Zichen Li†
,
Mengdi Zhang†,
Yingchun Yan*,
Weimin Zhang,
Xiuxia Meng and
Naitao Yang*
School of Chemical Engineering, Shandong University of Technology, Zibo 255049, China. E-mail: naitaoyang@126.com; yingchyan07@163.com
First published on 23rd August 2025
Sodium-ion batteries (SIBs) have emerged as a sustainable alternative to lithium-ion batteries (LIBs), benefiting from the natural abundance and low cost of sodium resources. However, the inherent safety risks of conventional organic liquid electrolytes necessitate the development of robust solid-state electrolytes (SSEs). NASICON-structure Na3Zr2Si2PO12 (NZSP) represents a promising SSE candidate due to its superior ionic conductivity, wide electrochemical stability window, and excellent chemical compatibility. Nevertheless, polycrystalline NZSP suffers from high grain boundary resistance and interfacial instability, limiting its practical application. Herein, a Sc3+/Zn2+ co-doped strategy was developed to synthesize Na3.2+2xZr1.8−xSc0.2ZnxSi2PO12 (NZSSP-Znx, x = 0–0.20), the incorporation of Sc3+ stabilizes the highly conductive rhombohedral phase and expands Na+ transport channels, while Zn2+ doping lowers the sintering temperature of monoclinic-to-rhombohedral transition in NZSSP-Zn0.10. The synergistic co-doped approach employing NZSSP-Zn0.10 demonstrates remarkable enhancement in both bulk and grain boundary conductivity. During thermal processing, the precursor materials undergo in situ transformation to generate a Na3PO4 secondary phase, exhibiting a high ionic conductivity of 2.41 × 10−3 S cm−1, and a low activation energy of 0.20 eV. The distribution of relaxation times (DRT) method was applied to analyze the time-resolved electrochemical impedance spectroscopy (EIS), elucidating the correlation between dynamic interfacial changes and electrochemical processes at the Na/NZSSP-Zn0.10 SSE interface. The Na|NZSSP-Zn0.10|Na symmetric cell achieves a high critical current density (CCD) of 0.9 mA cm−2 and stable Na plating/stripping for 550 h at 0.1 mA cm−2. The assembled Na|NZSSP-Zn0.10|NVP/C demonstrates excellent cycling stability (0.029% decay per cycle over 400 cycles at 0.5 C). This work provides a feasible co-doped approach to optimize NASICON electrolytes for high-performance solid-state sodium batteries.
Broader contextSolid-state sodium-ion batteries exhibit enormous potential in addressing the safety issues and performance challenges inherent in traditional liquid batteries, making them a key research direction for next-generation sodium-ion batteries. However, ideal solid-state electrolytes typically need to meet high ionic conductivity at room temperature (>10−4 S cm−1) and good interfacial compatibility with solid-state electrode materials. The rhombohedral-phase Na3Zr2Si2PO12 (NZSP) exhibits superior Na+ conductivity, attributed to its lower activation energy (Ea) for ion diffusion resulting from structurally expanded transport channels. In this study, a Sc3+/Zn2+ co-doping strategy was developed to synergistically reduce grain boundary resistance, enhance ionic conductivity, and improve interfacial compatibility in the solid electrolyte. The resulting Na3.2+2xZr1−xSc0.2ZnxSi2PO12 (NZSSP-Zn0.10)-based solid-state batteries exhibit exceptional cycling stability, positioning this material as a highly promising candidate for next-generation high-performance solid-state batteries. |
NASICON-structured Na3Zr2Si2PO12 (NZSP), possesses a structure consisting of corner-sharing ZrO6 octahedra and PO4/SiO4 tetrahedra. This arrangement forms a three-dimensional network, and the interconnected channels within the framework function as efficient Na+ transport pathways, enhancing the ionic conductivity of the electrolyte.9 The crystal structure of NZSP is highly tunable based on sintering conditions and composition. NZSP typically crystallizes into a rhombohedral phase at high temperatures (0 < x < 1.8, 1.8 < x < 2.3, space group Rc) and a monoclinic phase at low temperatures (1.8 ≤ x ≤ 2.3, space group C2/c).10,11 Critically, the rhombohedral phase exhibits superior Na+ conductivity due to its lower activation energy (Ea) for ion diffusion, derived from the expanded diffusion channels and optimized structural symmetry.12–14 Using NZSP as a basic structure, researchers have made many attempts to improve grain conductivity by doping metal ions (such as Mg2+, Al3+, Pr3+, Y3+, Lu3+, Ge4+) at the Zr site (rM = 0.72 Å) of NZSP. By doping with larger metal ions (rM > 0.72 Å), the Na+ transport channel can be effectively enlarged, stabilizing the formation of the rhombohedral phase and thereby reducing the activation energy of ion migration.12,15–19 At the same time, to compensate for the positive charge vacancies generated by the substitution of lower-valence cations, more Na+ carriers are introduced, thereby increasing the ionic conductivity.10,19–22 Despite the excellent bulk ionic conductivity (∼10−3 S cm−1) of NZSP, the high grain boundary resistance and instability at the electrode–electrolyte interface in their polycrystalline structure severely limit practical applications.
