was examined (Fig. S3†). At room temperature, the Li–Si alloy primarily exhibited Li13Si4 (PDF# 89-9) stoichiometry. With the increase in temperature, Li-deficient phases emerged, reflecting a continuous decrease in Li content. Concurrent with the diminishing Li–Si alloy signals, new diffraction peaks such as Li2O (PDF# 65-2972), LiOH (PDF# 1-1021), Li4SiO4 (PDF# 19-727), and Li2SiO3 (PDF# 70-330) were observed, which resulted from reactions between trace amounts of oxygen/moisture and the Li–Si alloy.68,69
Fig. 7(a) and (b) show the in situ XRD patterns of Ta- and Ga-LLZO powders mixed with Li–Si alloy powder, subjected to the same thermal treatment protocol. The Ta-LLZO + Li–Si alloy mixture maintained remarkable structural stability, with no significant alterations to the primary LLZO structure. The newly formed peaks are attributed to either La2Zr2O7 or the Li–Si alloy-derived impurities identified in the isolated Li–Si alloy study.
![]() | ||
Fig. 7 In situ XRD patterns of the pristine (a) Ta-doped LLZO + Li–Si alloy mixture and (b) Ga-doped LLZO + Li–Si alloy mixture; (c) magnified view of the 29–35° 2θ region shown in (b). |
Conversely, the Ga-LLZO + Li–Si alloy mixture underwent structural transformations. The cubic LLZO structure, which remained stable up to 300 °C, transitioned predominantly to the tetragonal phase at 500 °C. This phase transformation may have been induced by the interaction between Ga-LLZO and the Li–Si alloy, resulting in Ga segregation in the LLZO structure and causing the lattice to revert from the cubic configuration to the tetragonal configuration.38,39 Furthermore, the segregated Ga formed oxide compounds including Ga2O3 (PDF# 11-342), along with Li–Ga alloys such as LiGa (PDF# 65-9159) and Li2Ga (PDF# 65-1383). The high-temperature-induced changes persisted after cooling, suggesting that the reaction was irreversible. Fig. 7(c) presents a magnified view of the 29–35° 2θ region shown in Fig. 7(b), allowing detailed peak analysis. Despite insignificant peak shifts relative to the tetragonal reference pattern, the breakdown of crystal symmetry associated with the cubic-to-tetragonal phase transition is clearly evident. Distinct Ga2O3 peaks emerged at 29.9° and 31.1° at 500 °C, confirming the decomposition of the Ga-LLZO structure when in contact with Li–Si alloy at elevated temperatures. The formation of these products is consistent with the predictions made according to the SEM/EDS analysis in Section 3.4.
The observed degradation of Ga-LLZO upon interaction with the Li–Si alloy at high temperatures exhibits detrimental effects on its electrochemical performance, particularly its ionic conductivity. Initially, upon exposure to the discharge environment, Ga was released from the lattice structure of Ga-LLZO owing to interfacial instability, as indicated by the DFE calculations. The leached-out Ga promotes rapid penetration of Li from the Li source into the bulk of the solid electrolyte via Li–Ga alloying, as evidenced by the prompt blackening of Ga-LLZO. This leads to a structural collapse from the cubic to tetragonal phase as Li captures oxygen atoms from the lattice. Under similar conditions, bandgap closure due to oxygen vacancies during this transition has also been reported in previous studies.39,70 Moreover, LiGaO2, which can form through the interaction of extracted oxygen and Ga during these processes, is expected to play a key role in the subsequent generation of impurities. LiGaO2 can initiate a violent reaction when in contact with Li, generating Li–Ga alloys alongside oxides such as Li2O and Ga2O3.38,39,41,62,63 Thus, the conversion from the highly conductive cubic phase to the less-conductive tetragonal phase, coupled with the formation of poorly conducting oxide species, creates significant barriers to Li-ion transport. Moreover, the volume changes associated with these phase transformations and formation of secondary phases can induce microstructural cracking, further compromising the mechanical integrity and electrochemical functionality of the solid electrolyte.
Fig. 8(a) shows the DSC analysis results of Ta- and Ga-LLZO mixed with FeS2 (cathode material). Regardless of the dopant type, both samples exhibited two endothermic peaks. A small peak appeared in the 400–420 °C range, whereas a larger peak spread across the 500–540 °C range. These peaks are attributed to the thermal decomposition of FeS2, which exhibits similar endothermic peaks in these temperature ranges.71 This suggests that the thermal events observed in the mixed samples were primarily due to inherent degradation of FeS2 rather than reactions with the LLZO materials.
![]() | ||
Fig. 8 DSC analysis of (a) Ta- and Ga-LLZO mixed with FeS2, (b) Ta- and Ga-LLZO mixed with LK salt, and (c) Ta- and Ga-LLZO mixed with the Li–Si alloy. |
The DSC analysis results of Ta- and Ga-LLZO mixed with LK molten salt are presented in Fig. 8(b). Similar to the cathode mixture results, the two samples exhibit consistent thermal behaviour regardless of the dopant. A small endothermic peak is observed between 275 and 295 °C, followed by a stronger endothermic peak spanning the 310–360 °C range. The second larger peak is assigned to the melting of the LK eutectic salt, which has a reported melting point of 353 °C in a pure state.72 However, in the presence of LLZO, the eutectic composition may be altered owing to surface chemical reactions. The LiCl–LiOH binary system exhibits melting within the temperature range of 269–292 °C at a composition of 32 mol% LiCl and 68 mol% LiOH.73 When the LiOH that is formed on the LLZO surface combines with the LK molten salt, the eutectic point of the resulting ternary system (LiCl–KCl–LiOH) is lower than that of pure LK, producing the smaller endothermic peak.
Fig. 8(c) shows significant differences in thermal behaviour between different doped LLZO materials when mixed with the Li–Si alloy (anode material). Both samples exhibited a broad exothermic peak between 150 and 300 °C, which is attributed to the reaction of protonated LLZO that may have formed despite our moisture-control efforts.74–76 Importantly, Ga-LLZO exhibited distinctly different thermal behaviour from Ta-LLZO at higher temperatures. Ga-LLZO exhibited a broad peak between 320 and 400 °C and a strong, intense peak above 500 °C. These peaks strongly suggest the occurrence of reactions between the doped LLZO materials and Li–Si alloy.
Cubic LLZO possesses high configurational entropy owing to its disordered Li+ distribution, while the tetragonal phase adopts ordered Li+ sites.12,15,17,77 The spontaneous transition (ΔG < 0) should be driven by a negative enthalpy change (ΔH < 0) that compensates for the entropy loss (ΔS < 0) associated with Li+ sublattice ordering in the tetragonal phase. Therefore, the observed exothermic DSC peaks may correlate with the cubic-to-tetragonal phase transition of LLZO, as confirmed by in situ XRD at 500 °C (Fig. 7(b) and (c)).
Furthermore, the aforementioned reaction mechanism between Ga-LLZO and Li metal, where precipitated Ga forms LiGaO2 or gallium oxides, subsequently reacting with Li to produce Li–Ga alloys and Li2O, can be summarised as follows:39
LiGaO2 + 5Li → Li2Ga + 2Li2O. | (5) |
The 0 K DFT calculations pertaining to the above reaction, as shown in Table S2,† confirm the exothermic nature of the solid-state process, with a negative reaction enthalpy of −41.4 kJ mol−1, which may correspond to exothermic peaks observed in the high-temperature region of the DSC curves. This reaction involves the solid phases of LiGaO2, Li, Li2Ga, and Li2O, as listed in the Materials Project database (data retrieved from the Materials Project for LiGaO2 (mp-5854), Li (mp-135), Li2Ga (mp-29210), and Li2O (mp-1960) from database version v2025.04.10), which exhibit well-predicted lattice parameters with <4% deviation in volume.78
The DSC results clearly demonstrate that Ta-LLZO has superior thermal stability and compatibility with the Li–Si alloy compared to Ga-LLZO. Given these points, the diminished ionic conductivity in Ga-LLZO stems from multiple concurrent mechanisms that outweigh any beneficial effects of Li penetration: (1) structural transformation of the doped LLZO from the cubic phase to less-conductive tetragonal phase, (2) formation of various oxide species that impede ionic transport, and (3) substantial volume expansion leading to microcrack formation within the solid electrolyte. The competition between these enhancing and deteriorating mechanisms ultimately results in a net reduction of ionic conductivity for Ga-LLZO. This suggests that Ta-doped LLZO is the preferred choice for high-temperature solid-state battery applications where thermal stability and minimal reactivity with electrode materials are critical.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ra03917g |
This journal is © The Royal Society of Chemistry 2025 |