Ion transport scenarios under various vacancy and Na-ion concentration
The crystal structures of superionic conductors critically determine their ion transport capabilities and migration bottleneck sizes. In crystalline inorganic SEs, ionic hopping predominantly occurs between adjacent sites with the lowest energy barriers, whether through single-ion or multi-ion concerted migration30. Furthermore, charge carriers typically conduct through the grain bulk and along boundaries via interconnected diffusion channels formed by the anion framework. Broadly, the conductivity of inorganic SEs can be succinctly expressed by Eq. (1):
σ = qnµ (1)
Where q, n, and µ are indexed to the elementary charge, charge carrier concentrations, and the mobility of carriers, respectively. Notably, carrier mobility is influenced by various factors, such as defect concentration, interatomic forces, disorder, and channel size3. In halogen anion frameworks, ion conduction mechanisms are shaped by vacancy and mobile charge carrier concentrations31, which typically exhibit a negative correlation. Excess charge carriers hinder ionic hopping, while too many vacancies increase migration distances. Thus, theoretical concentration-balanced thresholds for carriers and vacancies likely exist, with SEs achieving maximum conductivity when these thresholds are met. As illustrated in Fig. 1a, we disentangled the impact of carrier and vacant-site contents on Na-ion diffusion through three designed scenarios. In scenario (Ⅰ), where Na-ion content > threshold (\(\:{\text{T}}_{{\text{Na}}^{\text{+}}}\)) and vacancy content < threshold (Tvacancy) in chloride frameworks, Na-ion hopping is hindered due to the limited number of available vacancies. Thus, the ion conduction process is dominated by vacancies. On the contrary, scenario (Ⅱ) shows that when Na-ion content < \(\:{\text{T}}_{{\text{Na}}^{\text{+}}}\) and vacancy content > Tvacancy, ion hopping is limited by the reduced mobility of Na-ion charge carriers. Consequently, the ion conduction process becomes ion-dominated. Scenario (Ⅲ) demonstrates that when both Na-ion and vacancy content are at their respective thresholds, Na-ion conduction is most efficient, indicating optimal ionic transport kinetics. Notably, a strong correlation exists between the concentration thresholds of vacancies and charge carriers in the structure. As vacancy content nears its threshold, charge carrier concentrations also approach their theoretical limits. Conversely, when these concentrations deviate from their thresholds, the variation trends of vacancies and charge carriers become diametrically opposite.
Meanwhile, the experimental results strongly support these viewpoints. As illustrated in Fig. 1b, optimizing vacancy and charge carrier contents using a NaCl-poor method yields compositions such as Na0.5ZrCl4.5, Na0.6Er0.4Zr0.6Cl4.2, and Na0.4Yb0.25Zr0.75Cl4.15, delivering higher conductivities of 8.1 × 10− 5 S cm− 1, 5.7 × 10− 5 S cm− 1, and 2.0 × 10− 4 S cm− 1 than their original samples at 25 ℃, respectively. As a proof-of-concept study, the representative trigonal phase structure (space group: P—3m1) NaxZrCl4+x SEs were selected to research their structure changes induced by the NaCl-poor process. Then, the electrochemical behaviors of ASSIBs based above electrolyte systems will be systematically evaluated.