Previous studies have demonstrated that incorporating a conductive second phase at the grain boundaries of NZSP effectively reduces grain boundary resistance and promotes the migration of Na+ at the grain boundaries. The in situ self-reconstruction of mesophase components (Na2ZrSi2O7 and Na3PO4) facilitates the formation of an amorphous NASICON interphase at grain boundaries.23 This structural evolution significantly enhances intergranular connectivity, resulting in exceptional grain-boundary conductivity reaching 36 mS cm−1. This value represents a sevenfold enhancement compared to the baseline NZSP electrolyte (5.1 mS cm−1). Moreover, La-doped NZSP can form a Na3PO4 second phase, which facilitates ion transport along the grain boundaries. However, this approach may also introduce insulating impurity phases, such as LaPO4 and La2O3 may be generated during the doping process, which can seriously hinder the transport of Na+.24 Therefore, developing a strategy for self-forming conductive second phases without extrinsic dopants is crucial for improving the grain boundary conductivity of NZSP.
In this work, Sc3+/Zn2+ co-doped Na3.2+2xZr1.8−xSc0.2ZnxSi2PO12 (NZSSP-Znx, x = 0–0.20) were designed by conventional solid-state reaction (Fig. 1a). The incorporation of Sc3+ serves to expand Na+ transport channels while stabilizing the rhombohedral phase, whereas Zn2+ doping effectively lowers the sintering temperature of rhombohedral NZSP and enhances electrolyte densification. Sc3+/Zn2+ co-doped creates additional Na+ vacancies, yielding remarkable improvements in ionic conductivity (2.41 × 10−3 S cm−1) and reduced activation energy (0.20 eV). During the sintering process, raw materials in situ form a Na3PO4 secondary phase, further optimizing the Na+ transport dynamics at the grain boundaries. This phase significantly improves Na+ transport dynamics, as revealed by the distribution of relaxation times (DRT) analysis of in situ electrochemical impedance spectroscopy (EIS) data. The optimized NZSSP-Zn0.10 develops a stable SEI layer during cycling by uniform Na deposition and stripping. When implemented in a Na|NZSSP-Zn0.10|NVP/C full cell, this electrolyte demonstrates outstanding rate capability and long cycling stability (only 0.029% capacity decay in each cycle over 400 cycles at 0.5 C).
The surface morphology of NZSSP-Znx (x = 0–0.20) pellets was analyzed via Scanning Electron Microscopy (SEM) for microstructural characterization (Fig. 1c). As shown in Fig. S1b, the surface SEM image of NZSP, reveals that the electrolyte pellets consist of numerous small grains with distinct grain boundaries. Similarly, the SEM image of NZSSP-Zn0 also shows small grains and clear grain boundaries. Interestingly, with the progressive increase in Zn ion content, the grain size of the electrolyte gradually enlarges, accompanied by a corresponding blurring of the grain boundaries. This structural evolution effectively diminishes the obstacles encountered by Na+ during transport, thereby significantly enhancing the overall Na+ mobility and facilitating more efficient ion conduction within the electrolyte.10,17,28 Notably, when Zn ion doping exceeds a certain level (x > 0.10), the grains of the electrolyte pellets exhibit anisotropic overgrowth, resulting in the formation of crack defects within the grains. This phenomenon disrupts the densification of the electrolyte pellets and impedes Na+ transport across grain boundaries.15,29 To confirm the successful incorporation of Sc and Zn ions into the main phase of the electrolyte, elemental distribution in the NZSSP-Zn0.10 electrolyte was analyzed via energy-dispersive spectroscopy (EDS). The elemental mapping demonstrates a homogeneous distribution of Na, Zr, Si, P, Sc, and Zn (Fig. 1d), confirming the successful integration of Sc3+ and Zn2+ into the main phase of the electrolyte.
