Next Article in Journal
Sustainable Composite Materials Based on Carnauba Wax and Montmorillonite Nanoclay for Energy Storage
Previous Article in Journal
Femtosecond Laser Fabrication of High-Linearity Liquid Metal-Based Flexible Strain Sensor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Chlorine-Rich Na6−xPS5−xCl1+x: A Promising Sodium Solid Electrolyte for All-Solid-State Sodium Batteries

1
College of Chemistry and Chemical Engineering, Huanggang Normal University, Huanggang 438000, China
2
School of Chemistry, Xi’an Jiaotong University, Xi’an 710049, China
*
Authors to whom correspondence should be addressed.
Materials 2024, 17(9), 1980; https://doi.org/10.3390/ma17091980
Submission received: 25 March 2024 / Revised: 20 April 2024 / Accepted: 22 April 2024 / Published: 24 April 2024

Abstract

:
Developing argyrodite-type, chlorine-rich, sodium-ion, solid-state electrolytes with high conductivity is a long-term challenge that is crucial for the advancement of all-solid-state batteries (ASSBs). In this study, chlorine-rich, argyrodite-type Na6−xPS5−xCl1+x solid solutions were successfully developed with a solid solution formation range of 0 ≤ x ≤ 0.5. Na5.5PS4.5Cl1.5 (x = 0.5), displaying a highest ionic conductivity of 1.2 × 10−3 S/cm at 25 °C, which is more than a hundred times higher than that of Na6PS5Cl. Cyclic voltammetry and electrochemical impedance spectroscopy results demonstrated that the rich chlorine significantly enhanced the ionic conductivity and electrochemical stability, in addition to causing a reduction in activation energy. The Na5.5PS4.5Cl1.5 composite also showed the characteristics of a pure ionic conductor without electronic conductivity. Finally, the viability of Na5.5PS4.5Cl1.5 as a sodium electrolyte for all-solid-state sodium batteries was checked in a lab-scale ASSB, showing stable battery performance. This study not only demonstrates new composites of sodium-ionic, solid-state electrolytes with relatively high conductivity but also provides an anion-modulation strategy to enhance the ionic conductivity of argyrodite-type sodium solid-state ionic conductors.

1. Introduction

Sodium-ion batteries (SIBs) have been considered a highly promising alternative to commercialized lithium-ion batteries (LIBs) due to the natural abundance of sodium resources (2.64 wt% for Na against 0.0017 wt% for Li) and their higher safety levels with similar charge/discharge mechanism [1]. Unfortunately, conventional SIBs face environmental and safety issues originating from the effumability and leakage of flammable organic liquid electrolytes [2,3]. Employing solid-state electrolytes (SSEs) instead of liquid organic electrolytes is a viable solution to the battery issues, since inorganic SSEs are non-volatile and non-flammable [4,5,6]. Therefore, the rational design of advanced sodium SSEs has been treated as a reliable way to achieve high-performance SIBs.
Sodium-based SSEs have been well developed in recent year and can be roughly divided into oxide-type and sulfide-type SSEs. Oxide-type SSEs such as β-alumina [7,8], NASICON–type Na1+xZr2SixP3−xO12 [9,10], and Na2M2TeO6 (M = Zn, or Mg) [11,12,13] exhibit good ionic conductivity of 10−4–10−3 S/cm. However, their poor interfacial compatibilities with electrodes lead to high contact resistances, which hinder the practical fabrication of all-solid-state sodium batteries at room temperature [14]. In contrast, sulfide materials are highly mechanically deformable and have excellent interfacial contact with active materials simply upon cold pressing [15,16]. The development of sodium-based sulfide conductors can be dates to 1992, when a homolog of a thio-LISICON-type, sulfide-based Na3PS4 sodium-ionic conductor with relatively high ionic conductivity was reported not long after the discovery of the well-known Li3PS4 lithium-ionic conductor in 1984 [17,18]. Since then, numerous studies have focused on the investigation of sulfide-based SSEs for SIBs [19,20,21]. Analogous ionic conductors such as Na3SbS4 have been synthesized and described in detail [22], while element-doping and defect-introducing strategies have been utilized to promote their electrochemical performance and air stability [2,15,23,24,25,26,27,28]. Moreover, inspired by the remarkable super-high ionic conductivity of thio-phosphate-type Li10GeP2S12 (LGPS, 12 mS/cm) sulfide electrolytes [29,30,31,32], a series of Na10MP2S12 (M = Ge, Si, Sn) sulfide compounds was predicted through first-principles calculations, and new Na-ion conductors, Na10+xSn1+xP2−xS12 (x = 0, 1) with tetragonal phases, have been successfully synthesized, possessing high ionic conductivity at ambient temperature [33,34,35,36]. However, Sn4+ in the SSE is potentially active when equipped with a highly reductive sodium metal anode, which might result in interfacial instability [37]. Therefore, a highly stable sodium SSE with high ionic conductivity is urgently needed.
Argyrodite-type conductors with high ionic conductivity and stable interfacial properties have been considered as promising electrolytes for all-solid-state batteries due to the absence of transition-metal species [38,39]. The Li6PS5X lithium-based conductor (X = Cl, Br, and I) showed their promising properties as an SSE, whereas few reports are available on its sodium analogs. Nevertheless, in previous reports, an argyrodite-type conductor, Na6PS5Cl, was synthesized [40]. Unexpectedly, the ion conductivity of Na6PS5Cl was only 1.2 × 10−5 S/cm, which was significantly lower than that of its Li6PS5Cl analog (3 × 10−3 S/cm) [41]. The electrochemical window of Na6PS5Cl is only 1.3 V, which makes it impossible for practical use. Increasing the degree of structure disorder by raising the halide content of electrolytes can strengthen the diffusivity of ions in the SSE [41,42]. Therefore, increasing the Cl and Na vacancy contents in Na6PS5Cl is favorable for the promotion of Cl/S2− disordering, as well as Na-ion mobility.
In this work, a highly conductive chlorine-rich sodium solid electrolyte, Na5.5PS4.5Cl1.5, was successfully synthesized by increasing Cl and Na vacancy contents with an outstanding conductivity of 1.2 × 10−3 S/cm at room temperature, which is almost ten-fold greater than that of pristine Na6PS5Cl. The introduction of Cl and Na vacancy contents in Na6PS5Cl induced a significant disordering of S2−/Cl arrangement, leading to the lowering of the activation barriers. Moreover, the electrochemical window of the Cl-rich SSE increased to 2 V, which makes the SSE a practical candidate for ASSBs. We demonstrated the availability of the chlorine-rich Na5.5PS4.5Cl1.5 SSE by assembling a Na3V2(PO4)3//Na5.5PS4.5Cl1.5//Na full battery, indicating the practical application value of the new Na5.5PS4.5Cl1.5 sulfide electrolyte for ASSBs.