Structural analysis, micro-morphology, and ion conduction
Figure 2a shows the ternary Na-Zr-Cl composition diagram, highlighting the stoichiometric ratios explored in this study (NaxZrCl4+x, where 0.2 ≤ x ≤ 2) with red markers. The NaxZrCl4+x SEs were synthesized using a mechanochemical ball-milling method with varying molar ratios of NaCl and ZrCl4. The powder X-ray diffraction (XRD) patterns (Fig. 2b) revealed that the crystal structures of the NaxZrCl4+x series correspond to typical trigonal phases, isostructural with Li2ZrCl6 and Li3YCl616,32. Furthermore, as the NaCl molar ratio decreases, the XRD diffraction peaks of the samples show reduced intensity and a slight rightward shift. This suggests partial volume contraction in the crystal structure and lower product crystallinity. The weak crystallinity of the NaxZrCl4+x samples primarily resulting from reduced grain size and strain-broadening effects induced by the intense mechanochemical ball-milling. During the synthesis process, the electrolyte particles would be crushed and creating structural disorder and defects, which contributed to promoting ion hopping33,34. Figures 2c-d and Figures S1-2 illustrate the XRD Rietveld refinements and corresponding crystal structures of representative Na0.5ZrCl4.5 and Na2ZrCl6 SEs. Detailed structural parameters are provided in Tables S1-2. The crystallographic atom sites within Na2ZrCl6 and Na0.5ZrCl4.5 SEs are identical, and both compounds share the same space group, P—3m1. This similarity indicates that the reduction in NaCl composition does not alter the electrolyte's microstructure, facilitating the examination of how vacant-site/Na-ion ratios affect the conductivities of HNISEs. Furthermore, Raman spectroscopy (Figure S3) reveals similar spectral profiles for Na2ZrCl6 and Na0.5ZrCl4.5 SEs, further supporting that the NaCl-poor process has minimal impact on the electrolyte structures. Distinctly, Na0.5ZrCl4.5 SEs exhibit more vacancies (Na and Cl vacant sites) and smaller lattice sizes compared to Na2ZrCl6 due to the NaCl-poor effect. The micro-morphologies of Na2ZrCl6, Na0.5ZrCl4.5, and Na3PS4 SEs are compared by scanning electron microscopy (SEM) images (Figure S4). The Na0.5ZrCl4.5 products present smaller particle sizes and lower porosity (under 300 MPa) than the others, indicating reduced crystallinity and improved deformability in NaCl-poor SEs.
Figure 2e compares the Arrhenius plots of the NaxZrCl4+x (0.2 ≤ x ≤ 2) series based on electrochemical impedance spectra (EIS) at different measuring temperatures. In addition, the correlative σ and Ea, along with their EIS Nyquist plots of all products at 25 ℃ are demonstrated in Fig. 2f, Figure S5, and Table S3. As a result, all NaCl-poor samples exhibit improved conductivities and reduced Ea. The optimized Na0.5ZrCl4.5 composition achieves the highest σ of 8.1 × 10− 5 S cm− 1 and lowest Ea of 0.35 eV at 25 ℃, representing a more than six-fold enhancement compared to the pristine Na2ZrCl6 sample (1.3 × 10− 5 S cm− 1 at 25 ℃). Moreover, the electron conductivities of NaxZrCl4+x SEs are estimated to be approximately 10− 9 S cm− 1 (Figure S6, Table S7), indicating their favorable insulating properties. The significant improvement in conductivity of the NaCl-poor SEs likely arises from an increased number of vacant sites within the structures, which provide more free volume and larger ion transport channels with reduced bottleneck size. To be noted, the conductivity change trend of NaxZrCl4+x SEs displays a distinct volcano-shaped curve, suggesting that neither insufficient nor excessive NaCl within the structures (correlated with vacancy and carrier concentrations) is optimal for Na-ion diffusion.
To verify the aforementioned conjecture, we compared the changes in Na-ion content and vacant-site concentrations as mole number x increased (Fig. 2g and Figure S7). The NaCl-poor condition in Na2ZrCl6 reduces the Na-ion content in the Cl sublattices while increasing the vacancy concentrations (both Na and Cl vacant sites). For example, when x = 0.2 (NaxZrCl4+x), Na-ion in the Na0.2ZrCl4.2 structure occupies only 5% of the total ion sites, while the concentrations of Na and Cl vacant sites reach 95% and 30%, respectively (Fig. 2g (Ⅰ)). When x increases to 0.5, the electrolytes achieve their highest conductivity. At this point, both the Na-ion (12.5%) and vacancy contents (Na vacancy: 87.5%, Cl vacancy: 25%) reach their optimal thresholds (Fig. 2g (Ⅱ)). In stark contrast, further increasing the Na/Cl ion content (x = 0.8, 2) leads to decreased Na and Cl vacancy concentrations (Fig. 2g (Ⅲ-Ⅳ)), resulting in a corresponding reduction in conductivity (Figs. 2e-f). Based on the above analysis, we concluded that an optimal balance between mobile charge carrier and vacancy concentrations is crucial for efficient ion conduction in sodium superionic conductors. Building on this inference, we developed a series of Na-ion halide SEs systems with significantly improved conductivities by adjusting carrier and vacancy contents through facile poor-NaCl strategies. As demonstrated in Figures S8-9 and Tables S4-S5, the optimized compositions of Na0.6Er0.4Zr0.6Cl4.2 and Na0.4Yb0.25Zr0.75Cl4.15 exhibited superior conductivities of 5.7 × 10− 5 S cm− 1 and 2.0 × 10− 4 S cm− 1, respectively, compared to their pristine counterparts (Na2.4Er0.4Zr0.6Cl6: 1.9 × 10− 5 S cm− 1; Na2.25Yb0.25Zr0.75Cl6: 2.9 × 10− 5 S cm− 1) at 25 ℃. The above results strongly demonstrate that enhancing the conductivity of inorganic solid electrolytes by balancing vacancy and carrier concentrations is a universal and feasible strategy.