The EIS profiles of NZSSP-Znx (x = 0–0.20) electrolyte pellets were measured over a temperature range of 30–100 °C (Fig. 2a, S2a, and S3). The ionic conductivity (σt) of the electrolyte pellets exhibits an initial increase followed by a subsequent decrease as the Zn ion content gradually increases. Specifically, the σt of NZSSP-Zn0 is measured to be 9.80 × 10−4 S cm−1, which is significantly higher than that of NZSP (2.13 × 10−4 S cm−1). With the incorporation of Zn ions, the σt of NZSSP-Zn0.10 reaches 2.41 × 10−3 S cm−1. However, with further increments in Zn ion content, the σt of NZSSP-Zn0.15 and NZSSP-Zn0.20 decreased to 1.25 × 10−3 S cm−1 and 1.02 × 10−3 S cm−1, respectively (Table S1 demonstrates detailed date on ionic conductivity for NZSSP-Znx (x = 0–0.20) electrolyte pellets), which attributed to the formation of crack defects within the grains due to grain anisotropic overgrowth, thereby compromises ionic conductivity.17,30,31 The enhanced ionic conductivity achieved through the doping of Na3Zr2Si2PO12 with Sc and Zn ions can be attributed to two primary factors: (1) the slightly larger ionic radius of Sc ions (0.745 Å) and Zn ions (0.74 Å) compared to Zr ions (0.72 Å), stabilize the rhombohedral phase with a larger unit cell volume, thereby creating a more spacious pathway for Na+ transport; (2) the introduction of Zn ions promotes the densification of NZSSP-Zn0.10, while the presence of Na3PO4 as a secondary phase at the grain boundaries reduces grain boundary resistance, further minimizing obstacles to Na+ migration.13,17
The activation energy for ion transport was calculated according to the Arrhenius equation (Fig. 2b and S2b), NZSSP-Zn0.10 with rhombohedral in NASICON phase exhibits an exceptionally low activation energy of 0.20 eV, significantly lower than those of NZSP (0.31 eV) and NZSSP-Zn0 (0.24 eV). This reduction in activation energy reflects that NZSSP-Zn0.10 creates a more spacious migratory pathway, facilitating the transport of Na+. Furthermore, the presence of Na3PO4 at the grain boundaries reduces the resistance experienced by Na+ during grain boundary transport, thereby significantly accelerating the Na+ transport process. Due to this structural optimization, the Na+ transference number (tNa+) of NZSSP-Zn0.10 is 0.98, higher than that of NZSSP-Zn0 (0.79), indicating a stronger Na+ transport capability in NZSSP-Zn0.10 (Fig. 2c and S4).32,33 The electrochemical window of the NZSSP-Znx (x = 0–0.20) electrolyte was evaluated by linear sweep voltammetry (LSV) (Fig. 2d). NZSSP-Zn0.10 exhibits an optimal electrochemical stability window of 4.9 V, exceeding the stability windows of other NZSSP-Znx (x = 0, 0.05, 0.15, 0.20) electrolytes (Fig. S5). The critical current density (CCD) serves as a key indicator for evaluating interfacial stability in Na|NZSSP-Zn0|Na and Na|NZSSP-Zn0.10|Na symmetric cells.34,35 The Na plating/stripping plots with progressively increasing current densities are shown in Fig. 2e. Compared with NZSSP-Zn0, which exhibits a CCD of only 0.1 mA cm−2. The irregular voltage fluctuations of Na|NZSSP-Zn0|Na symmetric cell can be ascribed to the unstable Na plating/stripping and the formation of dead Na.23 While the CCD of NZSSP-Zn0.10 displays a significantly enhanced CCD of 0.9 mA cm−2, demonstrating that its dense structure can suppress the growth of Na dendrites and thereby enhance interfacial stability. The performance metrics achieved in this work belong to a high level among the existing literature reports (Fig. 2f and Table S2).12,16,26,28,36,37