2. Materials and Methods

2.1. Materials Synthesis

Na6−xPS5−xCl1+x (x = 0, 0.1, 0.2, 0.3, 0.4, 0.5, and 0.6) was synthesized by a high-temperature, solid-state reaction method. Stoichiometric Na2S, P2S5, and NaCl (all chemicals with a purity of 99.99%; Macklin Biochemical Co., Ltd., Shanghai, China) were weighed by chemical stoichiometry in an Ar-filled glovebox (<1 ppm of O2, H2O) and sealed in a 45 mL ZrO2 pot with 15 ZrO2 balls with 10 mm diameters and 40 ZrO2 balls with 5 mm diameters. The raw materials were ball-milled with a planetary ball mill apparatus (YXQM-0.45L, Changsha Mitrcn Instrument Equipment Co., Ltd., Changsha, China) at a rotational speed of 400 rpm for 8 h. Then, the obtained precursors were pelleted at 20–30 MPa in an Ar-filled glovebox (<1 ppm of O2, H2O), sealed in an evacuated Pyrex tube at pressures below 2 × 10−3 Pa, and sintered at 450 °C for 12 h with a heating rate of 2 °C/min. After naturally cooling to room temperature, the as-synthesized Na6−xPS5−xCl1+x was transferred to a glovebox and mixed by hand with mortar and pestle for 15 min for further characterization and battery assembly.

2.2. Methods

Powder XRD measurements were conducted using a diffractometer equipped with CuKα1 radiation (XRD-6010, Shimadzu, Tokyo, Japan) to ascertain the phase compositions of the synthesized powders. Prior to XRD measurements, samples were prepared in a glove box and sealed with polyimide film (Dongguan Meixin Co., Ltd., Dongguan, China) to shield them from moisture during the measurements. The diffraction data were collected in the 2θ range of 15° to 50° with step widths of 0.01°. The morphology and elemental mapping analyses for the composite powders were obtained using energy-dispersive X-ray spectroscopy (EDS, Bruker QUANTAX, Berlin, Germany), and the obtained samples were examined using a scanning electron microscope (SEM) (JEOL Ltd., JSM-6390, Tokyo, Japan).
The ionic conductivities of the cylindrical solid electrolyte (SE) pellets were assessed using the AC impedance technique with a cell (stainless steel/SE/stainless steel), employing pressurizable sealing molds and measured between 25 and 85 °C. This measurement was repeated two or three times, applying 15 mV in the frequency range of 7 MHz to 1 Hz and utilizing a potentiostat electrochemical interface (Bio-Logic SAS, VSP-300, Claix, France). SE pellets with a diameter of 10 mm were fabricated under a pressure of 303 MPa using a polyaryletherketone mold. Electrochemical window measurements were conducted using sodium/solid electrolyte/sodium cells. Typically, 100 mg of SE was pressed at 150 MPa to form a solid pellet. Na metal foil (purity: 99.9%, thickness: 0.1 mm, diameter: 5 mm; 99.9%, MTI Corporation, Shenzhen, China) was placed on the top side of the electrolyte pellet and further pressed at 150 MPa. The electrochemical stability of Na6−xPS5−xCl1+x series was determined over a wider potential window, and a cyclic voltammetry (CV) test was conducted at a scan rate of 0.1 mV/s between −2.5 and 2.5 V for comparison of the voltage stability region. The stack pressure applied during the tests was about 5 MPa.
An all-solid-state battery employing the obtained chlorine-rich sodium-ionic conductor with the highest ionic conduction as the electrolyte was assembled under a pressure of 6 MPa within an Ar-filled glovebox to evaluate its charge–discharge performance. The battery assembly method and the utilized cathode materials were akin to those documented in previous reports [43]. The details are as follows: Na metal foil served as the anode, while a composite mixture comprising Na3V2(PO4)3 (NVP) and Na5.5PS4.5Cl1.5 (NPSC-6) (NVP:NPSC-6 = 7:3) was employed as the cathode. Prior to use, a sieve with a 10 μm mesh was utilized to eliminate large-particle-size powder in Na5.5PS4.5Cl1.5. A pellet with a diameter of 10 mm was produced by cold pressing approximately 100 mg of Na5.5PS4.5Cl1.5 powder. Subsequently, Na foil (purity: 99.9%; thickness: 0.1 mm; diameter: 5 mm; 99.9%; MTI Corporation, Shenzhen, China) with a Cu mesh (purity: >99.9%; thickness: 0.045 mm; diameter: 8 mm; pore size: 0.4 mm × 1.5 mm; MTI Corporation, Hefei, China) was successively pressed onto one side of the pellet. On the opposite side of the pellet, the cathode composite (8 mg) with Al mesh (purity > 99.9%; thickness: 0.055 mm; diameter: 10 mm; pore size; 0.4 mm × 1.5 mm; MTI Corporation, Hefei, China) and an Al foil current collector (thickness: 0.1115 mm; diameter: 10 mm; purity: >99.9%, MTI Corporation, Hefei, China) were then pressed. Charge–discharge measurements were conducted at 25 °C, between 1.4 and 1.8 V, and with a current density of 0.03 mA cm−2 (0.05 C) using the LANHE CT2001A charge–discharge system (Wuhan LAND Electronics Co., Wuhan, China). The stack pressure applied during the tests was about 5 MPa.