To elucidate the effect of electrolyte structural features on ion conduction, the bond valence site energy (BVSE) method was employed to calculate ionic transport channels and corresponding migration energy barriers. In the Na2ZrCl6 structure (Fig. 3a), the [Na1-Na2-Na1] chain along the [001] direction was identified as the most efficient 1D transport channel, with a migration barrier of 0.497 eV (marked in red in Figs. 3c-d and g). Additionally, the Na-ion can also migrate along the [Na2-i1-Na2], [Na2-i2-i3-Na1], and [Na1-i4-Na1] directions (marked in purple, blue/orange, and green, respectively, in Figs. 3c-d and g). These paths connect the [Na1-Na2-Na1] chain to form a 3D transport network, albeit with higher migration barriers of 0.673 eV, 0.696 eV, and 0.722 eV, respectively. In contrast, the Na0.5ZrCl4.5 structure (Fig. 3b) exhibits the lowest ion migration barrier of 0.479 eV along the [Na1-i1-Na1] direction (marked in red in Figs. 3e-f and h), effectively interconnecting the 1D [Na1-Na2-Na1] transport channels to form the most efficient 3D ionic percolating networks. Moreover, the other ionic transport channels along the [Na1-Na2-Na1] and [Na1-i2-Na1] directions (marked in green and blue, respectively, in Figs. 3e-f and h) also exhibit relatively low migration barriers of 0.591 eV. These findings further corroborate the more favorable Na-ion transport environments within NaCl-poor phase structures. It should be noted that BVSE analysis primarily characterizes ionic diffusion behaviors in crystal structures with periodic atomic arrangements and does not account for contributions from other non-ideal factors, such as amorphous matrices, defects, and surface structures. These features are typically nonperiodic owing to the robust ball-milling reactions. However, the crystalline component in NaCl-poor samples delivers lower migration energy barriers compared to pristine samples prepared under the same conditions. Thus, the effective transport barriers calculated from BVSE analysis provide powerful evidence of the significantly improved conductivities in NaCl-poor SEs. Furthermore, nonperiodic compositions likely play an important role in ionic migration, warranting systematic research applying advanced characterization techniques with exceptionally fine spatial resolution, such as aberration-corrected transmission electron microscopy32. Additionally, the conductivities of NaCl-poor electrolyte systems could be further improved by optimizing the proportions of nonperiodic components within the structures.
The general design strategy of the HNISEs in enhancing (electro)chemical stability
Ordinarily, the fluorine-ion substitution can effectively improve the (electro)chemical stability of HNISEs but at the cost of sacrificed conductivities due to the lower polarizability and smaller ionic radii of F− (119 pm) compared to Cl− (167 pm) or Br− (182 pm)35. Interestingly, in this study, the reduction in conductivity of fluorinated HNISEs can be well mitigated by the amorphization-induced protocol. Therefore, to further improve the electrochemical compatibility of NaCl-poor SEs with electrodes, varying amounts of F− were incorporated into the Na0.5ZrCl4.5−yFy (0 ≤ y ≤ 1) series through mechanochemical synthesis. XRD results (Fig. 4a) show that the pattern peaks diminish as the F− content increases, indicating that these fluorinated SEs tend to become amorphous during the synthesis procedures. Concurrently, Raman spectrums further suggest the successful fluorination of Na0.5ZrCl4.5 without introducing additional impurities (Fig. 4b). Furthermore, the peak positions in the spectra of fluorinated samples emerge slight left shift (higher wavenumber) compared to pristine Na0.5ZrCl4.5 (~ 328, 158, and 124 cm− 1), primarily due to the formation of shorter F-Zr bonds within the structures. The SEM image (Fig. 4c) of Na0.5ZrCl4F0.5 reveals small particle sizes (< 5 µm) consisting of irregular micro- and nano-particles. The reduced average particle and non-crystalline structure result in denser cold-pressed pellets (~ 98.2% relative density, Table S8) with decreased porosity (Figure S10a). The improved compressibility of the powder samples is more favorable for intimate solid-solid interface contact between electrode materials and SEs36. Furthermore, the corresponding energy-dispersive X-ray spectroscopy (EDS) mapping implies the uniform dispersion of Zr, Cl, and F elements within the samples, indicating successful chemical fluorination of Na0.5ZrCl4.5 rather than physical mixing of fluorides (Figures S10b-d).