Long-term Na plating/stripping cycling tests of Na|SSEs|Na symmetric cells were performed to further assess the interfacial stability between metallic Na and both NZSSP-Zn0 and NZSSP-Zn0.10 electrolytes (Fig. 3a). At a current density of 0.1 mA cm−2 and 30 °C, the Na|NZSSP-Zn0|Na symmetric cell experienced a micro-short circuit after only 70 h, accompanied by high overpotential (125.5 mV) during testing. In contrast, Na|NZSSP-Zn0.10|Na shows steady Na insertion/extraction processes for over 650 h with significantly lower overpotential (66.5 mV). As shown in Fig. 3b and S6, the Tafel curves reveal superior interfacial kinetics for NZSSP-Zn0.10 with an exchange current density (2.21 × 10−2 mA cm−2), which is higher than that of NZSSP-Zn0 (0.96 × 10−2 mA cm−2). These results demonstrate that Zn incorporation substantially improves the interfacial contact between the Na metal and the NZSSP-Zn0.10 electrolyte, establishing a more kinetically stable charge transfer interface.38 Furthermore, the interfacial stability of the Na||Na symmetric cells was systematically investigated through time-resolved EIS during galvanostatic cycling. Specifically, the Na|NZSSP-Zn0.10|Na symmetric cell demonstrated stable operation for 150 h, with time-resolved EIS capturing the temporal evolution of interfacial properties. Parallel testing of the Na|NZSSP-Zn0|Na symmetric cell was conducted for 100 h under identical experimental conditions to enable direct comparison of interfacial behaviors. As illustrated in Fig. 3c and Table S3, the time-dependent Nyquist plots for the Na|NZSSP-Zn0.10|Na symmetric cell reveal distinct electrochemical characteristics. The intercept on the real axis corresponds to the bulk resistance (Rb) of the NZSSP-Zn0.10 electrolyte pellet. The first semicircle represents the grain boundary resistance (Rgb) of NZSSP-Zn0.10, while the second semicircle is attributed to the interfacial resistance (Rint) at the Na/NZSSP-Zn0.10 interface.39 The Na|NZSSP-Zn0|Na symmetric cell exhibited an initial interfacial resistance of 1551 Ω, which increased dramatically to 4596 Ω after 100 h. This interfacial degradation leads to higher overall resistance, corresponding to uneven Na plating and stripping. In contrast, the Na|NZSSP-Zn0.10|Na symmetric cell exhibited an initial interfacial resistance of 407.8 Ω and a total resistance of 567.1 Ω. During cycling up to 30 h, Rint increases gradually to 567.2 Ω with a corresponding rise in total resistance to 740.98 Ω, attributed to the formation of intermediate phases at the interface. After 50 h, the interfacial resistance stabilizes at 567.3 Ω with a corresponding total resistance of 764.8 Ω. Notably, the interfacial impedance remained relatively constant until 150 h without significant fluctuations (Fig. 3d and Table S4), indicating better Na/NZSSP-Zn0.10 interface contact compared to the undoped NZSSP-Zn0 electrolyte. Such remarkably stable behavior confirms that NZSSP-Zn0.10 develops robust interfacial properties during prolonged cycling.
