Chih−Long
Tsai
*a,
Ngoc Thanh
Thuy Tran
b,
Roland
Schierholz
a,
Zigeng
Liu
a,
Anna
Windmüller
a,
Che-an
Lin
c,
Qi
Xu
ad,
Xin
Lu
ad,
Shicheng
Yu
a,
Hermann
Tempel
a,
Hans
Kungl
a,
Shih-kang
Lin
bce and
Rüdiger-A.
Eichel
adf
aInstitut für Energie– und Klimaforschung (IEK–9: Grundlagen der Elektrochemie), Forschungszentrum Jülich, D–52425 Jülich, Germany. E-mail: c.tsai@fz-juelich.de
bHierarchical Green-Energy Materials (Hi-GEM) Research Center, National Cheng Kung University, Tainan 70101, Taiwan
cDepartment of Materials Science and Engineering, National Cheng Kung University, Tainan 70101, Taiwan
dInstitut für Materialien und Prozesse für elektrochemische Energiespeicher– und wandler, RWTH Aachen University, D–52074 Aachen, Germany
eProgram on Smart and Sustainable Manufacturing, Academy of Innovative Semiconductor and Sustainable Manufacturing, National Cheng Kung University, Tainan 70101, Taiwan
fInstitut für Energie– und Klimaforschung (IEK–12: Helmholtz–Institute Münster, Ionics in Energy Storage), Forschungszentrum Jülich, D–48149 Münster, Germany
First published on 12th April 2022
Ga-substituted Li7La3Zr2O12 (LLZO) garnet is among the most promising solid electrolytes for next-generation all-solid-state Li battery (SSLB) applications due to its very high Li-ion conductivity. However, the attempts to use Ga-substituted LLZO as a solid electrolyte for SSLBs are not yet successful. Here, the research results show that Li6.4Ga0.2La3Zr2O12 can be reduced by Li at 25 °C when the surface of the material is properly cleaned. The experimental results suggest that Ga leached out of the garnet structure to form the Li–Ga alloy, which apparently would short-circuit the battery if Ga-substituted LLZO is used as a solid electrolyte. When low concentration Ga-substitution is applied, e.g. Li6.45Ga0.05La3Zr1.6Ta0.4O12, the material seems stable against Li at ambient temperature but not at high temperatures, where heat treatment is usually used to reduce the interfacial resistance between Li and LLZO. The experimental results are also supported by density functional theory calculations to show that the Ga-substituted LLZO/Li interface tends to transform into LLZO and the Li2Ga intermetallic compound. The results highlight the importance of substitution selection for LLZO, for which the Ga-substituted LLZO solid electrolyte may not be suitable for direct contact with metallic Li.
In the LLZO-type solid electrolytes, Ga-substituted LLZO represents one of the most studied materials due to its very high Li-ion conductivity at room temperature, >1 mS cm−1, which is much higher than that for other popular substitutions such as Al or Ta.1 The fundamental mechanism leading to this higher conductivity is not yet clear. Specifically, nuclear magnetic resonance (NMR) spectroscopy has been used to study the local environment of Ga to understand its site preference. Rettenwander et al. applied ultrahigh magnetic field NMR spectroscopy for studying Ga and Al co-substituted LLZO suggesting that Ga and Al have similarity in site preference in tetrahedral 24d and octahedral 96h sites.12 It was hypothesized before that Ga is solely located on the 96h site irrespective of the amount of Ga-substitution, which would not interfere with the diffusion of Li-ions as opposite to the Al-substitution that occupied the 24d site to act as a blockade for Li-ion conduction.13 The results weaken the previous assumption that substituent's site preference is responsible for the higher Li-ion conductivity of Ga- than Al-substituted LLZO. Furthermore, Karasulu et al. performed solid-state 17O, 27Al, and 71Ga magic angle spinning (MAS) NMR spectroscopy combined with density functional theory (DFT) calculations to suggest that the sharp and broad features in 27Al and 71Ga NMR spectra are from the tetrahedral 24d site with varying local distortions, whereas the octahedral 96h/48g sites are energetically inaccessible and not experimentally observed.14 This result is also in agreement with that of Bernuy-Lopez et al. from their 71Ga MAS NMR study.15
Neutron powder diffraction (NPD) was also used for understanding the site occupation for Ga-substitution in LLZO by taking advantage of knowing that the scattering length of Ga is positive while that of Li is negative. The refinements of the NPD results are in agreement with Ga-ions being located on the tetrahedral 24d site while Li-ions are occupying both tetrahedral 24d and octahedral 96h sites, with more Li-ions at the 96h site than that at the 24d site.16–18 Combining the 6Li MAS NMR results, which suggests that the Li-ions at the octahedral 96h site have a higher mobility than those at the tetrahedral 24d site, and NDP results, Wu et al. proposed that the strong coulombic repulsion between immobile Ga-ions and the nearby Li-ions shifts Li-ions away from their central position within the octahedral 96h site in the garnet structure. Hence, the higher Li-ion conductivity of Ga-substituted LLZO can be attributed to the enhanced mobility of Li-ions that increased from the coulombic repulsion between Ga-ions and Li-ions.17