The ion conduction properties of the Na0.5ZrCl4.5−yFy (0 ≤ y ≤ 1) compositional series were assessed via EIS and Arrhenius plots (Figures S11 a-b). The conductivities of Na0.5ZrCl4.5−yFy SEs reveal a pattern characterized by an initial increase followed by a subsequent decrease. At an F doping content near 0.5, the electrolyte achieves the highest σ of 1.12 × 10− 4 S cm− 1 at 25 ℃ (Fig. 4d). Furthermore, the Na0.5ZrCl4F0.5 SEs display electron conductivities of approximately 10− 10 S cm− 1 (Figure S11c, Table S7), exhibiting excellent electron insulating properties. In brief, when the F doping content (y) is ≤ 0.5, the contribution of the amorphous effect to ion transport exceeds the constrained force between F− anion and Na+ cation. Thus, the conductivities of Na0.5ZrCl4.5−yFy (0 < y ≤ 0.5) SEs exhibited a growth trend. Whereas, for y > 0.5, the conductivities dramatically decrease because of the stronger attraction between F− and carriers. Recent studies have demonstrated that the amorphization of SEs can significantly improve their ion conduction capabilities, owing to the unique local chemistry environments and ion migration mechanisms in amorphous domains11,37–40. Consequently, the increased conductivity observed with partial substitution of F− is mainly attributed to the amorphization effect induced by fluorination. However, excessive F− tends to form robust bonds with mobile charges, impeding the effective diffusion of Na+ within crystal structures. Furthermore, the mechanism underlying the conductivity enhancement in amorphous fluorinated-based SEs requires further investigation in future studies. Given the favorable balance between ion conduction and (electro)chemical stability in fluorinated SEs, the optimized Na0.5ZrCl4F0.5 samples will be the primary focus of subsequent research.
Another critical challenge for HNISEs is their sensitivity to moisture due to the deliquescent properties of most halides41. In this work, the moisture stability of various SEs was assessed using EIS Nyquist plots and XRD patterns. After 10 min of exposure to ambient relative humidity (RH, 35 ± 5%), the XRD pattern of Na0.5ZrCl4F0.5 showed no additional impurity peaks (Figure S12). Furthermore, the conductivity of Na0.5ZrCl4F0.5 remained stable after humidity exposure (Fig. 4e), demonstrating good moisture tolerance. In contrast, the XRD pattern of the Na0.5ZrCl4.5 pellet delivered attenuated peak intensity and reduced conductivity following similar moisture treatments (Figure S13). The improved humidity stability of Na0.5ZrCl4F0.5 can be ascribed to the stronger Zr/Na-F bonds formed within the structures through F substitution24,42,43.
The electrochemical stability of Na0.5ZrCl4F0.5 was evaluated using cyclic voltammetry (CV) tests in an asymmetric configuration. Vapor-grown carbon fibers (VGCF)/SEs composites were used as the working electrode, while a Na15Sn4 (0.1 V vs Na/Na+) alloy served as the reference electrodes. In composite electrodes, the incorporation of VGCF allows for continuous and efficient electron transport, enabling precise monitoring of electrochemical reaction signals44. Figure 4f demonstrates the positive scan curves for various SEs at a scan rate of 0.1 mV s− 1. Notable, Na0.5ZrCl4F0.5 exhibits the lowest integrated current of 1.34 mAV g− 1, in contrast to Na2ZrCl6 (4.35 mAV g− 1) and Na0.5ZrCl4.5 (4.26 mAV g− 1), manifesting the fluorination process significantly enhances the oxidation stability of HNISEs within high-voltage regimes. X-ray photo-electron spectroscopy (XPS) was employed to assess the high-voltage stability of these HNISEs. When oxidation is up to 5 V, oxidation peaks related to \(\:{\text{Cl}}_{\text{x}}^{\text{-}}\) are evident in the spectra of Na2ZrCl6 and Na0.5ZrCl4.5, indicating their relative instability at elevated voltages (Figs. 4g-h). In contrast, the spectra for Na0.5ZrCl4F0.5 (Figs. 4i-j) show no obvious oxidation peaks for \(\:{\text{Cl}}_{\text{x}}^{\text{-}}\) and \(\:{\text{F}}_{\text{x}}^{\text{-}}\), highlighting its exceptional anti-oxidation properties attributed to the stronger electronegativity of fluorine. Additionally, Zr 3d spectra exhibit minimal variation, indicating the maintenance of Zr in its highest oxidation state, Zr4+ (Figure S14).