To elucidate the dynamic interfacial changes at the Na/SSE interface and their correlation with interfacial electrochemical processes, the DRT method was employed to analyze the time-resolved EIS. Each relaxation time in the DRT spectrum corresponds to a distinct electrochemical process. The DRT analysis of the Na|SSEs|Na symmetric cells revealed three prominent peaks, denoted as τ1, τ2, and τ3 (Fig. 3e and f). Specifically, the relaxation time τ1 is associated with the resistive properties of grain boundaries in the SSE, directly reflecting the efficiency of Na+ transport within the grain boundary regions of the electrolyte. τ2 is attributed to the solid–electrolyte interphase (SEI) at the Na/SSE interface. Meanwhile, τ3 corresponds to the charge transfer process of Na+ at the Na/SSE interface.18,40,41 In comparison to the Na|NZSSP-Zn0|Na symmetric cell, the Na|NZSSP-Zn0.10|Na symmetric cell exhibited a lower τ1 peak intensity, suggesting reduced Na+ transport resistance across the grain boundaries of the NZSSP-Zn0.10 electrolyte. As the cycling progressed, τ2 shifted slightly toward higher frequencies, which was attributed to the formation of the SEI layer during the initial stages of cycling. Concurrently, the τ3 peak intensity stabilizes after 50 h, indicating that the Na|NZSSP-Zn0.10|Na symmetric cell gradually forms a stable SEI layer during cycling. These findings suggest that the Na|NZSSP-Zn0.10|Na symmetric cell develops a stable SEI layer during cycling, which enhances Na+ transport kinetics, promotes uniform Na deposition and stripping, and effectively suppresses interfacial side reactions.39,42 In contrast, the Na|NZSSP-Zn0|Na symmetric cell exhibits a continuous increase in τ3 peak intensity, indicative of ongoing side reactions at the Na/NZSSP interface. These reactions result in progressive deterioration of interfacial conditions and a sustained increase in resistance.
To investigate the chemical composition of the SEI at the Na/NZSSP-Zn0.10 interface during Na plating/stripping processes, X-ray photoelectron spectroscopy (XPS) analysis was employed to analyze the electrolyte surface before and after cycling. A clear increase in the Na 1s peak intensity is observed in the XPS spectra of NZSSP-Zn0.10 after cycling, reflecting Na enrichment at the interface. This change directly correlates with the detected alterations in the core-level spectra. As verified in the positions of Zr 3d, Sc 2p, and Zn 2p XPS spectra, the binding energy peaks remain largely invariant, demonstrating the exceptional chemical stability of NZSSP-Zn0.10 against the corrosive effects of highly reducing Na metal. Moreover, the Si 2p peak exhibits a shift toward lower binding energy after cycling, indicative of a slight reduction of Si4+.37 The P 2p spectra consistently exhibited a P peak at approximately 133.1 eV both before and after cycling, stemming from the P–O bonds in NZSSP-Zn0.10 and Na3PO4.27,43 Notably, a new peak emerged in the post-cycling P 2p spectra at 129.8 eV, corresponding to the Na3P peak, and a distinct O 1s peak at 529.7 eV attributed to Na2O (Fig. 4a and S7).44,45 The formed SEI layer ensures uniform electric field distribution, suppresses continuous side reactions at the interface, and accelerates Na+ transport at the interface, thereby achieving stable cycling of the Na/NZSSP-Zn0.10 interface. In contrast, after cycling, the Si 2p peak of NZSSP-Zn0 shifted slightly to lower binding energy, indicating Si4+ reduction.19 Meanwhile, the peaks of Zr 3d, Sc 2p, and P 2p shifted to higher binding energy, suggesting oxidation of NZSSP-Zn0 during cycling (Fig. 4b and S8). Such an unstable interface is prone to causing non-uniform Na plating/stripping, leading to increased polarization and dendrite formation, which can pose potential risks to battery performance and safety.24 The surface morphology of the cycled NZSSP-Zn0.10 electrolyte was observed by SEM to verify its inhibitory effect on Na dendrite growth (Fig. 4c and d). The NZSSP-Zn0 electrolyte exhibits severe Na dendrite formation, with dendritic structures propagating through internal pore channels. In contrast, only a small number of vitreous Na dendrites were obtained along the grain surfaces within the NZSSP-Zn0.10 electrolyte. This significant difference strongly demonstrates the superiority of the NZSSP-Zn0.10 electrolyte in inhibiting Na dendrite growth by regulating uniform Na plating and stripping.46–49 Furthermore, to directly assess the structural stability of the NZSSP-Zn0.10electrolyte, XRD analysis was carried out on the electrolyte after cycling in the Na|NZSSP-Zn0.10|Na symmetric cell (Fig. S9). Critically, the XRD patterns reveal that the crystal structure of NZSSP-Zn0.10 remains largely unchanged after prolonged cycling, showing no detectable signs of decomposition or phase transformation into less conductive secondary phases. This observation provides direct experimental evidence for the exceptional structural integrity of the Sc/Zn co-doped NASICON electrolyte during electrochemical cycling.