3. Results and Discussion

3.1. Physical Characterization of Na6−xPS5−xCl1+x

Na6−xPS5−xCl1+x (x = 0, 0.1, 0.2, 0.3, 0.4, 0.5, or 0.6, denoted as NPSC-0, NPSC-1, NPSC-2, NPSC-3, NPSC-4, NPSC-5, and NPSC-6, respectively) samples were prepared by a solid-state reaction. The XRD patterns of the NPSC-0 sample were similar to those of the related argyrodite-type Cu6PS5Br, which displayed a monoclinic phase with a Cc space group [44,45]. A structural model of atomic Na6PS5Cl was simulated by replacement of Cu and Br in Cu6PS5Br with Na and Cl through structural refinement using Vesta software (version 3) [46], as displayed in the inset of Figure 1a. Two different Na+ sites were tetrahedrally coordinated. The Na1 site was surrounded by four S atoms, while the Na2 site was encircled by three S atoms and one Cl atom. The obtained NPSC-0 product was a solid block with chartreuse color (Figure 1b), while the outer and inner layers turned yellow in the NPSC-5 sample (Figure 1c,d). Comparisons of XRD patterns with those of other related compounds (such as tetragonal Na3PS4, cubic Na3PS4, and NaCl) are displayed in Figure S1, proving that no Na3PS4 or NaCl impurities were included in the NPSC-0 sample. With the increase in x in Na6−xPS5−xCl1+x, the (400) peak gradually shifted to lower 2θ angles, while the ( 2 - 22) peak migrated to higher 2θ angles, indicating the formation of solid solutions in Na6−xPS5−xCl1+x. Further introduction of additional chlorine into the Na6−xPS5−xCl1+x structure caused exsolvation of NaCl impurity at x = 0.6 (Figure S2), resulting a decrease in ionic conductivity. Thus, under the present experimental synthesis conditions, it was shown that Na5.5PS4.5Cl1.5 (x = 0.5) is the limit of solid solution formation in the Na6−xPS5−xCl1+x system. In other words, Na6−xPS5−xCl1+x formed solid solutions in the range of 0 ≤ x ≤ 0.5.
The morphology analysis results showed that the particle sizes of the NPSC-0, NPSC-1, NPSC-2, NPSC-3, NPSC-4 and NPSC-5 samples were in the range of 5 to 50 μm, which was gradually reduced with the increase in the Cl/S2− ratio, as shown in Figure 2a and Figure S3. The elemental mapping analyses shown in Figure 2b–f indicate the presence and uniform distribution of Na, P, S, and Cl elements within the NPSC-5 sample.

3.2. Electrochemical Performance of Na6−xPS5−xCl1+x Electrolyte

The electrochemical stability of Na6−xPS5−xCl1+x was determined for comparison of the voltage stability region, the results of which are shown in Figure 3. The NPSC-0 sample showed a small stable region only at 1.33 V (Figure 3a). With the increase in x in Na6−xPS5−xCl1+x, the voltage stability region widened gradually. When x = 0.1, the stable region was at 1.52 V; when x = 0.2, the stable region shifted to 1.68 V. It kept increasing to 1.79 and 1.91 V in the NPSC-3 and NPSC-4 samples, respectively. The NPSC-5 sample exhibited a maximum stable voltage of 2.02 V, indicating that promoting the Cl content in Na6−xPS5−xCl1+x favors the widening of the operating potential window.
Figure 4 illustrates the complex impedance spectra of the uniaxially cold, isostatic pressed powder samples obtained in the study. The impedance plots exhibit a semicircle pattern accompanied by a spike observed at high and low frequencies, respectively [43]. These features correspond to the total resistance (combining bulk and grain boundary contributions) and electrode resistances [47]. The semicircle represents ionic transport within the materials, while the spike is indicative of ion-blocking behavior at the interfaces between the electrodes and electrolyte. The equivalent circuits used for impedance analysis, as depicted in Figure S4, were derived from the ZView program [47]. Additionally, the semicircles observed at higher frequencies, displaying capacitances of 10−10 F, suggest total resistance, with a combined contribution from both bulk and grain boundaries [48].
The ionic conductivity of Na6−xPS5−xCl1+x as analytical material was characterized using the AC impedance method. All the impedance spectra contained formed semicircle pattern at high frequencies attributed to the charge-transfer resistance and an inclined line at low frequencies associated with diffusion processes in the electrodes. With the increase in x in Na6−xPS5−xCl1+x (0 ≤ x ≤ 0.5), the semicircle diminished gradually, leading to a reduction in resistances by introducing rich Cl contents and facilitating Na+ diffusion in Na6−xPS5−xCl1+x (Figure 4a). The calculated ionic conductivities rose gradually with increasing content of Cl, namely 9.4 × 10−5 S/cm for NPSC-0, 1.4 × 10−4 S/cm for NPSC-1, 2.1 × 10−4 S/cm for NPSC-2, 4.2 × 10−4 S/cm for NPSC-3, 7.9 × 10−4 S/cm for NPSC-4, and 1.2 × 10−3 S/cm for NPSC-5. The ionic conductivity of the NPSC-5 sample (Na5.5PS4.5Cl1.5 with x = 0.5) showed the highest ionic conduction, which was over a hundred times higher than that of Na6PS5Cl (1.0 × 10−5 S/cm [40]), proving the importance of increasing the Cl content for Na+ migration. The NPSC-06 sample displayed a lower ionic conductivity value (2.9 × 10−4 S/cm) than that of NPSC-5 (Figure S4), which could be attributed to the low ionic conductivity of NaCl impurity (Figure S2). The whole testing and disassembly process of the steel/NPSC-5/steel configuration was completed in one shot, which can be observed in Video S1.
All the obtained samples exhibited linear relationships with ionic conductivity in the range of 25–85 °C. Figure 5a illustrates the temperature dependence of ionic conductivity. All obtained samples exhibited linear relationships between ionic conductivity and reciprocal temperature within the studied temperature range, indicating adherence to the Arrhenius law. By applying the Arrhenius equation (σT = σ0exp(−Ea/kBT), where σ0 references a pre-exponential factor) [49], the activation energy (Ea) of Na6−xPS5−xCl1+x (x = 0, 0.1, 0.2, 0.3, 0.4, and 0.5) for total conduction was determined from the slope of the plots. In Figure 5b, the composition dependence of the samples’ activation energies (Ea) are depicted. AC impedance analysis results revealed that the ionic conductivity of all chlorine-rich Na6−xPS5−xCl1+x samples was higher than that of the primordial Na6PS5Cl (x = 0) sample, with ionic conductivity increasing from x = 0 to 0.5. These findings indicate that a rise in chlorine contents can enhance the ionic conductivity of Na6PS5Cl with a decrease in Ea. Table S1 summarizes the ionic conductivities and Ea values reported in the literature, as well as the values obtained in this study. Compared with the reported Na6PS5Cl (0.01 × 10−3 S/cm) [40], the modified Na5.5PS4.5Cl1.5 exhibited superior ionic conductivity of 1.2 × 10−3 S/cm, which is the same order of magnitude as the Na11Sn2PS12 sulfide electrolyte (4 × 10−3 S/cm) reported in [35] and the vacancy-contained Na3SbS4 (3 × 10−3 S/cm) reported in [22]. Therefore, Na5.5PS4.5Cl1.5 (x = 0.5, NPSC-5) shows very high values of ionic conductivity among the typical sodium ion conductors that have been reported.

3.3. Electrochemical Performance of NVP//NPSC-5//Na ASSB

Na5.5PS4.5Cl1.5 (x = 0.5, NPSC-5) showed the highest ionic conductivity and lowest activation energy in the Na6−xPS5−xCl1+x system, so the electrochemical performance of a laboratory-grade, pressurized NVP//NPSC-5//Na cell with NPSC-5 as a sodium solid electrolyte and commercial Na3V2(PO4)3 (NVP, 99.9%, MTI Corporation, Shenzhen, China) and Na metal foil as cathode and anode, respectively, was determined, as depicted in the schematic diagram in Figure 5a. Since NVP can offer a 1.6 V voltage plateau through the V3+/V2+ redox reaction, which lies within the voltage stability region of the NPSC-5 electrolyte, the galvanostatic charge–discharge of the NVP//NPSC-5//Na ASSB was tested in the potential range of 1.4–1.8 V at a current density of 0.05 C. All charge–discharge profiles exhibited a 1.6 V plateau, corresponding to the phase transition of Na3V2(PO4)3/Na4V2(PO4)3 with one Na+ intercalation/deintercalation from the host lattices [50]. The right figure in Figure 6a shows a physical electronic photograph of an assembled all-solid-state sodium-ion battery that can normally light up an LED light bulb at room temperature, indicating that NPSC-5 is a pure sodium-ionic conductor material with negligible electronic conductivity and functions as a solid electrolyte. The additional plateau at about 1.42 V shown in Figure 6b can be attributed to the formation of a stable NPSC-5/Na interface during the first cycle, which was similar to reports of other electrochemical investigations sodium ASSBs [2,51,52,53]. The NPSC-5/Na interface resulted in a low coulombic efficiency (58%) during the first cycle (The initial discharge capacity was about 81.9 mAh/g, whereas the charge capacity was only 47.3 mAh/g). The second charge and discharge capacities were 43.1 mAh/g and 42.2 mAh/g, respectively, indicating the rebound of coulombic efficiency (98%). The 10th cycle displayed charge and discharge capacities of 34.3 and 33.7 mAh/g, respectively; nearly 73% of the initial charge capacity was retained, and the coulombic efficiency was maintained at 98%, demonstrating that the synthesized material facilitated the cyclic charging and discharging capabilities of the battery. The energy density of the full cell was about 70 Wh/Kg based on the mass of the active cathode. Further modifications, such as to the tips for battery assembly, compositing the NVP cathode with graphene, or replacing the Na disc with a Na-Sn alloy to improve the performance of the NVP//NPSC-5//Na ASSB, should be analyzed and explored in depth in the future.

4. Conclusions

In summary, halide-rich solid solutions of argyrodite-type Na6−xPS5−xCl1+x were synthesized and formed solid solutions in the range of 0 ≤ x ≤ 0.5. Increasing Cl contents effectively reduced the activation barrier and increased the Na-ion mobility, as well as electrochemical stability. The optimal solid solution composition of Na5.5PS4.5Cl1.5 exhibited particularly high Na-ion conductivity in the Na6−xPS5−xCl1+x in this work, namely 1.2 mS/cm at 25 °C, which is more than a hundred times higher than that of Na6PS5Cl with a regular composition. Meanwhile, Na5.5PS4.5Cl1.5 is more suitable for high-voltage cathodes than Na6PS5Cl due to its wider voltage stability region up to 2 V, and the smooth charge–discharge cycling performance of Na3V2(PO4)3//Na5.5PS4.5Cl1.5//Na at room temperature verifies that the obtained material is a pure sodium-ion conductor and its potential as a sodium-ionic solid electrolyte for all-solid-state sodium batteries. To date, the reported sulfide electrolyte, Na11Sn2PS12 (4 × 10−3 S/cm) [35] and vacancy-contained Na3SbS4 (3 × 10−3 S/cm) [22] have exhibited excellent conductivity. Both references proposed the strategy of designing Na vacancy for the improvement of conductivity. Therefore, we suppose that the ionic conductivity of Na5.5PS4.5Cl1.5 could be further promoted by designing Na vacancy in the crystal structure.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma17091980/s1, Figure S1: Comparison of XRD patterns for Na6PS5Cl, cubic Ag15P4S16Cl, tetragonal Na3PS4, cubic Na3PS4, and NaCl; Figure S2: XRD patterns of NPSC-5 and NPSC-6 samples and standard-NaCl; Figure S3: SEM images of (a) NPSC-0, (b) NPSC-1, (c) NPSC-2, (d) NPSC-3, and (e) NPSC-4 samples; Figure S4: The equivalent circuit used to extract the value of resistance from EIS spectra with two semicircles. All EIS spectra of Na6−xPS5−xCl1+x, in general, exhibit two semicircular patterns. The conductivity was computed from the resistance using the following formula: Conductivity = Thickness/(Area × Resistance); Figure S5: Impedance plot of NPSC-6 sample at 25 °C; Table S1: Comparison of ionic conductivity and activation energy (Ea) obtained in this study with those previously reported; Video S1: Impedance test of NPSC-5 sample.

Author Contributions

Conceptualization, G.Z.; Data curation, Y.Z., H.Z. (Haoran Zheng), J.Y., A.J.K., L.G. and G.Z.; Funding acquisition, Y.Z., L.G. and G.Z.; Investigation, Y.Z., L.G. and G.Z.; Methodology, G.Z.; Project administration, G.Z.; Validation, Y.Z., H.Z. (Hongyang Zhao), L.G. and G.Z.; Visualization, Y.Z., H.Z. (Haoran Zheng), J.Y., H.Z. (Hongyang Zhao), A.J.K. and G.Z.; Writing—original draft, Y.Z. and G.Z.; Writing—review and editing, H.Z. (Hongyang Zhao) and G.Z. All authors have read and agreed to the published version of the manuscript.