Even though LLZO is generally regarded as electrochemically stable when in contact with Li, recent research studies have discovered that the stability of LLZO toward Li is highly dependent on its substitutions. Theoretically, DFT calculations are used to determine the thermodynamic electrochemical stability window of LLZO.19–21 The results show that the reduction decomposition of LLZO only occurred at a voltage below 0.05 V vs. Li/Li+, where LLZO is reduced into Li2O, Zr3O and La2O3. When reducing the potential down to 0.004 V vs. Li/Li+, Zr3O can be further reduced into Zr metal. Furthermore, DFT calculations also suggest that small amounts of substitutions such as Al, Ta and Nb do not have large effect on the reduction potential of LLZO. Therefore, the good stability of LLZO against Li should be a result of formation of the surface passivation layer by Li2O, La2O and other oxides. Experimentally, Zhu et al. used X-ray photoelectron spectroscopy (XPS) and electrochemical impedance spectroscopy (EIS) to show the interfacial reactivity between LLZO and Li.22 Their XPS results show that Zr4+ can be reduced by Li into Zr0 or Zr2+ on the surface of Al-, Ta- and Nb-substituted LLZO. The formation of an “oxygen-deficient interphase” that protects Al- and Ta-substituted LLZO from further reduction into their bulk, and, therefore, stabilized the systems. In contrast, the Nb-ions in Nb-substituted LLZO have a thermodynamic preference to segregate to the interface between Nb-substituted LLZO and Li. The presence of the enhanced Nb content at the interface combined with significant reduction of Nb-ions by Li enables the reduction reaction propagation into the bulk Nb-substituted LLZO. Similarly, Rettenwander et al. revealed that Li-deficient tetragonal LLZO is formed when Fe-substituted LLZO is in contact with Li by using Raman spectroscopy.23 Black surface coloration at the interface between Fe-substituted LLZO and Li can be observed due to the reduction of Fe3+ into Fe2+ in Fe-substituted LLZO.
Thin film approaches have been proposed for LLZO-type solid electrolytes in SSLB fabrications for achieving a better energy density due to the use of heavy La and Zr elements, that gives a relatively high theoretical density of ∼5.3 g cm−3 for LLZO,24 when compared to sulfide-, e.g. Li3PS4: 1.83 g cm−3,25 phosphate-, e.g. LiPON: 2.30 g cm−3 (ref. 26) and halide-, e.g. Li3InCl6: 2.69 g cm−3,27 types of solid electrolytes. However, the thin film fabrications are not as straightforward as what was expected due to Li evaporation and reaction with the used substrates at high processing temperatures as well as inability to completely crystallize the deposited precursor into the garnet structure, which leads to the reported LLZO thin films usually having Li-ion conductivities at least one order lower than that of the bulk LLZO.3,4,7 Nevertheless, Romanyuk et al. were able to deposit a Ga-substituted LLZO thin film with a thickness of ∼1 μm and a record high ionic conductivity of 1.9 × 10−4 S cm−1 at room temperature.6 Unfortunately, demonstration of the Ga-substituted LLZO thin film as a solid electrolyte for thin film SSLBs without any assistance of a liquid electrolyte is still not yet available.5 Although the studies of Ga-substituted LLZO had been quite popular, the report of electrochemical stability of Ga-substituted LLZO toward Li is lacking, especially the real electrochemical stability of Ga-substituted LLZO toward Li is often hidden by its surface impurities, such as Li2CO3, due to Li-proton exchange.1
Here, the electrochemical stability of Ga-substituted LLZO toward Li has been studied. The results show that the electrochemical stability of Ga-substituted LLZO toward Li is much lower than what is expected due to the leaching out of Ga into the grain boundary when in contact with Li. Especially, a low substitution of 0.05 mole Ga to the Li site also de-stabilizes LLZO at high temperatures when compared to that by using 0.05 mole Al substituent. The conclusion from the experimental results is also supported by density functional theory calculations. The results point out the importance of substituent selection when the LLZO thin film is expected to be used as a solid electrolyte for SSLB developments.
Fig. 1(a) shows photos of the reduction reaction of Li6.4Ga0.2La3Zr2O12 by using Li after placing it onto a 250 °C heating plate. It can be clearly seen that the colour of Li6.4Ga0.2La3Zr2O12 turned from ivory to black from the interface of Li/Li6.4Ga0.2La3Zr2O12 and propagated through the whole 1.5 cm thick pellet, also see movie S1.†
The Li stripe on top of Li6.4Ga0.2La3Zr2O12 indicates that the reduction reaction starts at a temperature lower than the melting temperature of Li, i.e. 180.5 °C. In contrast, Li6.45Al0.05La3Zr1.6Ta0.4O12 did not show any visible change of colour or reaction throughout the experiment. For Li6.45Ga0.05La3Zr1.6Ta0.4O12, the reduction reaction is much slower than that of Li6.4Ga0.2La3Zr2O12. Fig. 1(b) shows photos of the results from heating Li/LLZO/Li symmetric cells of Li6.45Al0.05La3Zr1.6Ta0.4O12 and Li6.45Ga0.05La3Zr1.6Ta0.4O12 at 250 °C for 45 minutes. Both the pellets have a thickness of 0.6 mm. After removing the Li electrodes inside a glove box, the Li6.45Ga0.05La3Zr1.6Ta0.4O12 pellet was also found in black colour while a small part of the material at the central part of the pellet remained ivory in color. This indicates that the reduction reaction of Li6.45Ga0.05La3Zr1.6Ta0.4O12 is much slower than that of Li6.4Ga0.2La3Zr2O12 but still not stable towards Li. On the other hand, Li6.45Al0.05La3Zr1.6Ta0.4O12 showed only a thin layer of coloration, which can be explained by the reduction of Zr4+ to Zr2+ to form a stable oxygen-deficient interphase layer to block the continuous reduction reaction within Li6.45Al0.05La3Zr1.6Ta0.4O12.22
The electrochemical stabilities of Li6.45Ga0.05La3Zr1.6Ta0.4O12 and Li6.4Ga0.2La3Zr2O12 toward Li were also examined at 50 °C and 25 °C, respectively, by using EIS for monitoring the reactions. The surfaces of the LLZO pellets were polished in a glove box to remove any possible Li2CO3/LiOH layer from Li-proton exchange before Li electrodes were attached to the pellets. Isostatic pressing with a pressure of 500 MPa and 30 second dwell time was used to reinforce the attachment of Li electrodes onto LLZO pellets after Li electrodes were attached onto LLZO by hand. However, short circuiting of Li6.4Ga0.2La3Zr2O12 symmetric cells was always observed right after pressing. After removal of Li electrodes, black surface coloration can be seen in some parts of the Li6.4Ga0.2La3Zr2O12 pellets, as shown in Fig. S2.†
The short circuiting of the Li6.4Ga0.2La3Zr2O12 symmetric cells shows the importance of proper surface cleaning to see the reduction reaction of Li6.4Ga0.2La3Zr2O12 from Li. Nevertheless, Li electrodes were hand pressed onto Li6.4Ga0.2La3Zr2O12 pellets for the experiment. Fig. 1(c) shows the long term stability test of Li6.45Ga0.05La3Zr1.6Ta0.4O12 against Li at 50 °C for 60 days. The as prepared symmetric cell has a bulk resistance of ∼65 Ω and interface resistance of ∼32 Ω at 50 °C. With the increase of time at 50 °C, the interface resistance reduced from ∼32 Ω to only ∼4.7 Ω while the bulk resistance remained the same at ∼65 Ω. The decrease of the interface resistance can be explained by the improvement of the contact between Li and Li6.45Ga0.05La3Zr1.6Ta0.4O12 due to Li deposition/dissolution during the EIS measurements and the possible surface reaction between Li and Li6.45Ga0.05La3Zr1.6Ta0.4O12. It could be understood that an electric current of ∼150 μA cm−2 or higher was sent between the Li electrodes when a 20 mV AC bias was used in the EIS measurements. Li can be deposited/stripped onto/from the interface for improving the contact resistance especially when the measuring frequency was 1 Hz. Since the bulk resistance and the colour of the surface remained the same after the measurements, the surface reaction between Li and Li6.45Ga0.05La3Zr1.6Ta0.4O12 is unlikely to happen at 50 °C but cannot be eliminated completely at the nanoscale.
Fig. 1(d) shows the EIS measurement of the prepared Li/Li6.4Ga0.2La3Zr2O12 /Li symmetric cell at 25 °C by using 1 mm thick Li6.4Ga0.2La3Zr2O12 pellet. The EIS spectrum of the as prepared cell can be fit by using a circuit consisting of one resistor in series with three other R-CPE circuits. It is important to mention again that the separation of the bulk and grain boundary contributions was challenging even when ultra-high frequency EIS and ultra-low temperatures were used for the measurement, as shown in Fig. S1.† Hence, the assignment of one of the R-CPE circuits with C = 2.02 × 10−9 F solely to the contribution from the grain boundary would not be convincing. Therefore, the contributions from three different R-CPE circuits were assigned to the bulk/grain boundary of LLZO (Cbulk/GB = 2.02 × 10−9 F), interface resistance (Cint = 1.79 × 10−7 F) and charge transfer (Cct = 4.22 × 10−2 F), where the total resistance of the Li6.4Ga0.2La3Zr2O12 solid electrolyte was ∼160 Ω, i.e. the resistance (Z′) from x = 0 to the intercept at low frequency for the Rbulk/GB–Cbulk/GB circuit, as shown in Fig. 1(e). The interface resistance was ∼348 Ω, which is much higher than that of the Li6.45Ga0.05La3Zr1.6Ta0.4O12 cell (∼32 Ω) because the attachment of Li electrodes was only done by hand. With the increase of time, all the three contributions that were recognized from fitting were dramatically reduced, as shown in Fig. 1(d). After 32 hours, the contributions from the interface and charge transfer were no more recognizable, while the resistance from the Li6.4Ga0.2La3Zr2O12 pellet was reduced to ∼15 Ω only. If we take the real part of EIS spectra at 7 MHz as the bulk resistance contributions from the Li6.4Ga0.2La3Zr2O12 pellet solely and 1 Hz as the total cell resistance, which includes the bulk resistance and interface resistances of the cell, it can be seen that the cell resistance decreased much faster than that for the bulk Li6.4Ga0.2La3Zr2O12 pellet solely in the first 32 hours. This indicates that the reduction reaction of the Li6.4Ga0.2La3Zr2O12 pellet by Li electrodes starts from the surfaces into the bulk and finally stabilized to ∼5 Ω for the whole cell, Fig. 1(f). After the measurements, Li6.4Ga0.2La3Zr2O12 was black in colour throughout the pellet with some ivory colour single crystals decorated on the pellet, as shown in the inset picture in Fig. 1(d). The small cell resistance and the colour change of the pellet both indicate that the reduction reaction of Li6.4Ga0.2La3Zr2O12 by Li can happen at room temperature.