The NZSSP-Zn0.10/NZSSP-Zn0 solid electrolyte was applied in the Na|NASICON SSEs|Na3V2(PO4)3/C (NVP/C) cell to verify the practical application (Fig. 5a). As illustrated in Fig. 5b and c, the Na|NZSSP-Zn0.10|NVP/C cell exhibits specific discharge capacities of 111.05, 109.73, 102.81, 96.76, 87.21, 74.32, and 69.77 mAh g−1 at current densities of 0.1, 0.2, 0.5, 1, 2 and 4 C (1 C = 117.6 mAh g−1). Even at a high current density of 5 C, the cell still delivers a capacity of 69.77 mAh g−1, higher than that of Na|NZSSP-Zn0|NVP/C cell (61.13 mAh g−1), respectively. Notably, when the current density was returned to 0.1 C, the discharge capacity recovered to 111.82 mAh g−1, demonstrating exceptional reversibility and stability. The voltage differential capacity (dQ/dV) curve based on 0.1 C was employed to analyze the charging and discharging behavior of the solid-state battery (Fig. 5d and e). The Na|NZSSP-Zn0.10|NVP/C cell displayed a charge/discharge plateau at 3.40/3.38 V, corresponding to the reversible redox of V3+/V4+ in Na3V2(PO4)3, indicating that the NZSSP-Zn0.10 electrolyte facilitates rapid Na+ reaction kinetics during the Na+ insertion/extraction process. In contrast, the voltage difference of the charge/discharge plateau corresponding to the dQ/dV curve of the Na|NZSSP-Zn0|NVP/C cell was 42.38 mV, higher than that of Na|NZSSP-Zn0.10|NVP/C cell (37.21 mV). This discrepancy stems from the low Na+ transport efficiency and inherent kinetic retardation in the NZSSP-Zn0 electrolyte, collectively limiting ion migration kinetics during redox reactions.
To further evaluate the long-term cycle stability of the Na|SSEs|NVP/C cell, galvanostatic cycling tests were conducted. As shown in Fig. 5f, the Na|NZSSP-Zn0.10|NVP/C cell delivered an initial discharge capacity of 108 mAh g−1 at 0.2 C, which corresponds to a high initial coulombic efficiency of 97.67%. Impressively, it maintained a reversible capacity of 106.61 mAh g−1 after 100 cycles, corresponding to 98.71% capacity retention relative. Even at a higher current density of 0.5 C, the Na|NZSSP-Zn0.10|NVP/C cell exhibited a high initial capacity of 105.72 mAh g−1 and demonstrated only 0.029% capacity decay per cycle over 400 cycles, accompanied by an average coulombic efficiency exceeding 99.99% (Fig. 5g and h). In contrast, the Na|NZSSP-Zn0|NVP/C cell exhibited a rapid capacity decay after 200 cycles, reactions at the electrode–electrolyte interface. These irreversible side reactions triggered progressive interfacial degradation, ultimately compromising the structural integrity of the electrolyte and accelerating performance deterioration. To comprehensively evaluate the practical viability of NZSSP-Zn0.10 SSE, Na|NZSSP-Zn0.10|NVP/C pouch cell with dimensions of 5.5 cm × 5.5 cm was assembled, and its performance was systematically characterized under mechanical stress (Fig. 5i and S10). The pouch cell consistently powered an LED light even when subjected to extreme deformations, including “flat”, “folded”, “rolled”, and “sheared”, demonstrating the excellent mechanical reliability and flexibility of the NZSSP-Zn0.10 SSE. The corresponding cycling performance was shown in Fig. S11. These results show its vast development potential and promising prospects in the field of flexible electronics.
The SI file contains additional materials characterization and electrochemical performance data (including Fig. S1–S11 and Tables S1–S4). See DOI: https://doi.org/10.1039/d5eb00075k.
Footnote |
† These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |