Open Access Article
Heather A. Ritchie
,
Ioanna M. Pateli
,
Oxana V. Magdysyuk
,
Aaron B. Naden
,
Heitor S. Seleghini
,
Federico Grillo
,
Gavin Peters,
Sharon E. Ashbrook
,
John T. S. Irvine
and
Venkataraman Thangadurai
*
EaStCHEM, School of Chemistry, University of St Andrews, North Haugh, St Andrews, KY16 9ST, UK. E-mail: vt36@st-andrews.ac.uk
First published on 1st October 2025
Garnet-type electrolytes are a promising class of solid-state oxide materials for next-generation lithium batteries. In this study, parent-phase garnets, Li5La3M2O12 (M = Nb, Ta, Nb/Ta), were synthesised via solid-state reaction at 900 °C and 1100 °C, enabling comprehensive structural and chemical stability characterisation. Powder X-ray diffraction (PXRD) and Raman spectroscopy confirmed single-phase garnets with Ia
d symmetry, while scanning electron microscopy (SEM) imaging revealed improved densification at 1100 °C. Electrochemical impedance spectroscopy showed the high ionic conductivities of 4.8 × 10−5 S cm−1 at 55 °C for Nb and Ta co-substituted Li5La3NbTaO12 prepared at 1100 °C. X-ray photoelectron spectroscopy and thermogravimetric analysis identified a surface lithium carbonate layer formed under ambient conditions, which was not discussed in the original garnet reported in 2003. Solid-state nuclear magnetic resonance spectroscopy provided an insight into the lithium environment in the surface and bulk of the samples and confirmed aluminium contamination in samples sintered at the base of alumina crucibles at 1100 °C, with LaAlO3 identified as the dominant secondary phase, corroborated by PXRD, SEM, and energy dispersive X-ray spectroscopy analysis. Nb-based garnet showed the most severe reaction with the alumina crucible. The use of a sacrificial mother powder of targeted garnet oxides, an approach commonly used for the preparation of Li-garnets, or sintering at 900 °C effectively reduced Al contamination. This work delivers a detailed evaluation of the parent-phase garnet, offering renewed insights into phase and structure two decades on from its initial development in 2003.
In 2003, Thangadurai et al.5 reported lithium-ion conducting original garnet-type oxides with nominal composition Li5La3M2O12 (M = Nb, Ta). This work highlighted that Li5La3Ta2O12 is chemically stable to lithium metal, which is a major advantage over many of the other promising solid electrolytes, such as La0.67−xLi3x□0.33−2xTiO3 (□ = A-site vacancy; x = 0.1 member exhibits bulk ionic conductivity of about 10−3 S cm−1 at room temperature) and Li1.3Ti1.7Al0.3(PO4)3.5 Cussen demonstrated through neutron diffraction study the three-dimensional network of Li diffusion within the garnet structure and highlighted the role of partially occupied octahedral sites in facilitating Li mobility.6 The combination of high ionic conductivity, chemical stability, and compatibility with lithium metal attracted significant interest, and over two decades on from their initial conception, the Li5-composition has earned the title of ‘parent-phase lithium garnet’ electrolytes. Cerium-doping in the parent phase Li5La3Nb2O12 was shown to enhance ionic conductivity by nearly two orders of magnitude, representing the highest value reported among the Li5 garnet family.7 The development of further garnet-like structures with enhanced conductivities was achieved by partially substituting the trivalent La3+ ions with divalent cations, leading to an increased lithium-ion concentration within the structure. This modification gave rise to a new subset of garnet-type electrolytes with the general formula Li6ALa2M2O12 (A = Mg, Ca, Sr, Ba, Sr0.5Ba0.5).8 A breakthrough in highly conducting garnet-type realised in 2007 when Murugan, Thangadurai and Weppner reported the cubic Li7La3Zr2O12 (LLZO), which showed a bulk conductivity of 3 × 10−4 S cm−1 at 25 °C, over an order of magnitude higher ionic conductivity than that of parent Li5La3M2O12 garnets.9 Karasulu et al.10 used experimental and computational ab initio analysis to understand the structure of doped and undoped cubic LLZO and demonstrated that Al and Ga dopants preferentially occupy Li1 sites in garnet, giving rise to distinct NMR signatures associated with local distortions.
Following this discovery, subsequent studies reported the formation of a less conductive tetragonal phase when sintering was performed at lower temperatures and in Pt crucible,11 indicating that the garnet-type structure is highly dependent on both the synthesis and processing conditions. The development of the Li7-phase marked a significant advancement, offering enhanced ionic conductivity and sparking a wave of research into the optimisation and application of garnet-type electrolytes, a field that is actively explored today.12–14 The surface chemistry of an SSE directly impacts the solid–solid interfacial properties in a solid-state battery, which in turn can significantly impact the battery performance. Initially, garnet-type electrolytes were thought to be chemically stable in ambient conditions; however, it has now been highlighted that surface layers of Li2CO3 formation in air results in low ionic conductivities and low densities.15 Further, the Li2CO3 layer is known to affect the wettability of Li metal on the electrolyte surface,16 which creates challenges during cell assembly.
The aim of this revisitation is to conduct a comprehensive investigation of the parent-phase garnet electrolyte Li5La3M2O12 to build a deep fundamental understanding of its structural, chemical, and electrochemical properties. This study offers a renewed evaluation of this emerging material through the synthesis of original Li5La3M2O12 and 1
:
1 solid solution of Li5La3Nb2O12 and Li5La3Ta2O12 and revisits the material purity through state-of-the-art solid-state characterisation methods. Whilst investigating surface contamination that arises during processing, particularly Li2CO3 formation due to air exposure, and examining intrinsic Al-doping introduced from alumina crucibles during high temperature processing, which have not been as widely studied for the parent phase Li5-composition garnet, unlike Li-stuffed LLZO and its derivatives.
17 was used for Rietveld refinement analysis.18,19
Scanning electron microscopy (SEM) images were obtained using the JSM-IT800 Schottky Field Emission Scanning Electron Microscope by JEOL coupled with JEOL Energy Dispersive X-ray Spectrometer (EDS). Images captured in both secondary and back-scattered electron detectors were made using 5 kV and 10 mm working distance. For elemental composition analysis, 15 kV and analysis current mode were used.
X-ray photoelectron spectrometry (XPS) measurements were collected in an ultrahigh vacuum, with base pressure ca. 1 × 10−10 mbar, chamber using a Kratos Axis Ultra DLD photoelectron spectrometer with monochromatic Al Kα (1486.6 eV) anode operating at 4 mA × 15 kV. The system was operated in hybrid mode with the slot aperture, giving a measuring area of ca. 700 micron × 300 micron. Charge compensation was used during data collection; the binding energy (BE) scale was referenced to adventitious carbon species at 285.0 eV and BE values are reported ± 0.1 eV. Peak fitting was carried out using the CasaXPS software,20 with Kratos supplied sensitivity factors, using Gaussian–Lorentzian functions, after subtracting a Shirely-type background.
Solid-state NMR spectra were recorded using a Bruker Avance III spectrometer equipped with a 14.1 T wide-bore magnet. Powdered samples were packed into 3.2 mm outer diameter ZrO2 rotors from Bruker and rotated at a magic angle spinning (MAS) rate of 20 kHz. The 27Al MAS NMR spectra were recorded using a short-flip angle pulse (shorter than π/20) with a radiofrequency field strength of ∼100 kHz to ensure that the spectra are quantitative and are the result of averaging 256 transients with a recycle interval of 5 s. The 7Li MAS NMR spectra were recorded using a short-flip angle pulse (shorter than π/12) with a radiofrequency field strength of ∼106 kHz to ensure quantification and are the result of averaging 32 transients with a recycle interval of 10 s. 27Al NMR spectra and 7Li NMR spectra are shown relative to 1.1 M Al(NO3)3 in D2O and 1 M LiCl in D2O, respectively. For the 27Al NMR spectra acquisition, Al(acac)3 (δiso = 0.0 ppm) was used as a secondary reference.
Scanning transmission electron microscopy (STEM) measurements were performed on an FEI Titan Themis operated at 200 kV, equipped with a CEOS DCOR probe corrector, a SuperX energy dispersive X-ray spectrometer (EDX) and a 4k × 4k Ceta CMOS camera. High-Angle Annular Dark-Field (HAADF) images were acquired with a probe convergence angle of 20 mrad and inner/outer collection angles of 60 and 200 mrad, respectively. Selected area electron diffraction patterns (SADPs) were indexed and fitted using Recipro.21 Samples for transmission electron microscopy were prepared by dispersion of powder in methyl isobutyl ketone (largely immiscible with water to avoid formation of lithium carbonate) before drop casting onto lacey C/Cu grids purchased from EM Resolutions Ltd, UK.
d space group, as the main phase. Fig. S1–S3 (see SI) show the PXRD data for Nb, Ta and Nb/Ta phases synthesis and corresponding precursors compounds.
The unit cell parameters for all samples were found to be 12.73–12.77 Å. These values are comparable to the literature,5,23 further confirming that cubic garnet structures were synthesised. Additionally, it was observed that the garnets containing Nb showed a slight decrease in lattice parameter at the higher sintering temperature, whereas the Ta only sample showed the same cell parameter at both sintering temperatures. The unit cell of the garnet may decrease due to the formation of small amounts of impurity as a secondary phase, preventing the formation of the garnet in the desired stoichiometric ratio or if a material decomposes at elevated temperatures, in turn changing the structure. However, the observed changes in the unit cell parameter are found to be rather small value, and therefore, the proposed compositional changes are only speculative, requiring further investigation.
Investigation of the phase formation of Li5La3Nb2O12 began using the mixture, which had been pre-annealed ex situ in the laboratory oven for 10 hours at 750 °C. Therefore, in the starting powder at room temperature (RT), a small amount of the targeted phase Li5La3Nb2O12 was present. Thus, even if considerable amounts of Li5La3Nb2O12 had formed during the ex situ pre-annealing step at 750 °C, it was largely decomposed to La2LiNbO6 (84 wt%) and Li3NbO4 (9 wt%) on cooling back to RT. The marginally small amount of targeted phase Li5La3Nb2O12 (4 wt%) as well as minor impurities La2O3 (1 wt%) of non-reacted high-temperature decomposition phase of La(OH)3 and La3NbO7 (2 wt%, probably formed at 750 °C during ex situ pre-annealing and preserved during cooling to RT) were also observed in the starting pre-annealed ex situ material.
It is observed that the ex situ pre-annealed at 750 °C mixture quickly transforms before 700 °C (Fig. 2), giving 54 wt% of the targeted Li5La3Nb2O12 phase at 700 °C, while undesirable La2LiNbO6 completely disappears by 800 °C via decomposition to La2O3 and other possible amorphous phases. The target phase Li5La3Nb2O12 reaches 88 wt% at 1075 °C before unexpected complete decomposition into La2LiNbO6 and Li3NbO4 after 1075 °C. Although the very little impurity of La3NbO7 was formed during the pre-annealing step, it reacted with other phases before reheating during in situ PXRD to 700 °C. These results allow us to speculate that using a very high temperature of 1100 °C for the synthesis of Li5La3Nb2O12 may not be necessary, and extended annealing at a lower temperature can be preferable and more convenient.
Although the maximal amount of Li5La3Nb2O12 obtained during in situ PXRD experiment was 88 wt%, it is possible that the slightly different conditions, which are necessary for in situ measurement prevented 100 wt% transformation. During ex situ synthesis, the powder is pressed and a larger volume of the powder is used, while during in situ PXRD synthesis the powder was loose and a small amount of powder (cylinder of dia. 1 mm and of length 10 mm) was used. This scaling down of the material during synthesis may have prevented the full transformation because part of the powder may have stuck to the capillary tube walls and, therefore, was not involved in the reaction. To further understand whether the multiple annealing steps between 700 °C and 1100 °C (during which PXRD data were recorded) had a substantial influence on the transformation behaviour, the pre-annealed ex situ at 750 °C mixture was used for a different temperature profile: heating from RT to 700 °C for a short time to record one diffraction data at this temperature and subsequent heating up to 1100 °C and holding at that temperature for 4 hours while continuously recording diffraction data (Fig. 3).
The room-temperature results for the freshly made ex situ pre-annealed mixture were almost identical to the previous mixture (Fig. 2 and 3), demonstrating 88 wt% of La2LiNbO6 and 8 wt% of Li3NbO4 with only 4 wt% of the targeted phase Li5La3Nb2O12. The first data point at 700 °C indicated that the transformation proceeded in the identical way to the previous experiment, with the targeted phase Li5La3Nb2O12 reaching 48 wt%. After that, the mixture was fast heated up to 1100 °C and measured at this temperature during hold. Very unexpected results indicate 91 wt% of targeted phase Li5La3Nb2O12 within 1 hour after reaching 1100 °C and subsequent complete decomposition of Li5La3Nb2O12 by the end of 3 hours hold at 1100 °C. The decomposition products were La2LiNbO6 (87 wt%) and Li3NbO4 (13 wt%), i.e. the same mixture was formed in situ after 3 hours hold at 1100 °C, as the starting mixture after ex situ pre-annealing at 750 °C. Using such a pre-annealed mixture allowed to synthesise the almost pure targeted phase Li5La3Nb2O12 (with restrictions imposed by scaling down the reaction for the in situ PXRD experiment). To check whether this mixture can again produce the targeted phase Li5La3Nb2O12, a second heating profile was used for in situ PXRD measurements (Fig. 3). Despite the same temperature profile for the second cycle, only decomposition products La2LiNbO6 and Li3NbO4 were observed without any variation in their amount. Attempt was made to synthesise the targeted phase Li5La3Nb2O12 during in situ PXRD measurements without an ex situ pre-annealing step at 750 °C in an ordinary laboratory oven. Two different experiments were performed with: (i) only 2 annealing steps (700 °C and 1100 °C) to simulate the standard ex situ synthesis process (Fig. 4) and (ii) multiple intermediate temperature steps to clarify the fine details of the solid-state synthesis process (Fig. 5).
In situ PXRD measurements of as-prepared mixture (Fig. 4) contained at RT the mixture of starting materials Nb2O5, Li2CO3, La(OH)3 (formed during preparation of the initial mixture from the correctly weighted La2O3 via hydration in air). Fast heating up to 700 °C indicated the fast formation of the targeted phase Li5La3Nb2O12 (25 wt%) with impurities La2LiNbO6 (7 wt%) and Li3NbO4 (9 wt%) and unreacted La2O3 (60 wt%). Subsequent annealing at 700 °C for 3.5 h had a minor impact on these formed phases. Ramping the temperature to 1100 °C allowed to increase the concentration of the targeted Li5La3Nb2O12 to 47%. Previous experiments (Fig. 2 and 3) demonstrated that longer annealing at 1100 °C would result in the complete decomposition of Li5La3Nb2O12, hence, the mixture was cooled to RT through one intermediate short temperature step at 700 °C to record one diffraction pattern. In situ diffraction data indicate almost complete decomposition of the targeted phase already on cooling to 700 °C (Fig. 4), which perfectly correlates with results from ex situ pre-annealing (first RT results on Fig. 2 and 3). Second identical heating cycle (Fig. 4) was performed for in situ PXRD with the same temperature profile, but the results indicated that the more stable phase of La2LiNbO6 does not participate in the reaction, and its amount remained almost unchanged during the second heating cycle. The amount of the targeted phase Li5La3Nb2O12 slightly increased during the second heating cycle to 5 wt%, but the complete decomposition of this phase was observed on the second heating to 1100 °C.
There is an acceptable correlation between completely in situ synthesis during PXRD measurements (Fig. 4) and in situ synthesis with ex situ pre-annealing step at 750 °C. The formation and decomposition of the targeted phase Li5La3Nb2O12 at 700 °C and the irreversible decomposition of the targeted phase Li5La3Nb2O12 during the second heating cycle, and similar decomposition phases and intermediate phase during all heat treatments were observed. The interesting observation is the possibility to synthesise the almost pure targeted phase Li5La3Nb2O12 from the mixture of La2LiNbO6 and Li3NbO4 after an ex situ pre-annealing step at 750 °C (Fig. 2 and 3), while using the same mixture after heating it up to 750 °C makes synthesis of the targeted Li5La3Nb2O12 phase impossible (Fig. 3 and 4).
To clarify the fine details of the solid-state synthesis process, the as-prepared mixture was used for in situ synthesis during PXRD measurements with multiple temperature steps (Fig. 5). As expected, the initial as-prepared mixture at RT (Fig. 5) contained the starting materials Nb2O5, Li2CO3, La(OH)3 (formed during preparation from La2O3). The diffraction peaks from a crystalline Nb2O5 disappear after 500 °C. The formation of LiNbO3 can be assigned to the reaction between Nb2O5 and Li2CO3 around 500 °C. At 650 °C, the transformation of LiNbO3 into Li3NbO4 is observed; this transformation was previously observed during the synthesis of lithium niobates doped by transition metal.24 An important observation is a substantial amorphisation of the starting materials (Li2CO3, La(OH)3, Nb2O5) around 400 °C: all diffraction peaks at this temperature are very broad; La(OH)3 is undergoing the transformation to crystalline La2O3 via a substantially amorphous intermediate state (Fig. 5). At 500 °C, the highly crystalline La2O3 and LiNbO3 with low-crystalline Nb2O5 are observed. Subsequent heating up leads to decreasing the amount of LiNbO3 at 650 °C (with subsequent disappearance after heating to 875 °C), formation of Li3NbO4, La2LiNbO6, and Li5La3Nb2O12. The targeted Li5La3Nb2O12 is formed only within a temperature range 650–750 °C. During this slow heating process, the amount of this phase remains very small due to the competitive formation of more stable Li3NbO4 and La2LiNbO6 phases. Thus, the very slow heating via multiple annealing temperature steps, to record the diffraction patterns at each temperature, changed the synthesis conditions considerably and prevent the detailed analysis of the synthesis process by in situ PXRD. It is important to underline that no quantifiable amount of LaAlO3 was observed during in situ PXRD experiments.
To analyse the possible influence of the Ta substitution, additional in situ PXRD measurements were performed during synthesis of Li5La3NbTaO12 (Fig. 6). The initial mixture was pre-annealed ex situ in the laboratory oven for 10 hours at 750 °C, and the resulting powder at RT showed a significant amount of targeted phase Li5La3NbTaO12 (54 wt%) in contrast to all ex situ pre-annealed experiments for synthesis of Li5La3Nb2O12 (Fig. 2 and 3). Subsequent heating to 700 °C during in situ PXRD measurements allowed for an increase in the amount of targeted phase to 85 wt%, and further heating allowed achieving an almost pure sample with 94 wt% of targeted Li5La3NbTaO12 at 975 °C. Further heating up resulted in the partial decomposition into more stable La2LiNb0.5Ta0.5O6 and other possible amorphous phases. Li5La3NbTaO12 and La2LiNb0.5Ta0.5O6 were observed by in situ PXRD as crystalline phases at 1050 °C and above, as well as during subsequent cooling to RT. Still, a substantial amount of the targeted Li5La3NbTaO12 (85 wt%) was observed at RT at the end of the in situ synthesis (Fig. 6), which contradicts the results for the targeted phase Li5La3Nb2O12 (Fig. 2–4).
The Raman spectra further confirm formation of a cubic garnet structure across both sintering temperatures for each composition. This is evidenced by the presence of broad, characteristic peaks associated with the garnet structure.25 Additionally, due to Raman spectroscopy's sensitivity to local bonding environments, subtle variations in M-site coordination can be resolved, allowing differentiation between Nb and Ta samples. The spectra indicate the successful formation of a mixed M-site garnet, which is not easily distinguishable by PXRD. It should be noted that for all compositions after sintering at 1100 °C, a shoulder peak emerges at 110 cm−1, indicating the existence of small amounts of lower symmetry tetragonal phase.
| Element and orbital | Li5La3Nb2O12 | Li5La3NbTaO12 | Li5La3Ta2O12 | |||
|---|---|---|---|---|---|---|
| BE (eV) | ASC (%) | BE (eV) | ASC (%) | BE (eV) | ASC (%) | |
| La 3d5/2 | 833.0 | 1.3 | 832.9 | 0.4 | 833.1 | 0.9 |
| O 1s | 531.7 | 47.8 | 531.8 | 49.9 | 531.7 | 45.3 |
| 528.7 | 7.1 | 528.7 | 3 | 529.0 | 6.9 | |
| Nb 3p3/2 | 363.7 | 0.8 | 363.8 | 0.2 | — | — |
| C 1s | 289.9 | 15.7 | 289.9 | 18.3 | 290.0 | 18.3 |
| 285.0 | 2.7 | 285.0 | 2.5 | 285.0 | 2.5 | |
| Ta 4d5/2 | — | — | 228.5 | 0.3 | 228.7 | 0.8 |
| Li 1s | 55.0 | 24.7 | 55.1 | 25.4 | 55.0 | 27.6 |
When comparing the M-site elements (Nb and/or Ta), differences in both binding energy and atomic percentage are observed for the mixed composition vs. both the single substituted samples. In the pure Nb and Ta samples, the M-site elements are present at approximately 0.8 atomic %, whereas the mixed composition shows a lower value of 0.5 atomic %. This may indicate a reduced amount of the garnet phase exposed at the surface in the mixed sample, likely due a thicker layer of lithium carbonate. Interestingly, the mixed M-site sample also exhibits slightly lower binding energies for both Nb and Ta compared to the single-substitution samples. This could indicate that Nb and Ta may reside in a more electron-rich environment within the mixed-phase garnet, potentially due to modified local bonding environments or variations in oxygen coordination. However, the observed binding energy shifts of 0.1 eV for Nb and 0.2 eV for Ta, are relatively small, suggesting that these changes are not significant, and perhaps just within the experimental error.
The O 1s core level peaks show a profile which is best fitted with two components (FWHM ca. 1.7 eV): the higher BE component, ca. 531.7 eV, is assigned to oxygen belonging to the carbonate anion, whereas the lower BE component, ca. 529 eV, is assigned to the garnets lattice oxygen.34 The C 1s core level region is fitted with two components: the higher BE component, ca. 289.9 eV (FWHM ca. 1.2 eV), is assigned to carbon belonging to the carbonate anion, whereas the lower BE component, 285 eV (FWHM ca. 1.6 eV), is assigned to adventitious carbon,35 and used as internal reference for the BE scale. For all the investigated garnets, the Li 1s core level peak is fitted with a single component with maximum at 55.0 eV (FWHM ca. 1.5 eV) and assigned to the presence of Li(I).30,31 Overall, the presence of the garnet phase elements is confirmed, however, the measured atomic concentrations deviate from those expected for a bulk stoichiometric garnet phase, particularly showing an excess of lithium, carbon and oxygen. Since XPS is a highly surface-sensitive technique, probing in the range of about 30 nanometres, it is particularly susceptible to detecting surface composition. The presence of excess Li, C and O, suggests that the garnets' surfaces are coated in a layer of Li2CO3 which is likely due to surface reaction of the exposed Li with atmospheric H2O and CO2.
The resistance derived from the Nyquist plot at low-frequency intercept to the real axis, the thickness of the pellet, l, and the surface area of the electrode, A, were used to calculate the ionic conductivity, σ, of the investigated three garnet-type samples:
![]() | (1) |
The activation energy for the lithium conductivity is calculated by fitting the data to the Arrhenius equation:
![]() | (2) |
Lithium-ion conductivity of the parent garnets Li5La3M2O12 (M = Nb, Ta, Nb/Ta) can be obtained via Arrhenius plots. Fig. 12 presents the Arrhenius behaviour of Li5La3Nb2O12 sintered at both 900 °C and 1100 °C. The conductivity data, collected during both heating and cooling cycles, follow a linear trend (dashed lines in the plots), indicating that the measurements reflect equilibrium behaviour. The overall trends align well with those reported in literature for similar materials. The Li5La3Nb2O12 sample prepared in 2006 by Thangadurai et al. sintered at 1000 °C lies between the two sintering conditions used in this study, offering a useful point of comparison.38 In general, the Nb- and Ta-only samples show higher conductivities when sintered at 900 °C compared to 1100 °C. Conversely, the mixed Nb/Ta sample displays improved conductivity when sintered at 1100 °C. These results highlight the intricate balance between phase purity and microstructure. While higher sintering temperatures promote densification, which should enhance ionic transport, in this study they also generate impurity phases, which can reside particularly at grain boundaries, increasing resistance and thus reducing overall conductivity. The samples sintered at 900 °C all exhibit comparable conductivities, as expected for samples with similar phase composition and fewer impurities. However, the 1100 °C samples show a broader range of conductivity values, indicating varying responses to high-temperature processing. The highest conductivity at 55 °C was observed for the Li5La3NbTaO12 composition sintered at 1100 °C, with a value of 4.8 × 10−5 S cm−1. Above 300 °C, the conductivities of all samples converge, this high-temperature effect is beyond the scope of the present study, and the analysis focuses on the Arrhenius behaviour between 25 °C and 200 °C.
![]() | ||
| Fig. 12 Arrhenius plots of the total conductivity of all Li5La3Nb2O12 (Nb-LLO); Li5La3Ta2O12 (Ta-LLO); and Li5La3NbTaO12 (NbTa-LLO) sintered at 900 °C (a) and sintered at 1100 °C (b). For comparison, data from literature38 sintered at 1000 °C is included in both (a) and (b). | ||
From the Arrhenius plot, the activation energies for lithium-ion conduction were extracted (Table 2).5,23,38–44 All samples sintered at 900 °C have activation energies of 0.50 eV or lower, reflecting favourable lithium-ion mobility. However, the Ta-substituted sample sintered at 1100 °C exhibits a significantly higher activation energy of 0.62 eV, highlighting a difference in the sample compared to the other samples, and suggesting that a secondary phase, such as surface contamination, may be present. A possible explanation for the reduced conductivity and increased activation energy could be surface contamination with Li2CO3, which can hinder lithium-ion transport and lead to poor overall conductivity. The conductivity study yielded results that are comparable to the literature. However, the observed differences in conductivity across the two sintering temperatures underscore the importance of optimising phase composition and structural features. To fully understand and control the conductivity behaviour in lithium garnet electrolytes, careful consideration of secondary phase (contamination) formation is required.
| Composition (final sintering temperature) | σLi+(S cm−1) | Ea (eV) | References |
|---|---|---|---|
| Li5La3Nb2O12 (900 °C) | 2.9 × 10−5 (55 °C; air) | 0.43 | This study |
| Li5La3Nb2O12 (1100 °C) | 1.9 × 10−5 (55 °C; air) | 0.44 | This study |
| Li5La3Nb2O12 (950 °C) | 2.3 × 10−5 (50 °C; air) | 0.55 | 38 |
| Li5La3Nb2O12 (950 °C) | 8.0 × 10−6 (50 °C; air) | 0.43 | 5 |
| Li5La3Nb2O12 (950 °C) | 2.5 × 10−7(25 °C; air) | 0.51 | 39 |
| Li5La3Nb2O12 (1100 °C) | 5.1 × 10−6 (22 °C; air) | 0.60 | 40 |
| Li5La3NbTaO12 (900 °C) | 6.8 × 10−6 (55 °C; air) | 0.50 | This study |
| Li5La3NbTaO12 (1100 °C) | 2.6 × 10−5 (55 °C; air) | 0.45 | This study |
| Li3.43La3Nb1.07Ta0.93O12−y (1000 °C) | 7.1 × 10−8 (25 °C; air) | 0.59 | 23 |
| Li5La3Ta2O12 (900 °C) | 3.3 × 10−5 (55 °C; air) | 0.42 | This study |
| Li5La3Ta2O12 (1100 °C) | 8.6 × 10−7 (55 °C; air) | 0.62 | This study |
| Li5La3Ta2O12 (850 °C) | 5.0 × 10−6 (40 °C; air) | 0.55 | 41 |
| Li5La3Ta2O12 (950 °C) | 1.2 × 10−6 (50 °C; air) | 0.56 | 5 |
| Li5La3Ta2O12 (1000 °C) | 2.6 × 10−6 (60 °C; air) | 0.6 | 42 |
| Li5La3Bi2O12 (775 °C) | 4.0 × 10−5 (22 °C; Ar) | 0.54 | 43 |
| Li5La3Sb2O12 (950 °C) | 8.2 × 10−6 (24 °C; Ar) | 0.51 | 44 |
In the O 1s spectra, the carbonate oxygen peak at ∼532 eV is significantly reduced after heat treatment, while the lattice oxygen peak associated with the garnet phase at ∼529 eV becomes more prominent. Similarly, in the C 1s region, the carbonate carbon peak at ∼290 eV and the atmospheric carbon peak at ∼285 eV both diminish after heating.47 Initially, XPS measurements were restricted to room temperature due to gas evolution that occurred on heating, which disrupted the vacuum at elevated temperatures. However, partial removal of the carbonate layer via heating enabled higher-temperature measurements by minimising decomposition-related outgassing.
In the sample without additional heat treatment, the XPS spectrum was dominated by Li2CO3, effectively masking signals from key garnet elements such as La and Nb. After heating, the carbonate layer was substantially reduced, revealing signals from the underlying garnet phase (Fig. S11, see SI). Based on the surface sensitivity of XPS and comparison with literature, the thickness of the surface Li2CO3 layer is estimated to be approximately 30 nm–100 nm.15,48,49
Although heating of the sample was effective in immediately reducing the carbonate layer, an order of tens of nanometres of carbonate remained, and the layer is expected to regrow over time due to continued surface reactions with atmospheric H2O and CO2. Since the carbonate layer was observed across all samples during XPS analysis on samples stored in the lab, it is likely that without removing this layer, the ionic conductivity was impacted due to the lithium-ion transport being impeded by the carbonate layer at the pellet surfaces. Fig. S12 (see SI) presents fractured cross-sectional SEM images of an Ar-treated pellet, stored in the dry room for 45 days, and an untreated pellet, stored for more than four months. The images reveal the extent of surface contamination, with a contamination layer of 50–120 nm in the former case and a thicker contamination layer of 500–800 nm in the latter (Fig. S12, see SI).
These findings emphasise the need for future focus on both the complete removal and prevention of surface carbonate formation. It was found that storing samples at room temperature in a desiccator in a dry room alone was insufficient. This area is growing interests, with strategies such as protective coatings, optimised atmospheres during synthesis, or post-synthesis carbonate layer control employed to mitigate excessive Li2CO3 formation and preserve the electrochemical performance of garnet-type solid electrolytes.47,50–52
| Li5La3M2O12 + H2O → “Li5−xHxLa3M2O12·xLiOH” | (3) |
The protonic exchanged garnet subsequently reacts with carbon dioxide and forms a surface layer of lithium carbonate, as proposed in eqn (4).
![]() | (4) |
A TGA experiment was designed to replicate the thermal treatment applied to the XPS samples, by heating the previously synthesised (sintered at 1100 °C) Li5La3M2O12 (M = Nb, Ta, Nb/Ta) pellets to 800 °C for one hour, followed by cooling to 100 °C, with the cycle repeated once. All compositions (Fig. S13, see SI) exhibit three distinct weight loss events occurring at approximately 100 °C, 250 °C, and 550 °C, indicating a similar interaction with atmospheric species across all samples. These features are consistent with the behaviour reported in literature for garnet-type materials:53,54
• The first weight loss (∼100 °C) corresponds to the desorption of surface-adsorbed water.
• The second (∼250 °C) is attributed to the release of structural H2O arising from proton incorporation into the garnet lattice.
• The third event (∼550 °C) reflects the decomposition of surface Li2CO3, with the release of CO2.
The ∼250 °C water loss is attributed to H+ ions substituting for Li+ within the garnet structure, supporting the concept of a Li+/H+ exchange mechanism.53 This mechanism, when coupled with exposure to ambient CO2, results in the preformed LiOH layer reacting to form Li2CO3 on the surface (eqn (4)). Upon heating, the decomposition of the surface carbonate species is believed to occur via loss of carbon dioxide (eqn (5)), and the back-insertion of the lithium from lithium carbonate into the garnet returns to the original “stoichiometry”.
![]() | (5) |
Importantly, the garnet reactions in ambient conditions are reversible, and at the end of the cycle, the original garnet is recovered.55 Interestingly, during both heating cycles, a slight mass increase is observed between 100 and ∼250 °C. This may reflect adsorption or formation of LiOH and Li2CO3 on the surface during heating from residual atmospheric amounts of water and carbon dioxide, which is subsequently lost at higher temperatures. This observation may suggest that carbonate formation is initiated during the cooling phase of synthesis,56 particularly as the sample passes through the 100–250 °C range, where the reaction is thermodynamically favourable (Fig. 14). Additionally, the presence of unreacted lithium, intentionally added during synthesis to compensate for lithium volatilisation at high sintering temperatures, may contribute to the formation of surface Li2CO3. As the amount of excess lithium was not precisely optimised, residual reactive Li2O within the structure may facilitate Li2CO3 formation upon exposure to ambient air. TGA analysis shows that heating the samples to 800 °C for one-hour results in a two-step decomposition of surface Li2CO3. This thermal behaviour aligns with the changes observed in the XPS spectra, confirming the progressive removal of carbonate species from the surface. Understanding the two-step Li2CO3 formation and decomposition process under ambient conditions is essential for guiding future efforts to optimise synthesis protocols, reduce surface contamination, and explore alternative strategies that mitigate carbonate formation more effectively than the current approach.
From the ex situ PXRD patterns of Li5La3Nb2O12, it is evident that the sample sintered at 1100 °C with a sacrificial mother powder and the sample sintered at 900 °C without a mother powder both yield a clean, single-phase garnet structure. In contrast, the sample sintered at 1100 °C without a mother powder shows the garnet phase along with multiple impurity phases, one of which is identified as LaAlO3. Notably, when the sintering time at 1100 °C is extended to 40 hours without a mother powder, the Al-based secondary phase becomes the dominant phase, indicating extensive reaction between the garnet and the alumina crucible. The PXRD analysis of Li5La3Ta2O12 and Li5La3NbTaO12 (Fig. S14 and S15, see SI) sintered without mother powder at 900 °C and 1100 °C, also shows the presence of a major secondary phase in the base pellet sintered at 1100 °C in the alumina crucible.
Al has been widely employed to stabilise cubic phase garnets, both doping intrinsically via alumina crucibles,57–59 or doping extrinsically via the addition of alumina,60–62 where Al substitutes into the structure in place of Li. This substitution creates charge-compensating lithium vacancies, which are essential for stabilising the high-conductivity cubic phase.58 Al2O3 has been employed directly as both a sintering aid and phase stabiliser in the synthesis of garnet-type materials such as Li7La3Zr2O12 (LLZO).62,63 Al2O3 promotes densification by forming a secondary phase at grain boundaries, which has been shown to suppress lithium dendrite growth and enhance overall ionic conductivity.64 The process of intrinsic doping happens at high sintering temperatures, when reactive species, such as Li2O or La2O3 can interact with the crucible, producing transient melt phases capable of dissolving Al2O3.58 This process facilitates the diffusion of Al into the garnet, leading to the formation of secondary Al-containing phases.65 In addition, it has been shown that one significant reason for reaction of garnet with the crucible during high temperature sintering is the Li volatilization inducing Al-doping to ensure charge balance.66
A previous study on LLZO garnets reported that reactions between the garnet and alumina crucibles begin around at ∼1050 °C.67 This aligns with our observations, where Al-containing phase LaAlO3 emerges as a major phase in samples sintered without mother powder at 1100 °C but is only minor in the sample sintered at 900 °C. These results highlight the sensitivity of garnet materials to sintering conditions, particularly the role of temperature in driving Al contamination from the crucible. Careful control over sintering environments is therefore essential to ensure phase purity and material stoichiometry. To minimise Al contamination, two effective strategies are demonstrated in this study: (i) lowering the sintering temperature below 1100 °C, which reduces alumina crucible interaction, though it may compromise pellet densification, and (ii) utilising a sacrificial mother powder, which acts as a protective buffer layer, enabling clean garnet formation even at elevated temperatures. Overall, Al-related secondary phases were only observed in the base pellet in direct contact with the crucible, therefore strongly indicating that Al diffusion and contamination are surface-limited and driven by physical contact during sintering. Using single-crystal sapphire tubes for synthesis during in situ PXRD measurements allowed to avoid the Al contamination of the material due to the considerably lower reactivity of the alumina single crystal surface.
The BSE images of all the analogues sintered at 900 °C and 1100 °C without mother powder for 24 h are presented in Fig. 16. As also observed before from the PXRD analysis, the pellets in contact with the crucible during sintering at 900 °C show no evidence of secondary phases for all garnets. However, sintering at 1100 °C led to the formation of Al-rich phases, most severely in the case of Nb-doped garnets. Line and point EDS analysis of Li5La3Nb2O12 sintered at 1100 °C (Fig. S16 and S17, see SI) reveals the most significant reaction with the alumina crucible. Three distinct phases are observed: the garnet with the targeted stoichiometry, an Al- and La-rich phase, and a Nb-rich phase. This highlights the severe destabilisation of the garnet due to Al contamination from the crucible, forming a LaAlO3 phase, as confirmed by XRD analysis.
Fig. S18 and S19 (see SI), show similar analyses for Li5La3NbTaO12 and Li5La3Ta2O12. For Li5La3NbTaO12, three phases are present: the intended garnet, a Ta- and Nb-rich phase, and an Al-rich phase containing minor La. This indicates notable, though less severe, interaction with the crucible compared to the Nb-based sample. Li5La3Ta2O12 exhibits the least contamination, with only two phases detected: the garnet and an Al- and La-rich secondary phase. These findings demonstrate that Nb-based garnets are more prone to crucible interactions at temperatures above 900 °C, owing to the greater tendency of niobium to undergo reduction at high temperatures.
Combining the findings from PXRD, SEM, and EDS, it is evident that sintering at 1100 °C without sacrificial mother powder leads to the formation of secondary La–Al phases. This effect is intensified with prolonged sintering durations, as the 40-hours sample predominantly shows La–Al phase formation. As established earlier, achieving a single-phase garnet at 1100 °C requires the use of sacrificial mother powder. Beyond phase control, SEM analysis confirms that this approach also promotes better densification and reduces surface voids compared to samples prepared without a mother powder.
Future approaches to achieve highly dense solid electrolytes at elevated temperatures are the use of ZrO2, MgO or Pt crucibles, as they provide higher control on elemental composition of produced garnet,68 especially the latter two show lower reactivity with the Li2O gas.69 For example, it was shown by Jiang et al.70 that a smaller amount of Mg was incorporated in the Ga-doped LLZO structure at the same conditions of heat treatment compared to Al. Alternatively, lower sintering temperatures or lower needed dwelling time at higher temperatures may be employed through advanced techniques such as ultra-fast high-temperature sintering,71 flash sintering,72 cold sintering73 or spark plasma sintering74,75 which can enable densification while minimizing contamination.68,76
The high intensity resonance at ∼11 ppm, characteristic of octahedrally coordinated Al, can be assigned to LaAlO3,40 which is consistent with previous PXRD, SEM, and EDS results. Alongside this dominant peak, all samples exhibit a lower intensity resonance at ∼77 ppm, falling within the expected range for tetrahedrally coordinated aluminates (50–90 ppm).77 A study by Vema et al.62 identified a similar resonance at ∼79 ppm when investigating Al-doped LLZO, attributing it to γ-LiAlO2, therefore, the ∼77 ppm peak observed here could likely correspond to the same phase. Additionally, the Ta-only sample shows a low-intensity resonance at ∼67 ppm. Vema et al. also reported that a resonance at ∼68 ppm arises from Al substitution into the Li 24 d site in the garnet structure, this suggests that the resonance slightly above the noise in the Ta sample may reflect minor Al doping into the garnet lattice.62 Similar observations were reported by Karasulu et al.,10 whose DFT calculations demonstrated that Al preferentially occupies the Li1 site, giving rise to an NMR shift of ∼68.6 ppm and the observed NMR peak broadening originates from local structural distortions of the AlO4 tetrahedra, where the absence of corner-sharing LiO4 (Li2) neighbours increases the distortion index and thereby increasing the quadrupolar coupling constant.
The 27Al MAS NMR spectra were normalised by the sample mass in the rotor, allowing qualitative conclusions about the composition with the most Al-secondary phase to be drawn. The peak at ∼11 ppm, associated with LaAlO3, shows the greatest intensity in the Nb-only sample, followed by the mixed Nb/Ta, and the lowest in the Ta-only sample. This trend suggests that the Nb-only sample contains the highest concentration of LaAlO3 secondary phase, which aligns with the SEM and EDS analysis of the base pellets. Interestingly, the trend does not hold for the low-intensity peaks. The mixed Nb/Ta sample exhibits the most intense resonance at ∼77 ppm, suggesting it contains the most γ-LiAlO2. The solid-state NMR spectroscopy study indicates that aluminium incorporation from the alumina crucible is present across all compositions, predominantly in an octahedral LaAlO3 environment, with the Nb-only sample containing the most secondary phase in the samples studied. This implies that Nb promotes greater reactivity with alumina crucibles. However, as the Li stoichiometry was not precisely determined and conclusions regarding Al incorporation mechanisms remain incomplete. As a study by Lan et al. indicates that Li content strongly influences Al uptake,67 therefore, further analysis of the sample stoichiometry is required to determine whether M-site substitution is driving Al contamination.
d symmetry. SEM imaging showed better densification of the samples sintered at 1100 °C, than those sintered at 900 °C. The in situ PXRD study of the Li5La3Nb2O12 garnet gave an insight into the route of the phase formation. Under the conditions studied, the Nb garnet phase proceeds via La2LiNbO6, Li3NbO4 and La3NbO7, and the product begins to form below 700 °C. The investigation clearly highlighted that at elevated temperatures, above 1075 °C, decomposition of the garnet phase begins, and on the in situ scale, the product was almost entirely decomposed. This study identifies that such elevated temperatures to achieve the garnet structure is not the most efficient and that a lower temperature for a longer dwelling time may be sufficient.
It was observed that samples sintered at 900 °C showed similar conductivities to each other, while those sintered at 1100 °C displayed more variation between the garnet compositions. The highest conductivity measured in this study was 4.8 × 10−5 S cm−1 at 55 °C, for Li5La3NbTaO12 sintered at 1100 °C. This is comparable to Li5 phases previously reported in the literature.35 7Li MAS NMR spectra showed a single resonance for all samples, indicating that all Li environments are equivalent due to fast Li-ion hopping, rendering the Li environments indistinguishable. 27Al MAS NMR spectra revealed Al incorporation from alumina crucibles, in all samples sintered directly on the crucible base at 1100 °C. The main contaminant phase appears to be LaAlO3, which was confirmed by PXRD, SEM, and EDS analysis. This work showed that to reduce Al contamination, sintering with a sacrificial mother powder or at 900 °C is recommended. This study provides a comprehensive analysis of the Li5- garnet phase, a composition often overlooked in favour of the more commonly studied Li7La3Zr2O12 (LLZO). By revisiting the parent garnet structure, this work offers valuable insights that could inform the design and optimisation of next-generation garnet-based solid-state lithium-ion electrolytes.
Supplementary information: X-ray diffraction, scanning electron microscopy and elemental analysis, XPS spectroscopy, TGA analysis and electrochemical impedance data. See DOI: https://doi.org/10.1039/d5ta06905j.
| This journal is © The Royal Society of Chemistry 2025 |