Funding

The work performed at Huanggang Normal University was supported by financial support from the National Natural Science Foundation of China (22109049); the Provincial Natural Science Foundation for Distinguished Young Scholars of Hubei Province, China (2023AFA064); the Key Project of Scientific Research Program of Hubei Provincial Department of Education, China (D20222901); and the Key Laboratory of Catalysis and Energy Materials Chemistry of the Ministry of Education and Hubei Key Laboratory of Catalysis and Materials Science (202306).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Vaalma, C.; Buchholz, D.; Weil, M.; Passerini, S. A Cost and Resource Analysis of Sodium-Ion Batteries. Nat. Rev. Mater. 2018, 3, 18013. [Google Scholar] [CrossRef]
  2. Moon, C.K.; Lee, H.-J.; Park, K.H.; Kwak, H.; Heo, J.W.; Choi, K.; Yang, H.; Kim, M.-S.; Hong, S.-T.; Lee, J.H.; et al. Vacancy-Driven Na+ Superionic Conduction in New Ca-Doped Na3PS4 for All-Solid-State Na-Ion Batteries. ACS Energy Lett. 2018, 3, 2504–2512. [Google Scholar] [CrossRef]
  3. Zhang, Y.; Zhang, R.; Huang, Y. Air-Stable NaxTMO2 Cathodes for Sodium Storage. Front. Chem. 2019, 7, 335. [Google Scholar] [CrossRef] [PubMed]
  4. Lu, Y.; Li, L.; Zhang, Q.; Niu, Z.Q.; Chen, J. Electrolyte and Interface Engineering for Solid-State Sodium Batteries. Joule 2018, 2, 1747–1770. [Google Scholar] [CrossRef]
  5. Wang, J.; Okabe, J.; Komine, Y.; Notohara, H.; Urita, K.; Moriguchi, I.; Wei, M. The optimized interface engineering of VS2 as cathodes for high performance all-solid-state lithium-ion battery. Sci. China Technol. Sci. 2022, 65, 1859–1866. [Google Scholar] [CrossRef]
  6. Chong, P.; Zhou, Z.; Wang, K.; Zhai, W.; Li, Y.; Wang, J.; Wei, M. The Stabilizing of 1T-MoS2 for All-Solid-State Lithium-Ion Batteries. Batteries 2023, 9, 26. [Google Scholar] [CrossRef]
  7. Yu, X.; Yao, Y.; Wang, X.; Cen, S.; Li, D.; Ma, H.; Chen, J.; Wang, D.; Mao, Z.; Dong, C. Rapid ultrasound welding toward compact Na/Beta-Al2O3 interface realizing room-temperature solid-state sodium metal battery. Energy Storage Mater. 2023, 54, 221–226. [Google Scholar] [CrossRef]
  8. Lei, D.; He, Y.-B.; Huang, H.; Yuan, Y.; Zhong, G.; Zhao, Q.; Hao, X.; Zhang, D.; Lai, C.; Zhang, S.; et al. Cross-linked beta alumina nanowires with compact gel polymer electrolyte coating for ultra-stable sodium metal battery. Nat. Commun. 2019, 10, 4244. [Google Scholar] [CrossRef]
  9. Liu, L.; Su, J.; Zhou, X.; Liang, D.; Liu, Y.; Tang, R.; Xu, Y.; Jiang, Y.; Wei, Z. Conductivity versus structure dependence investigation of Na1+xZr2SixP3-xO12 (0 ≤ x ≤ 3) through composition optimization by adjusting Si/P ratio. Mater. Today Chem. 2023, 30, 101495. [Google Scholar] [CrossRef]
  10. Park, H.; Jung, K.; Nezafati, M.; Kim, C.-S.; Kang, B. Sodium Ion Diffusion in Nasicon (Na3Zr2Si2PO12) Solid Electrolytes: Effects of Excess Sodium. ACS Appl. Mater. Interfaces 2016, 8, 27814–27824. [Google Scholar] [CrossRef] [PubMed]
  11. Li, Y.; Deng, Z.; Peng, J.; Chen, E.; Yu, Y.; Li, X.; Luo, J.; Huang, Y.; Zhu, J.; Fang, C.; et al. A P2-Type Layered Superionic Conductor Ga-Doped Na2Zn2TeO6 for All-Solid-State Sodium-Ion Batteries. Chemistry 2018, 24, 1057–1061. [Google Scholar] [CrossRef] [PubMed]
  12. Li, Y.; Deng, Z.; Peng, J.; Gu, J.; Chen, E.; Yu, Y.; Wu, J.; Li, X.; Luo, J.; Huang, Y.; et al. New P2-Type Honeycomb-Layered Sodium-Ion Conductor: Na2Mg2TeO6. ACS Appl. Mater. Interfaces 2018, 10, 15760–15766. [Google Scholar] [CrossRef] [PubMed]
  13. Huang, H.; Yang, Y.; Chi, C.; Wu, H.-H.; Huang, B. Phase stability and fast ion transport in P2-type layered Na2X2TeO6 (X = Mg, Zn) solid electrolytes for sodium batteries. J. Mater. Chem. A 2020, 8, 22816–22827. [Google Scholar] [CrossRef]
  14. Dong, Y.F.; Wen, P.C.; Shi, H.D.; Yu, Y.; Wu, Z.S. Solid-State Electrolytes for Sodium Metal Batteries: Recent Status and Future Opportunities. Adv. Funct. Mater. 2023, 21, 2213584. [Google Scholar] [CrossRef]
  15. Hayashi, A.; Masuzawa, N.; Yubuchi, S.; Tsuji, F.; Hotehama, C.; Sakuda, A.; Tatsumisago, M. A sodium-ion sulfide solid electrolyte with unprecedented conductivity at room temperature. Nat. Commun. 2019, 10, 5266. [Google Scholar] [CrossRef] [PubMed]
  16. Li, F.; Hou, M.; Zhao, L.; Zhang, D.; Yang, B.; Liang, F. Electrolyte and interface engineering for solid-state sodium batteries. Energy Storage Mater. 2024, 65, 103181. [Google Scholar] [CrossRef]
  17. Tachez, M.; Malugani, J.-P.; Mercier, R.; Robert, G. Ionic conductivity of and phase transition in lithium thiophosphate Li3PS4. Solid State Ionics 1984, 14, 181–185. [Google Scholar] [CrossRef]