![]() | ||
Fig. 3 Microstructure and EDS mapping images of pristine Li6.4Ga0.2La3Zr2O12, (a)–(e), and Li6.45Ga0.05La3Zr1.6Ta0.4O12, (f)–(l). |
7Li and 71Ga MAS NMR spectra were used for characterizing the pristine and reduced Li6.4Ga0.2La3Zr2O12. The samples were collected from the central part of the 1.5 cm thick pellet to avoid any possible residual Li electrode for the MAS NMR measurements. The 7Li MAS NMR spectrum of pristine Li6.4Ga0.2La3Zr2O12, Fig. 2(e), shows one resonance at 1.2 ppm, which can be assigned to the Li-ions in the two Li sites with a fast exchanging rate in Li6.4Ga0.2La3Zr2O12. Apart from the 7Li resonance of Li6.4Ga0.2La3Zr2O12, 3 new resonances were observed in the 7Li MAS NMR spectrum of the reduced Li6.4Ga0.2La3Zr2O12 polycrystalline sample. The far left one at 263.2 ppm is assigned to the Li metal signal and the two middle resonances at 67.2 and 45.8 ppm, respectively, are tentatively assigned to the Li–Ga alloy. The result indicates that Li metal was infiltrated into the pellet and reacted with Ga to form the alloy. Furthermore, 2D exchange spectroscopy (EXSY) experiments were performed for the reduced Li6.4Ga0.2La3Zr2O12 sample. Only diagonal peaks were observed for Li metal and Li6.4Ga0.2La3Zr2O12 resonances, while off-diagonal peaks were obtained for the Li–Ga alloy resonances with the mixing time between 1 and 30 ms, as shown in Fig. S4,† implying that the two alloy signals are exchanged at the millisecond timescale. The exchange rate determined by fitting the cross peak build up curve is 8.9 s−1 at 295 K. Thus, the 7Li MAS NMR spectrum suggests that the Li–Ga alloy is likely the pathway for Li metal to penetrate into Li6.4Ga0.2La3Zr2O12. For the reduced single crystal samples, collected by cleaning out the black color material using absolute ethanol, as shown in Fig. S2(b),† only the Li-ion signal of Li6.4Ga0.2La3Zr2O12 can be detected in the 7Li MAS NMR spectrum, suggesting that the infiltration of metallic Li occurs only along the grain boundaries instead of in closed porosities within the crystals. 71Ga MAS NMR spectra of pristine and reduced Li6.4Ga0.2La3Zr2O12 samples, as shown in Fig. 2(f), show two resonances: one sharp resonance at ∼238 ppm is assigned to Ga-ions at the ordered tetragonal 96h site and the broad resonance between 50 ppm and 200 ppm is assigned to Ga-ions at the distorted tetragonal 24d site within Li6.4Ga0.2La3Zr2O12.12–15,37 When compared with the pristine sample, the reduced Li6.4Ga0.2La3Zr2O12 shows relatively less Ga-ions in the ordered tetragonal 96h site, implying that some of the Ga was leached out of the garnet structure during the reduction process. It has to be emphasized again that the chemical shift from the 71Ga MAS NMR spectrum directly points to the Ga-ions in the garnet structured Li6.4Ga0.2La3Zr2O12 without influence of the impurity phase at grain boundaries.
Fig. 4 shows the microstructures of reduced Li6.4Ga0.2La3Zr2O12 and Li6.45Ga0.05La3Zr1.6Ta0.4O12. The material with dark contrast in the BSE micrograph, i.e. light molar mass elements, fills up the grain boundaries and covers most of the grains of reduced Li6.4Ga0.2La3Zr2O12, as shown in Fig. 4(a). It is important to note that new cracks were formed and cut through the grains of Li6.4Ga0.2La3Zr2O12. These newly developed cracks were also filled up with the material of dark contrast, which suggests that the gains of Li6.4Ga0.2La3Zr2O12 were opened up by the formation of this dark contrast material. EDS mapping images were also used for identifying the elemental distributions of reduced Li6.4Ga0.2La3Zr2O12, as shown in Fig. 4(b)–(g). Nitrogen was specifically found on the gain boundary where the dark contrast material is located, as shown in Fig. 4(c), indicating that the dark contrast material is Li3N that was formed during transferring the reduced sample into the SEM chamber. More importantly, areas with high Ga concentration were not only found at the grain boundaries but also in the closed pores of Li6.4Ga0.2La3Zr2O12 grains where the oxygen concentration was detected to be minimum. The observation indicates that Ga was leached out of the garnet structure as metal during the reduction process as suggested by the NMR result. Since the alloying of Ga and Li is accompanied by volume expansion, it could be understood that the newly formed transgranular cracks could be a result of the Li–Ga alloy formation, which also provides the pathways for Li to infiltrate into the Li6.4Ga0.2La3Zr2O12 pellet to keep the reduction reaction propagating through the whole ceramic pellet as suggested by the 2D EXSY result. Furthermore, the reduction reaction of Li6.4Ga0.2La3Zr2O12 accompanied with Li metal expulsion was also observed during sample preparation for SEM by using Ar-ion milling under vacuum, as shown in Fig. 4(h) and (i). It was reported that the electron beam expulsed Li out of Li6.4La3Zr1.4Ta0.6O12 during the SEM investigation due to electron injection into the bulk LLZO.47 However, the positive charge of the Ar-ion beam and inert nature of Ar make the reason for Li metal expulsion unclear. From the change of Li6.4Ga0.2La3Zr2O12 colour after Ar-ion milling, it could be speculated that the valence of some elements in Li6.4Ga0.2La3Zr2O12 changed to a lower state while Li-ions need to be expelled as Li metal to keep the system neutral.
Different from Li6.4Ga0.2La3Zr2O12, the microstructure of reduced Li6.45Ga0.05La3Zr1.6Ta0.4O12 did not have any noticeable change from that of the pristine one. No newly developed transgranular cracks can be found within the investigated sample cross-section. The grain boundaries of reduced Li6.45Ga0.05La3Zr1.6Ta0.4O12 are tightly binding the grains to hold the whole structure of the sintered pellet even after the pellet was immersed into water, as shown in Fig. S2(c).† EDS mapping images of the reduced Li6.45Ga0.05La3Zr1.6Ta0.4O12 pellet show similar results as that of Li6.4Ga0.2La3Zr2O12, as shown in Fig. 4(k)–(q). Areas with higher oxygen concentration can be found, which are overlapped with higher nitrogen concentration areas. The high oxygen and nitrogen concentration also indicate that Li was infiltrated into the sample along the grain boundaries, which is in agreement with the Ar-ion polished sample that Li3N was grown out of the reduced Li6.45Ga0.05La3Zr1.6Ta0.4O12 from the grain boundaries after exposing to air, as shown in Fig. 4(r). Unlike the reduced Li6.4Ga0.2La3Zr2O12 sample, high Ga concentration was not found in the close pores within the large grains and the areas where Li was infiltrated are not always superposed with LiGaO2. The result suggests that the infiltration paths for Li may not necessarily always go along with where LiGaO2 is.
Er (x) = Ef [competing phases] − xEf[Ga–LLZO] − (1 − x)Ef[Li] |
![]() | ||
Fig. 5 The convex hulls of (a) Ga0.25LLZO/Li and (b) Ga0.125LLZO/Li pseudo-binary systems, in which the most stable phase constituents (those with the lowest reaction energies) at various interface compositions marked as A–D and A′–C′ are listed. The detailed interfacial reaction equations can be found in Tables S2 and S3.† |
Based on the assumption of local equilibrium at the interface,48 the tie lines on the convex hull of Ga0.25LLZO/Li connect the stable sets of competing phases that are marked as points (A, B, C and D), while those above the hull are metastable at the interface of Ga0.25LLZO/Li. Obviously, the Ga0.25LLZO/Li interface is unstable and interfacial reactions may occur. A number of Li–Ga binary compounds may be formed at the interface, namely Li2Ga, LiGa, and Li2Ga7, indicating the driving force of Ga segregation and diffusion from Ga0.25LLZO to the interface to react with Li. On the other hand, the interfacial reaction also drives Li atoms to diffuse from the bulk Li to the interface and transfer into Li-ions to intercalate into the garnet structure to form tetragonal phase LLZO. Among the stable interfacial phase constituents, the most energetically favourable set of competing phases in the convex hull (point A) suggests that the Ga0.25LLZO/Li interface tends to transform into Li7La3Zr2O12 and the Li2Ga intermetallic compound. The formation of Li2Ga intermetallic compound also gives the reason that the reduction reaction can propagate through the whole pellet and allows the infiltration of metallic Li into the bulk. A similar result was obtained in the case of lower Ga-substitution (Ga0.125LLZO) as shown in Fig. 5(b); however, the reaction energies are less negative. This indicates that despite being an unstable interface, its interfacial stability against metallic Li is better than the higher Ga-concentration substitution one, which is in agreement with the observation that Li6.45Ga0.05La3Zr1.6Ta0.4O12 is more stable toward Li than Li6.4Ga0.2La3Zr2O12. It is worth mentioning that both Ga0.25LLZO and Ga0.125LLZO themselves are not in the ground state and could decompose when presuming that Li7La3Zr2O12 is a stable phase against Li for the calculations. The negative formation energy of Ga0.25LLZO and Ga0.125LLZO may also explain the experimental observation of LiGaO2 formation on the grain boundaries. The calculated results agree closely with experimental observation on the instability of Ga-substituted LLZO against Li metal, as shown in Fig. 1 and 3.
Here, the stability of Ga-substituted LLZO toward Li has been studied. The results show that Ga-substituted LLZO has relatively low stability when in contact with Li, even with a very low concentration substitution of Ga, i.e. Li6.45Ga0.05La3Zr1.6Ta0.4O12. It could dramatically reduce the stability of the whole system when compared to that of the Al-substituted one, i.e. Li6.45Al0.05La3Zr1.6Ta0.4O12. The reduction reaction was also found to be accompanied with the infiltration of Li into the bulk pellet along the grain boundaries. Since the reduction reaction was done without applying any external field, the infiltration of Li into the bulk pellets must be driven by the alloying effect with another element to keep Li in the metal form. From the EDS results, one possibility to create this pathway would be the reaction of LiGaO2 at the grain boundaries with Li to form the Li–Ga alloy and Li2O, where the Li–Ga alloy can keep the reaction continuing through the bulk pellet, also see ESI, Fig. S5 and movie S2† for more discussion. However, the reaction between LiGaO2 and Li is not able to explain the results from 71Ga MAS NMR spectroscopy that suggest that the Ga concentration in Li6.4Ga0.2La3Zr2O12 was less than that of the pristine one and the XRD results suggest that the phase of Ga-substituted LLZO was changed from cubic to the tetragonal phase after the reduction reaction. In combination with the results from NMR spectroscopy, XRD and EDS mapping, especially the observation of high Ga concentration in the close pores of reduced Li6.4Ga0.2La3Zr2O12 grains, it is reasonable to conclude that Ga leaches out of Ga-substituted LLZO to form the Li–Ga alloy while some Li atoms must oxidize to form Li-ions to fill into the Li-vacancies in LLZO to keep the material neutral when Ga-substituted LLZO is in contact with Li and considering that trivalent diffusion is possible within the oxide framework.49,50 The observed results are in good agreement with DFT calculations on the Ga-LLZO/Li interfacial stability. The leached out Ga further forms alloy with Li to keep the reduction reaction propagate through the bulk pellet by keeping its electrochemical potential the same as that of the Li electrode on the surface of the LLZO pellet due to the high electronic conductivity of the Li–Ga alloy. With the increase of the Li concentration in the garnet structure, a phase transformation from the cubic to tetragonal phase garnet structure can be expected when the Li-ion concentration in the garnet structure is higher than 6.6.45 Nevertheless, if only considering the short circuiting of the pellet, the contribution from the reaction of LiGaO2 and Li at the grain boundaries to form the Li–Ga alloy cannot be totally excluded. A similar leaching out behaviour of Al to LiCoO2 to form a highly resistant tetragonal phase from cubic phase Al-substituted LLZO was reported by Park et al.10 The transformation behaviour from cubic phase Fe-substituted LLZO into tetragonal phase Fe-substituted LLZO due to the reduction of Fe3+ to Fe2+ when in contact with Li was also reported by Rettenwander et al.23 Similarly, the observation of the reduction reaction of LLZO from Li to form a black surface coloration layer was reported for Fe-, Nb- and (Nb, Sc)-substituted LLZO.22,23,51
For Li6.45Ga0.05La3Zr1.6Ta0.4O12, if we assume that all Ga was leached out to form the Li–Ga alloy and Li-ions were taken up from Li metal to keep the material neutral, the chemical formula of the reduced material would be Li6.6La3Zr1.6Ta0.4O12, which should have a cubic phase as its major phase.45 However, XRD analysis suggests that the reduced Li6.45Ga0.05La3Zr1.6Ta0.4O12 consists of a major phase in the tetragonal form. This indicates that at least one other element than Ga in Li6.45Ga0.05La3Zr1.6Ta0.4O12 would need to reduce its valence so that more Li-ions can be taken up. On the other hand, Zhu et al. proposed that the formation of the oxygen-deficient interphase due to the reduction of Zr4+ and Ta5+ in Ta-substituted LLZO by Li could stabilize the interface between Ta-substituted LLZO and Li to prevent Ta-substituted LLZO to reduce into the bulk.22 The observation of Li6.45Ga0.05La3Zr1.6Ta0.4O12 being reduced into the bulk pellet also suggests that Ga in Li6.45Ga0.05La3Zr1.6Ta0.4O12 or Li6.4Ga0.2La3Zr2O12 could possibly behave as Nb in Nb-substituted LLZO which segregates to the LLZO surface to destabilize the formation of the oxygen-deficient interphase, even with a relatively low amount of Ga-substitution in LLZO.
Powder X-ray diffraction (XRD) was carried out by using an EMPYREAN (Panalytical) with Ni-filtered Cu Kα radiation. Samples for XRD were collected from the sintered and reduced pellets by grinding and loading the ground powders into front loading sample holders. The diffractogramms were collected in the 2θ range from 10° to 80° at 40 kV, 40 mA, with a step size of 0.008° 2θ and counting time of 20 seconds per step. Qualitative phase analysis was carried out with the QualX2.0 software package52 employing the crystallography open database.53–58 The phase compositions of the patterns were further analyzed within the GSAS II software package59 by using a combined Le Bail/Rietveld approach.60–62 Solid state nuclear magnetic resonance (NMR) spectra were acquired on a Bruker Advance NEO 800 MHz spectrometer using a 1.3 mm MAS probe spinning at 60 kHz with the Hahnecho pulse sequence. The 90° pulse length for 7Li and 71Ga was 1.6 and 1.5 μs, respectively. The recycle delay was 10 and 5 s, respectively. The 2D 7Li exchange spectroscopy (EXSY) experiment was carried out with the mixing time in the range of 1 to 300 ms. 7Li and 71Ga shifts were referenced to LiF (−1 ppm) and Ga2(SO4)3 (−87 ppm) powders. Microstructures and elemental chemistry distribution were investigated by using a Quanta 650 FEG scanning electron microscope (SEM) (FEI Company, USA) equipped with an energy dispersive spectrometer (EDS) (Ametek, USA).
EIS was used for determining the ionic conductivity of fabricated samples by using the Turnkey Broadband System Concept 80 (Novocontrol, Germany). The EIS measurements were done by using two different frequency generation systems. One used Agilent E4991B, which measured in the frequency range of 3 GHz ∼ 10 MHz, and the other used the Alpha-A analyzer, which measured in the frequency range of 10MHz ∼ 1 Hz, in the Broadband System Concept 80 system. Sputtered gold thin films were used as electrodes for ionic conductivity measurements. The impedance spectra were collected in the temperature range from −60 °C to 60 °C with a step size of 10 °C. For the stability studies of Li symmetric cells, a BioLogic VMP-300 multipotentiostat, combined with a climate chamber (Vötsch 4002 EMC), was used. The EIS data were analyzed by “ZView” (Scribner Associates, USA).
Density functional theory (DFT) calculations for Ga-substituted cubic LLZO and all possible competing compounds in Li–La–Zr–O–Ga are done by using the Vienna Ab initio Simulation Package (VASP).63 The electron–ion interactions are evaluated by the projector augmented wave (PAW) method,64 whereas the electron–electron interactions are calculated under exchange–correlation functional within the generalized gradient approximation of the Perdew–Burke–Ernzerhof method.65 All atomic coordinates are relaxed until the total energy difference is smaller than 10−5 eV and the Hellmann–Feynman force on each atom is less than 0.01 eV Å−1. The wave functions are built from the plane waves with a maximum energy cutoff of 520 eV, which is tested to achieve accuracy. The k-point mesh is set up as 2 × 2 × 2 for the 2 × 2 × 2 Ga-substituted LLZO supercell in the Monkhorst–Pack scheme and these k-point values might differ for other compounds depending on their cell sizes for converged formation energies. In this work, Li6.25Ga0.25La3Zr2O12 (Ga0.25LLZO) and Li6.625Ga0.125La3Zr2O12 (Ga0.125LLZO) models are built for the rich and dilute Ga substitutions to match with our designed 2 × 2 × 2 cubic LLZO supercell. The reaction energies and equilibrated competing phases are determined by applying the self-developed Matlab code on all possible compounds of the Li–La–Zr–O–Ga system (ESI, Table S1†).
Footnote |
† Electronic supplementary information (ESI) available. See https://doi.org/10.1039/d1ta10215j |
This journal is © The Royal Society of Chemistry 2022 |