In addition to impressive oxidation stability, Na0.5ZrCl4F0.5 also demonstrates a decreased absolute reduction current of 0.0266 mA at 0 V (Figure S15), corroborating the improvement of intrinsic reduction stability against Na15Sn4 anode by fluorination strategy. Further analysis of the interface kinetic stability between Na0.5ZrCl4F0.5 and Na15Sn4 involved cycling Na15Sn4/Na15Sn4 symmetrical cells. As illustrated in Figure S16, the initial plating/stripping potential of the Na0.5ZrCl4F0.5 and Na0.5ZrCl4.5-based cells were approximately 0.8 V at a current density of 0.1 mA cm− 2 and a capacity of 0.1 mAh cm− 2. Remarkably, the Na0.5ZrCl4F0.5-based symmetrical cell maintained stable plating/stripping for 800 h with minimal overpotential increases (~ 1 V), whereas its Na0.5ZrCl4.5-based counterpart delivered a progressive overpotential increased to approximately 2.6 V over the same duration. In contrast, the Na2ZrCl6-based symmetrical cell displayed a significant overpotential increase and shorter plating/stripping times (exceeding 5 V within 73 h), primarily due to the lowest conductivity and poor reduction stability of the electrolytes. Additionally, the improved cycling stability in Na15Sn4/Na0.5ZrCl4F0.5/Na15Sn4 cells likely originates from the in-situ formation of a kinetically stable interphase layer between the fluorinated SEs and the anode interface (e.g., NaF, see Figure S17), which helps mitigate ongoing parasitic reactions during cycling25,45–47.
Although Na0.5ZrCl4F0.5-based symmetrical cells exhibit reduced plating/stripping overpotentials, these remain relatively high (~ 0.8 V) to effectively release the full battery's capacity. Therefore, the Na3PS4 (NPS, ~ 1.4 × 10− 4 S cm− 1 at 25 ℃), known for its exceptional kinetic stability against Na-Sn alloys15,48, was incorporated as an intermediate electrolyte layer between the HNISEs and the Na15Sn4 anode. Moreover, recent studies have shown that fluorination of halides is advantageous for promoting interfacial compatibility with sulfides26. We explored this by comparing the electrochemical behaviors of two symmetrical cell configurations: both utilized HNISEs as interlayers and NPS as outer layers. As depicted in Figure S18, both configurations initially exhibited low overpotentials (~ 0.2 V). However, the Na0.5ZrCl4.5-based sandwich symmetrical cell presented increasing plating/stripping potentials over time, signaling interfacial incompatibility between Na0.5ZrCl4.5 and NPS during cycling. In contrast, the Na0.5ZrCl4F0.5-based sandwich symmetrical cell maintained excellent cycling stability with negligible overpotential increase, underscoring the beneficial impact of fluorine doping on halide/sulfide interfacial compatibility. Based on these findings, a dual-electrolyte layer strategy was adopted for ASSIBs, positioning Na0.5ZrCl4F0.5 adjacent to the cathode and NPS near the Na15Sn4 alloy anode.
Electrochemical energy storage behaviors of ASSIBs
The electrochemical performance of mechanochemically synthesized Na0.5ZrCl4F0.5, Na0.5ZrCl4.5, and Na2ZrCl6 HNISEs in ASSIBs was evaluated using Na3V2(PO4)3 (NVP) cathodes and VGCF conductive additives, as shown in Fig. 5a. Detailed material characteristics are provided in Figures S19-20. For comparative purposes, we constructed a single-layer cell configuration employing only NPS, without HNISEs (Figure S21a). As depicted in Fig. 5b, under a voltage window of 1.9 to 3.7 V at RT and a 0.1 C rate, the initial Coulombic efficiency (ICE) of the Na0.5ZrCl4F0.5 cell was markedly higher at 94.5%, with a discharge capacity of 86.4 mAh g− 1, in contrast to the Na0.5ZrCl4.5 and Na2ZrCl6 cells, which achieved ICE of 88.7% and 84.2% with discharge capacities of 70.3 mAh g− 1 and 57.7 mAh g− 1, respectively. The enhanced electrochemical stability of the Na0.5ZrCl4F0.5 electrolyte appears to shield the other HNISEs and NPS from oxidation through interaction with NVP. Nonetheless, the cell employing NPS electrolyte exhibited a rapid decline in discharge capacity and increased interfacial resistance following the initial charge, attributed to the poor cathodic stability under high voltages, as further detailed in Figures S21b-c49.
The rate and cycle capability of batteries employing Na0.5ZrCl4F0.5, Na0.5ZrCl4.5, and Na2ZrCl6 as catholytes demonstrate comparable capabilities, as shown in Figs. 5c and d. Specifically, the Na0.5ZrCl4F0.5 cell exhibits a higher specific capacity of 46.5 mAh g− 1 at 0.3 C, significantly outperforming the capacities of 25.3 and 4.6 mAh g− 1 observed in the Na0.5ZrCl4.5 and Na2ZrCl6 cells, respectively. When the current density is reduced back to 0.1 C, a reversible capacity of 82.3 mAh g− 1 is restored. The improved rate capacity of the Na0.5ZrCl4F0.5 cells primarily stems from their higher conductivity and enhanced compatibility with fluorinated electrolytes. It is noteworthy that the capacity limitations at higher rates are largely due to the suboptimal conductivities of the electrolyte layers, which are less than 1 mS cm− 1. The Na0.5ZrCl4F0.5-based cell also displays a substantial capacity retention of 94.4% over 300 cycles at 0.1 C, markedly surpassing the performance of Na0.5ZrCl4.5 and Na2ZrCl6 cells, which retain only 49.5% and 17.7% of their initial capacities after 150 and 100 cycles, respectively. Furthermore, the lower overpotential and enhanced cycling reversibility of the Na0.5ZrCl4F0.5 cell are further demonstrated in the charge/discharge voltage profiles and corresponding dQ/dV integral curves (Figures S22-23).
The EIS tests were conducted to assess the interfacial stabilities between cathodes and HNISEs in both pre- and post-long-term cycling. As illustrated in Figure S24, Nyquist plots present two distinct semicircles at high and mid frequencies, corresponding to the impedance of separating electrolyte layers and composite cathode interfaces, respectively50. Following extended cycling, the Na0.5ZrCl4F0.5 cell (~ 711 Ω) demonstrated significantly lower interfacial resistance compared to Na0.5ZrCl4.5 (~ 4237 Ω) and Na2ZrCl6 (~ 7346 Ω) cells, indicating superior electrochemical stability of Na0.5ZrCl4F0.5 catholytes during cycling. Moreover, the in-situ galvanostatic-EIS (GEIS) combined with the distribution of relaxation time (DRT) method was performed to evaluate the detailed electrochemical evolution of HNISE cells during charging. Figures S25a-c show no distinct changes in resistance during the charging process (2.9 to 3.7 V) for the Na0.5ZrCl4F0.5 cell, yet notable impedance growth is observed in Na0.5ZrCl4.5 and Na2ZrCl6 analogs. Semi-quantitative DRT results (Figures S25d-f) categorize relaxation time (τ) into four groups—S1, S2, S3, and S4—across these three cells. Peak changes in the S3 (10− 4 < τ < 100 s) and S4 (100 < τ < 101 s) regions are predominantly linked to the cathode/electrolyte interface and diffusion resistance, respectively51. For the Na0.5ZrCl4F0.5 cell, peak intensities in the S3 and S4 regions show minimal changes with relaxation time, underscoring excellent electrochemical stability within composite cathodes. Conversely, peak intensities in S3 and S4 regions increase notably with relaxation time for Na0.5ZrCl4.5 and Na2ZrCl6 cells, indicative of irreversible parasitic reactions at cathode/electrolyte interfaces, leading to deteriorating interface stability and heightened cell impedance. In conclusion, based on EIS and DRT analyses, Na0.5ZrCl4F0.5 catholytes exhibit superior high-voltage stability and interface compatibility with polyanionic-type positive electrodes, thereby boosting ASSIBs' performance.
To investigate the influences of chemical-mechanical failure on full cells, we compared the micro-morphological evolution of composite cathode surfaces using SEM images of pristine and cycled samples. Figure 6a illustrates that all three electrode types show similar microcrack areas on their surfaces before cycling, manifesting insufficient connection between HNISEs and cathode materials. After several hundred cycles, the Na0.5ZrCl4F0.5-based composite electrodes delivered a uniform and dense structure with no discernible microcracks, in contrast to the other two types, which displayed more and longer cracks. This conversion stems from the continuous enhancement of NVP/Na0.5ZrCl4F0.5 interfacial mechano-chemical stability during cycling, alongside the inherent high deformability of fluorinated SEs. Additionally, Na-ion diffusion coefficients (\(\:{\text{D}}_{{\text{Na}}^{\text{+}}}\)) for three configurations were calculated through the galvanostatic intermittent titration technique (GITT), a standard method for assessing ion transport kinetics within composite cathode52. As demonstrated in Fig. 6b and Figures S26, the Na0.5ZrCl4F0.5-based cathode composites display higher \(\:{\text{D}}_{{\text{Na}}^{\text{+}}}\) and lower overpotential during charging, affirming the favorable kinetic and thermodynamic stability of Na0.5ZrCl4F0.5 catholyte within the electrodes. Furthermore, the calculated interfacial reaction energies between HNISEs and electrodes provide further support for these observations (Fig. 6c and Table S9). The Na0.5ZrCl4F0.5 (~ 84 meV) catholytes and their phase equilibria products (Na2ZrF6 ~ 15.5 meV) exhibit a lower interface reaction driving force with NVP cathodes compared to the Na0.5ZrCl4.5 (~ 88 meV) counterpart, suggesting a more favorable interfacial compatibility between the fluorinated electrolyte and NVP. It is noteworthy that the mutual reaction energies between HNISEs and anodes (Na metal and Na15Sn4) are significantly higher than those at the cathode interface, highlighting the inherent thermodynamic instability of the HNISEs/anode interface. However, favorable interfacial decomposed insulating phases (such as NaF and NaCl) can be formed from fluorinated SEs during cycling, which helps inhibit the interfacial parasitic reactions and the continuous decomposition of HNISEs25,53. Thus, fluorinated SEs demonstrate greater stability at the anode interface, consistent with the stable cycling observed in Na0.5ZrCl4F0.5-based symmetrical cells (Figure S16). In a word, the Na halide SEs could achieve significantly improved RT conductivities using the NaCl-poor strategy owing to an effective balance of vacancy and carrier concentrations within their crystal structures. Moreover, the introduction of fluorine into structures further enhanced their (electro)chemical stability and interface compatibility. Surprisingly, the fluorinated electrolyte also demonstrated boosted conductivity, deformability, and moisture stability, largely attributed to the amorphous effect and stronger bonding within structures (as summarized and illustrated in Figs. 6d and e).
In conclusion, we have developed a universal strategy for enhancing ion transport in Na halide-based SEs by adopting a NaCl-poor method. This approach optimally adjusts vacancy and carrier concentrations within the lattice, establishing critical balancing concentration thresholds that significantly enhance Na+ diffusion efficiency. As a result, the conductivities of these NaCl-poor SEs substantially exceed those of their standard counterparts. Furthermore, integrating fluoride into zirconium-based chloride SEs not only augments their electrochemical stability and interface compatibility with electrodes but also increases conductivity and moisture tolerance. This improvement is primarily attributed to the formation of an amorphous phase and the strengthened Zr-F bonds within the electrolyte matrix. Employing Na3V2(PO4)3 as a cathode and Na0.5ZrCl4F0.5 as a catholyte, ASSIBs exhibit excellent cycling stability, maintaining approximately 94% of their initial capacity after 300 cycles at a discharge rate of 0.1 C. This study lays the groundwork for the design of advanced Na-ion halide SEs, optimizing both conductivity and (electro)chemical stability, thus promoting the durability and safety of ASSIBs.