  18. Jansen, M.; Henseler, U. Synthesis, Structure Determination, and Ionic-Conductivity of Sodium Tetrathiophosphate. J. Solid State Chem. 1992, 99, 110–119. [Google Scholar] [CrossRef]
  19. Wu, E.A.; Kompella, C.S.; Zhu, Z.; Lee, J.Z.; Lee, S.C.; Chu, I.-H.; Nguyen, H.; Ong, S.P.; Banerjee, A.; Meng, Y.S. New Insights into the Interphase between the Na Metal Anode and Sulfide Solid-State Electrolytes: A Joint Experimental and Computational Study. ACS Appl. Mater. Interfaces 2018, 10, 10076–10086. [Google Scholar] [CrossRef]
  20. Ma, Q.; Tietz, F. Solid-State Electrolyte Materials for Sodium Batteries: Towards Practical Applications. ChemElectroChem 2020, 7, 2693–2713. [Google Scholar] [CrossRef]
  21. Li, Y.; Arnold, W.; Halacoglu, S.; Jasinski, J.B.; Druffel, T.; Wang, H. Phase-Transition Interlayer Enables High-Performance Solid-State Sodium Batteries with Sulfide Solid Electrolyte. Adv. Funct. Mater. 2021, 31, 2101636. [Google Scholar] [CrossRef]
  22. Zhang, L.; Zhang, D.; Yang, K.; Yan, X.; Wang, L.; Mi, J.; Xu, B.; Li, Y. Vacancy-Contained Tetragonal Na3SbS4 Superionic Conductor. Adv. Sci. 2016, 3, 1600089. [Google Scholar] [CrossRef] [PubMed]
  23. Feng, X.; Chien, P.-H.; Zhu, Z.; Chu, I.-H.; Wang, P.; Immediato-Scuotto, M.; Arabzadeh, H.; Ong, S.P.; Hu, Y.-Y. Studies of Functional Defects for Fast Na-Ion Conduction in Na3−yPS4−xClx with a Combined Experimental and Computational Approach. Adv. Funct. Mater. 2019, 29, 1807951. [Google Scholar] [CrossRef]
  24. Tsuji, F.; Masuzawa, N.; Sakuda, A.; Tatsumisago, M.; Hayashi, A. Preparation and Characterization of Cation-Substituted Na3SbS4 Solid Electrolytes. ACS Appl. Energy Mater. 2020, 3, 11706–11712. [Google Scholar] [CrossRef]
  25. Yu, L.; Jiao, Q.; Liang, B.; Shan, H.; Lin, C.; Gao, C.; Shen, X.; Dai, S. Exceptionally high sodium ion conductivity and enhanced air stability in Na3SbS4 via germanium doping. J. Alloys Compd. 2022, 913, 165229. [Google Scholar] [CrossRef]
  26. Jalem, R.; Gao, B.; Tian, H.-K.; Tateyama, Y. Theoretical study on stability and ion transport property with halide doping of Na3SbS4 electrolyte for all-solid-state batteries. J. Mater. Chem. A 2022, 10, 2235–2248. [Google Scholar] [CrossRef]
  27. Yu, L.; Yin, J.; Gao, C.; Lin, C.; Shen, X.; Dai, S.; Jiao, Q. Halogen Doping Mechanism and Interface Strengthening in the Na3SbS4 Electrolyte via Solid-State Synthesis. ACS Appl. Mater. Interfaces 2023, 15, 31635–31642. [Google Scholar] [CrossRef] [PubMed]
  28. Fu, Y.; Gong, Z.; Li, D.; Liu, Y.; Zhou, X.; Yang, Y.; Jiao, Q. Zn doping for enhanced sodium-ion conductivity and air stability in Na3SbS4 solid electrolyte. J. Mater. Sci. 2024, 59, 3009–3017. [Google Scholar] [CrossRef]
  29. Kamaya, N.; Homma, K.; Yamakawa, Y.; Hirayama, M.; Kanno, R.; Yonemura, M.; Kamiyama, T.; Kato, Y.; Hama, S.; Kawamoto, K.; et al. A lithium superionic conductor. Nat. Mater. 2011, 10, 682–686. [Google Scholar] [CrossRef] [PubMed]
  30. Weber, D.A.; Senyshyn, A.; Weldert, K.S.; Wenzel, S.; Zhang, W.; Kaiser, R.; Berendts, S.; Janek, J.; Zeier, W.G. Structural Insights and 3D Diffusion Pathways within the Lithium Superionic Conductor Li10GeP2S12. Chem. Mater. 2016, 28, 5905–5915. [Google Scholar] [CrossRef]
  31. Dawson, J.A.; Islam, M.S. A Nanoscale Design Approach for Enhancing the Li-Ion Conductivity of the Li10GeP2S12 Solid Electrolyte. ACS Mater. Lett. 2022, 4, 424–431. [Google Scholar] [CrossRef] [PubMed]
  32. Kato, Y.; Hori, S.; Kanno, R. Li10GeP2S12-Type Superionic Conductors: Synthesis, Structure, and Ionic Transportation. Adv. Energy Mater. 2020, 10, 2002153. [Google Scholar] [CrossRef]
  33. Richards, W.D.; Tsujimura, T.; Miara, L.J.; Wang, Y.; Kim, J.C.; Ong, S.P.; Uechi, I.; Suzuki, N.; Ceder, G. Design and synthesis of the superionic conductor Na10SnP2S12. Nat. Commun. 2016, 7, 11009. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, Z.; Ramos, E.; Lalère, F.; Assoud, A.; Kaup, K.; Hartman, P.; Nazar, L.F. Na11Sn2PS12: A new solid state sodium superionic conductor. Energy Environ. Sci. 2018, 11, 87–93. [Google Scholar] [CrossRef]
  35. Duchardt, M.; Ruschewitz, U.; Adams, S.; Dehnen, S.; Roling, B. Vacancy-Controlled Na+ Superion Conduction in Na11Sn2PS12. Angew. Chem. Int. Ed. 2018, 57, 1351–1355. [Google Scholar] [CrossRef] [PubMed]
  36. Kandagal, V.S.; Bharadwaj, M.D.; Waghmare, U.V. Theoretical prediction of a highly conducting solid electrolyte for sodium batteries: Na10GeP2S12. J. Mater. Chem. A 2015, 3, 12992–12999. [Google Scholar] [CrossRef]
  37. Nunotani, N.; Ohsaka, T.; Tamura, S.; Imanaka, N. Tetravalent Sn4+ Ion Conductor Based on NASICON-Type Phosphate. ECS Electrochem. Lett. 2012, 1, A66. [Google Scholar] [CrossRef]
  38. Boulineau, S.; Courty, M.; Tarascon, J.-M.; Viallet, V. Mechanochemical synthesis of Li-argyrodite Li6PS5X (X=Cl, Br, I) as sulfur-based solid electrolytes for all solid state batteries application. Solid State Ionics 2012, 221, 1–5. [Google Scholar] [CrossRef]
  39. Deiseroth, H.J.; Kong, S.T.; Eckert, H.; Vannahme, J.; Reiner, C.; Zaiss, T.; Schlosser, M. Li6PS5X: A class of crystalline Li-rich solids with an unusually high Li+ mobility. Angew. Chem. Int. Ed. 2008, 47, 755–758. [Google Scholar] [CrossRef] [PubMed]
  40. Studenyak, I.P.; Pogodin, A.I.; Studenyak, V.I.; Kokhan, O.P.; Azhniuk, Y.M.; Solonenko, D.; Daróczi, L.; Kökényesi, S.; Zahn, D.R.T. Structural, electrical and optical properties of ion-conducting Na6PS5Cl, Na6PS5Br, and Na7PS6 compounds. J. Phys. Chem. Solids 2021, 159, 110269. [Google Scholar] [CrossRef]
  41. Adeli, P.; Bazak, J.D.; Park, K.H.; Kochetkov, I.; Huq, A.; Goward, G.R.; Nazar, L.F. Boosting Solid-State Diffusivity and Conductivity in Lithium Superionic Argyrodites by Halide Substitution. Angew. Chem. Int. Ed. 2019, 58, 8681–8686. [Google Scholar] [CrossRef] [PubMed]
  42. Stamminger, A.R.; Ziebarth, B.; Mrovec, M.; Hammerschmidt, T.; Drautz, R. Ionic Conductivity and Its Dependence on Structural Disorder in Halogenated Argyrodites Li6PS5X (X = Br, Cl, I). Chem. Mater. 2019, 31, 8673–8678. [Google Scholar] [CrossRef]
  43. Gao, L.; Xie, Y.; Tong, Y.; Xu, M.; You, J.; Wei, H.; Yu, X.; Xu, S.; Zhang, Y.; Che, Y.; et al. Lithium-site substituted argyrodite-type Li6PS5I solid electrolytes with enhanced ionic conduction for all-solid-state batteries. Sci. China Technol. Sci. 2023, 66, 2059–2068. [Google Scholar] [CrossRef]
  44. Haznar, A.; Pietraszko, A.; Studenyak, I.P. X-ray study of the superionic phase transition in Cu6PS5Br. Solid State Ionics 1999, 119, 31–36. [Google Scholar] [CrossRef]
  45. Nilges, T.; Pfitzner, A. A structural differentiation of quaternary copper argyrodites: Structure—Property relations of high temperature ion conductors. Z. Kristallogr. Cryst. Mater. 2005, 220, 281–294. [Google Scholar] [CrossRef]
  46. Momma, K.; Izumi, F. VESTA 3 for three-dimensional visualization of crystal, volumetric and morphology data. J. Appl. Crystallogr. 2011, 44, 1272–1276. [Google Scholar] [CrossRef]
  47. Zhao, G.; Suzuki, K.; Hirayama, M.; Kanno, R. Syntheses and Characterization of Novel Perovskite-Type LaScO3-Based Lithium Ionic Conductors. Molecules 2021, 26, 299. [Google Scholar] [CrossRef] [PubMed]
  48. Zhao, G.; Suzuki, K.; Hirayama, M.; Kanno, R. Synthesis and Lithium-ion Conductivity of Sr(La1−xLi3x)ScO4 with a K2NiF4 Structure. Electrochemistry 2022, 90, 017005. [Google Scholar] [CrossRef]
  49. Zhao, G.; Suzuki, K.; Seki, T.; Sun, X.; Hirayama, M.; Kanno, R. High lithium ionic conductivity of γ-Li3PO4-type solid electrolytes in Li4GeO4–Li4SiO4–Li3VO4 quasi-ternary system. J. Solid State Chem. 2020, 292, 121651. [Google Scholar] [CrossRef]
  50. Jian, Z.; Sun, Y.; Ji, X. A new low-voltage plateau of Na3V2(PO4)3 as an anode for Na-ion batteries. Chem. Commun. 2015, 51, 6381–6383. [Google Scholar] [CrossRef] [PubMed]
  51. Chu, I.-H.; Kompella, C.S.; Nguyen, H.; Zhu, Z.; Hy, S.; Deng, Z.; Meng, Y.S.; Ong, S.P. Room-Temperature All-solid-state Rechargeable Sodium-ion Batteries with a Cl-doped Na3PS4 Superionic Conductor. Sci. Rep. 2016, 6, 33733. [Google Scholar] [CrossRef] [PubMed]
  52. Wan, H.; Mwizerwa, J.P.; Qi, X.; Xu, X.; Li, H.; Zhang, Q.; Cai, L.; Hu, Y.-S.; Yao, X. Nanoscaled Na3PS4 Solid Electrolyte for All-Solid-State FeS2/Na Batteries with Ultrahigh Initial Coulombic Efficiency of 95% and Excellent Cyclic Performances. ACS Appl. Mater. Interfaces 2018, 10, 12300–12304. [Google Scholar] [CrossRef] [PubMed]
  53. Nguyen, H.; Banerjee, A.; Wang, X.; Tan, D.; Wu, E.A.; Doux, J.-M.; Stephens, R.; Verbist, G.; Meng, Y.S. Single-step synthesis of highly conductive Na3PS4 solid electrolyte for sodium all solid-state batteries. J. Power Sources 2019, 435, 126623. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of the Na6PS5Cl sample and argyrodite-Cu6PS5Br (inset represents the crystal structure model of Cu6PS5Br). (b,c) Digital images of the synthesized Na6−xPS5−xCl1+x (x = 0 and 0.5) solid block. (d) Comparison of inner color of Na6−xPS5−xCl1+x (x = 0 and 0.6 samples. (e) XRD patterns of Na6−xPS5−xCl1+x (x = 0, 0.1, 0.2, 0.3, 0.4, and 0.5, denoted as NPSC-0, NPSC-1, NPSC-2, NPSC-3, NPSC-4, and NPSC-5, respectively). (f) Enlarged XRD patterns in the 2θ range of 28–32° from the dotted boxes in (e), the arrows clearly show regular shifting in one direction with increasing Cl content, indicating the formation of solid solutions.
Figure 1. (a) XRD patterns of the Na6PS5Cl sample and argyrodite-Cu6PS5Br (inset represents the crystal structure model of Cu6PS5Br). (b,c) Digital images of the synthesized Na6−xPS5−xCl1+x (x = 0 and 0.5) solid block. (d) Comparison of inner color of Na6−xPS5−xCl1+x (x = 0 and 0.6 samples. (e) XRD patterns of Na6−xPS5−xCl1+x (x = 0, 0.1, 0.2, 0.3, 0.4, and 0.5, denoted as NPSC-0, NPSC-1, NPSC-2, NPSC-3, NPSC-4, and NPSC-5, respectively). (f) Enlarged XRD patterns in the 2θ range of 28–32° from the dotted boxes in (e), the arrows clearly show regular shifting in one direction with increasing Cl content, indicating the formation of solid solutions.
Materials 17 01980 g001
Figure 2. (a) SEM image of NPSC-5 sample. (bf) EDS element mapping of Na, P, S, and Cl in NPSC-5 sample.
Figure 2. (a) SEM image of NPSC-5 sample. (bf) EDS element mapping of Na, P, S, and Cl in NPSC-5 sample.
Materials 17 01980 g002
Figure 3. CV curves of (a) Na/NPSC-0/Na, (b) Na/NPSC-1/Na, (c) Na/NPSC-2/Na, (d) Na/NPSC-3/Na. (e) Na/NPSC-4/Na, and (f) Na/NPSC-5/Na with Na6−xPS5−xCl1+x as analytical materials (x = 0, 0.1, 0.2, 0.3, 0.4, and 0.5, denoted as NPSC-0, NPSC-1, NPSC-2, NPSC-3, NPSC-4, and NPSC-5, respectively) and a scan rate of 0.1 mV/s.
Figure 3. CV curves of (a) Na/NPSC-0/Na, (b) Na/NPSC-1/Na, (c) Na/NPSC-2/Na, (d) Na/NPSC-3/Na. (e) Na/NPSC-4/Na, and (f) Na/NPSC-5/Na with Na6−xPS5−xCl1+x as analytical materials (x = 0, 0.1, 0.2, 0.3, 0.4, and 0.5, denoted as NPSC-0, NPSC-1, NPSC-2, NPSC-3, NPSC-4, and NPSC-5, respectively) and a scan rate of 0.1 mV/s.
Materials 17 01980 g003
Figure 4. (a) Impedance plots of Na6−xPS5−xCl1+x samples (x = 0, 0.1, 0.2, 0.3, 0.4, and 0.5, denoted as NPSC-0, NPSC-1, NPSC-2, NPSC-3, NPSC-4, and NPSC-5, respectively) obtained at 25 °C. (b) Impedance plots of the complex impedance responses of (b) NPSC-0, (c) NPSC-1, (d) NPSC-2, (e) NPSC-3, (f) NPSC-4, and (g) NPSC-5 samples measured at 25, 45, 65, and 85 °C, respectively.
Figure 4. (a) Impedance plots of Na6−xPS5−xCl1+x samples (x = 0, 0.1, 0.2, 0.3, 0.4, and 0.5, denoted as NPSC-0, NPSC-1, NPSC-2, NPSC-3, NPSC-4, and NPSC-5, respectively) obtained at 25 °C. (b) Impedance plots of the complex impedance responses of (b) NPSC-0, (c) NPSC-1, (d) NPSC-2, (e) NPSC-3, (f) NPSC-4, and (g) NPSC-5 samples measured at 25, 45, 65, and 85 °C, respectively.
Materials 17 01980 g004
Figure 5. (a) Temperature dependence of the ionic conductivity of XRD patterns of Na6−xPS5−xCl1+x (x = 0, 0.1, 0.2, 0.3, 0.4, and 0.5, denoted as NPSC-0, NPSC-1, NPSC-2, NPSC-3, NPSC-4, and NPSC-5, respectively) samples. (b) Activation energies of Na6−xPS5−xCl1+x at different x values.
Figure 5. (a) Temperature dependence of the ionic conductivity of XRD patterns of Na6−xPS5−xCl1+x (x = 0, 0.1, 0.2, 0.3, 0.4, and 0.5, denoted as NPSC-0, NPSC-1, NPSC-2, NPSC-3, NPSC-4, and NPSC-5, respectively) samples. (b) Activation energies of Na6−xPS5−xCl1+x at different x values.
Materials 17 01980 g005
Figure 6. (a) Illustration of a Na3V2(PO4)3//NPSC-5//Na ASSB. (b) A digital picture of a Na3V2(PO4)3//NPSC-5//Na ASSB working at room temperature to make an LED diode bead light up. (c) Galvanostatic charge–discharge curves of a Na3V2(PO4)3//NPSC-5//Na ASSB at 25 °C with a charge–discharge speed of 0.05 C (1 C = 60 mA/g).
Figure 6. (a) Illustration of a Na3V2(PO4)3//NPSC-5//Na ASSB. (b) A digital picture of a Na3V2(PO4)3//NPSC-5//Na ASSB working at room temperature to make an LED diode bead light up. (c) Galvanostatic charge–discharge curves of a Na3V2(PO4)3//NPSC-5//Na ASSB at 25 °C with a charge–discharge speed of 0.05 C (1 C = 60 mA/g).
Materials 17 01980 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, Y.; Zheng, H.; You, J.; Zhao, H.; Khan, A.J.; Gao, L.; Zhao, G. Chlorine-Rich Na6−xPS5−xCl1+x: A Promising Sodium Solid Electrolyte for All-Solid-State Sodium Batteries. Materials 2024, 17, 1980. https://doi.org/10.3390/ma17091980

AMA Style

Zhang Y, Zheng H, You J, Zhao H, Khan AJ, Gao L, Zhao G. Chlorine-Rich Na6−xPS5−xCl1+x: A Promising Sodium Solid Electrolyte for All-Solid-State Sodium Batteries. Materials. 2024; 17(9):1980. https://doi.org/10.3390/ma17091980

Chicago/Turabian Style

Zhang, Yi, Haoran Zheng, Jiale You, Hongyang Zhao, Abdul Jabbar Khan, Ling Gao, and Guowei Zhao. 2024. "Chlorine-Rich Na6−xPS5−xCl1+x: A Promising Sodium Solid Electrolyte for All-Solid-State Sodium Batteries" Materials 17, no. 9: 1980. https://doi.org/10.3390/ma17091980

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop