Next Article in Journal
Collaborative Optimization Scheduling of Resilience and Economic Oriented Islanded Integrated Energy System under Low Carbon Transition
Previous Article in Journal
Study on the Spatiotemporal Evolution of the Ecological Landscape and Construction of an Ecological Network: A Case Study of Hebei Province
Previous Article in Special Issue
Synthesis and Electrochemical Properties of Co3O4@Reduced Graphene Oxides Derived from MOF as Anodes for Lithium-Ion Battery Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Characterization of Na3SbS4 Solid Electrolytes via Mechanochemical and Sintered Solid-State Reactions: A Comparative Study

by
Celastin Bebina Thairiyarayar
1,†,
Chia-Hung Huang
2,3,†,
Yasser Ashraf Gandomi
4,
Chien-Te Hsieh
5,6,* and
Wei-Ren Liu
1,*
1
Department of Chemical Engineering, R&D Center for Membrane Technology, Chung Yuan Christian University, 200 Chung Pei Road, Chungli District, Taoyuan City 32023, Taiwan
2
Department of Electrical Engineering, National University of Tainan, No. 33, Sec. 2, Shulin St., West Central District, Tainan City 700, Taiwan
3
Metal Industries Research and Development Centre, Kaohsiung 701, Taiwan
4
Department of Chemical Engineering, Massachusetts Institute of Technology, Cambridge, MA 02142, USA
5
Department of Chemical Engineering and Materials Science, Yuan Ze University, Taoyuan City 32003, Taiwan
6
Department of Mechanical, Aerospace, and Biomedical Engineering, University of Tennessee, Knoxville, TN 37996, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Sustainability 2023, 15(21), 15662; https://doi.org/10.3390/su152115662
Submission received: 19 September 2023 / Revised: 21 October 2023 / Accepted: 30 October 2023 / Published: 6 November 2023
(This article belongs to the Special Issue Advanced Energy Materials and Batteries Technology)

Abstract

:
A sulfide-based solid electrolyte is an enticing non-organic solid-state electrolyte developed under ambient conditions. Na3SbS4, a profoundly enduring substance capable of withstanding exceedingly elevated temperatures and pressures, emerges as a focal point. Within this investigation, we employ dual distinct techniques to fabricate Na3SbS4, encompassing ball milling and the combination of ball milling with sintering procedures. A remarkable ionic conductivity of 3.1 × 10−4 S/cm at room temperature (RT), coupled with a meager activation energy of 0.21 eV, is achieved through a bifurcated process, which is attributed to the presence of tetragonal Na3SbS4 (t-NSS). Furthermore, we delve into the electrochemical performance and cyclic longevity of the Na2/3Fe1/2Mn1/2O2|t-NSS|Na system within ambient environs. It reveals 160 mAh/g initial charge and 106 mAh/g discharge capacities at 0.01 A/g current density. Furthermore, a cycle life test conducted at 0.01 A/g over 30 cycles demonstrates stable and reliable performance. The capacity retention further highlights its enduring energy storage capabilities. This study underscores the sustainable potential of Na3SbS4 as a solid-state electrolyte for advanced energy storage systems.

1. Introduction

Solid electrolytes (SEs) are better separators for battery safety problems than liquid electrolytes [1,2]. Sulfide and selenide [3,4,5], oxide [6,7], β alumina [8,9], and NASICON [10,11], borohydride [12,13], and halo-aluminates [14] are the types of inorganic solid electrolytes which have been used so far as the solid electrolytes in sodium-ion batteries. Due to high ionic conductivity, low cost, good mechanical properties, wide electrochemical window, and formation of stable solid electrolyte interphase (SEI), sulfides received extensive research attention as one of the inorganic solid electrolytes in comparison to the other solid electrolytes [5,15]. Hayashi introduced Na3PS4 as one of the most well-known sulfide SE in all-solid-state sodium-ion batteries by substituting sodium for lithium in Li2S-P2S5 [3,16,17]. This material exhibits high ionic conductivity at room temperature (RT). It has also been investigated for the replacement of phosphorus by antimony as a Na3SbS4 [18,19]. The Na3SbS4 material was synthesized using various approaches, such as solid-state methods, wet processes [5,20,21], mechanochemical processes and then sintering [22,23,24], sintering and then mechanochemical processes [18,25,26], and sintering alone [27,28,29,30]. Different synthesis techniques affected its ionic conductivities and activation energies. Our primary research focus is on the synthesis of Na3SbS4, a sulfide solid electrolyte customized for application in sodium-ion batteries. This choice is motivated by the manifold advantages that sulfide electrolytes offer over their liquid counterparts, including enhanced battery safety and superior ionic conductivity [5,15]. An extensive literature review was conducted to comprehensively address critical aspects, such as ionic conductivity and electrochemical performance. Through our rigorous examination of various synthesis methods, it became evident that certain approaches proved remarkably effective.
For instance, Banerjee et al. reported the synthesis of Na3SbS4 powder using two different methods. The first method involved a solid-state process, where a mixture of Na2S, Sb2S3, and elemental sulfur was heat-treated at different temperatures in a vacuum-sealed quartz ampoule. The second method involved a solution-based process, where the solid powders were dissolved in anhydrous methanol or deionized water. The resulting solution was dried and heat-treated at different temperatures under a vacuum. The resulting Na3SbS4 solid electrolyte exhibited a high conductivity of 1.1 mS cm−1 at 25 °C and a low activation energy of 0.20 eV [5].
Rush et al. synthesized tetragonal Na3SbS4 using a ball milling and sintering method. The synthesis involved ball milling 4 g of Na3SbS4·9H2O with ZrO2 balls for 5 min, then heating the powder under vacuum at a ramp rate of 5 °C/min to 150 °C for 1 h. The resulting powder was transferred to an Argon-filled glove box for further processing. The synthesized tetragonal Na3SbS4 exhibited an impressive ionic conductivity of over 0.8 mS cm−1 at room temperature (RT). In comparison, Na-ion migration in tetragonal Na3SbS4 has an activation energy of 0.3 eV, while Na-ion migration in cubic Na3SbS3 has an activation energy of 0.5 eV [22]. Wu et al. successfully synthesized Na3SbS4 (NSS) through sintering and ball milling. They mixed Na2S, Sb powder, and S in a ratio and pressed the mixture into a pellet. The pellet was heat-treated in a box furnace and ball-milled for 20 h. The resulting NSS material had a tetragonal crystal structure with an ionic conductivity of 1.06 mS cm−1 at room temperature (RT) and an activation energy of 0.216 eV. These results suggested that NSS has high ionic mobility, making it a promising candidate for use in solid-state batteries. This study highlights the effectiveness of solid-state synthesis and ball milling for producing high-performance solid electrolytes [26]. Wang et al. investigated the relationship between pressure loading and the ionic conductivity of Na3SbS4 solid electrolytes prepared by sintering alone. The starting material, Na3SbS4·9H2O, was sintered at 150 °C under a vacuum and held for 1 h. The researchers found that the ionic conductivity of Na3SbS4 increased as pressure loading increased, mainly due to a decrease in grain boundary resistance. As the pressure gradually increased from 0.4 GPa to 0.9 GPa, the ionic conductivity increased from 0.7 mS cm−1 to 1 mS cm−1 and then tended to be constant beyond. Interestingly, the conductivity further increased during decompression due to decreased bulk and grain boundary resistance. After the pressure was released to 0.1 GPa, the impedance spectra were collected, and the ionic conductivity was found to be stable at around 1.6 mS cm−1 for 24 h. This work demonstrated that pressure can effectively increase the ionic conductivity of sulfide-based solid electrolytes [30]. Typically, ball milling is a type of grinding where chemicals are combined to create a fine powder while reducing particle size and enhancing material reactivity. There are no adverse reactions, and it has a wide range of applications and is suitable for continuous usage [31,32]. The mechanical characteristics were enhanced, and its porosity was reduced throughout the sintering process [33,34].
As far as the current literature is concerned, limited research specifically explores using ball milling and sintering methods to synthesize Na3SbS4 solid electrolytes with high ionic conductivity and low activation energy for sodium-ion batteries. Thus, in this study, we attempt to synthesize the t-Na3SbS4 using two methods, the first being ball milling at 5 to 20 h, the second being ball milling at 20 h, and then sintering at three different temperatures at 12 h. Further, we compare their results with characterizations of materials of the crystal phase structure and its properties from X-ray diffraction techniques (XRD) and Rietveld refinement. The surface morphology of the materials and electronic state of the atoms is obtained using scanning electronic microscopy/energy dispersive X-ray spectroscopy (SEM/EDX), X-ray Photoelectron Spectroscopy (XPS), and electrochemical performance of ionic conductivity. The activation energy is characterized with electron impendence spectroscopy (EIS). The evaluation of the performance and sensitivity of a sensor in detecting hydrogen sulfide gas is carried out by an H2S sensor test to demonstrate air stability of as-synthesized NSS solid electrolyte. The corresponding electrochemical tests, such as Galvanostatic charge/discharge tests and rate capability tests, are carried out by using Na2/3(Fe1/2Mn1/2)O2/NSS/Na coin cells. This study not only aims for technical advancements but also emphasizes the sustainability of the synthesized solid electrolyte.

2. Materials and Methods

2.1. Materials and Chemicals

Sodium sulfide (Na2S), a powder with a purity of 90%, was purchased from Alfa Aesar in Karlsruhe, Germany. The Antimony (III) sulfide (Sb2S3), with a purity of 98%, was also purchased from Alfa Aesar in Germany. Additionally, we obtained sulfur with a purity of 99.9% from Sigma-Aldrich in Schnelldorf, Germany.

2.2. Synthesis of Na3SbS4 Solid Electrolyte

The two methods for synthesizing Na3SbS4 (NSS) are shown in Scheme 1: ball milling (BM) alone and ball milling plus sintering (BM-H). Na2S, Sb2S3, and sulfur were the starting components in the first process, in a 3:1:2 ratio. ZrO2 balls with a thickness of 10 balls/10 mm were used to ball mill the mixture for 20 h at 510 rpm. Na3SbS4, also known as NSS-BM-20h, was the end product. In the second process, the final product from the first approach served as the starting material. It was sintered for 12 h under vacuum at three different temperatures (250, 300, and 350 °C) with a heating rate of 5 °C/min, resulting in samples designated as NSS-BM-20h-H250, NSS-BM-20h-H300, and NSS-BM-20h-H350 (also known as NSS-BM-H250, NSS-BM-H300, and NSS-BM-H350, respectively). To preserve the samples’ purity, the whole synthesis and post-processing stages were completed in a glovebox (LABstar, MBraun, München, Germany, H2O and O2 < 0.5 ppm). Additionally, we extended the ball milling time to 25 h and synthesized samples at lower temperatures of 150 and 200 °C, which are labeled as NSS-BM-25 h, NSS-BM20h-H150, and NSS-BM20h-H200, respectively.

2.3. Material Characterization

The crystallographic phase was identified using X-ray powder diffraction (XRD) on a Bruker eco D8 advance diffractometer with monochromatic CuK radiation (λ = 1.54060 Å). The samples were analyzed using closed XRD plates, and XRD data were collected at a scan rate of 2°min−1 in steps of 0.02° in the range of 10–80°. Rietveld structure refinement against XRD data was performed using GSAS–II (1.0.0.) software. The crystal structure was determined using the Diamond-Crystal and Molecular Structure Visualization software Version 3.2. Quantitative elemental analyses and elemental distributions were obtained using energy dispersive spectrometry and electron mapping, while images of the morphology and microstructure of the materials were captured using field emission scanning electron microscopy (FE-SEM, JSM-7600F; JEOL) (EDS, X-MAX, Oxford Instrument, Abingdon, UK). The K-Alpha XPS spectrometer (Thermo Scientific, Waltham, MA, USA) was used to conduct X-ray photoelectron spectroscopy (XPS) to determine the chemical valence states. A specific (Riken Keiki HS-04) H2S sensor was used to assess the air stability of NSS. A 180 mg NSS electrolyte pellet was prepared and pressed at a pressure of 360 MPa. The pellet was carefully inserted into a 12 cm × 12 cm × 12 cm airtight acrylic box from the glovebox. Consequently, the H2S sensor was placed inside the container, and measurements were made every 30 s while the sensor readings were being recorded. This experimental design made it possible to monitor H2S gas concentrations and offered insightful information on the sensor’s receptivity and sensitivity to the target gas.

2.4. Electrochemical Characterization

Using an E.C. Lab, Biologic Potentiostat, model SP-200, experiments were performed to measure the samples’ ionic conductivity and activation energy in the frequency range of 7 MHz–10 Hz at an A.C. amplitude of 10 mV using electrochemical impedance spectroscopy (EIS). With the help of the KP cell, ionic conductivities and electrochemical behaviors were explored, respectively. A PTFE tank with a diameter of 10 mm was used to assemble the KP cell. About 70 mg of NSS powder was placed inside, and it was cold-pressed into a pellet for one minute at a pressure of 360 MPa to test the ionic conductivity and activation energy of the NSS samples.

2.5. Cell Assembly for Sodium-Ion Battery Testing

In sodium-ion battery testing, a half-cell configuration was employed to assess the electrochemical properties. The electrode slurry, consisting of NFMO (Na2/3Fe1/2Mn1/2O2), PVDF and Super P in an 8:1:1 ratio, was thoroughly mixed overnight for homogeneity. It was then coated onto an aluminum foil substrate and subjected to vacuum heating at 120 degrees Celsius for 9 h to fabricate the electrode. The resulting electrode was transferred to a glovebox for further assembly. The cell assembly involved pressing a 70 mg of NSS electrolyte into a pellet under a pressure of 360 MPa, which was then sandwiched between two electrodes, completing the half-cell setup.
Additionally, a drop of 1M NaClO4 solution into a propylene carbonate (PC, Sigma-Aldrich, St. Louis, MO, USA) and 5% fluoroethylene carbonate (FEC, Sigma-Aldrich) was applied to enhance the contact between the cathode and solid electrolyte. Na metal was used as the anode material. The cell assembly was performed in an argon-filled glovebox with controlled oxygen and water levels. Galvanostatic charge–discharge (GCD) analysis was conducted at different current densities of 0.01, 0.02, and 0.05 A/g within a potential window of 2.0 V to 3.8 V vs. Na/Na+. The GCD analysis was carried out using a Neware multichannel battery tester at room temperature (RT) to evaluate the Na-ion storage capacity and cyclic stability. This comprehensive approach allowed for the systematic evaluation of the electrochemical performance of the sodium-ion battery system under controlled conditions.

3. Results and Discussion

Scheme 1 shows the flow chart of how we synthesized NSS by using two different methods. The precursors were ball-milled for 5, 10, 15, and 20 h, with the rationale being to investigate the effect of milling time on the resulting particle size and crystallite size. Figure 1a shows the XRD patterns of NSS at different ball milling time. The peak noted at 17° exhibited substantial intensities, while the presence of the tetragonal phase augmented in correlation with the duration of milling. Following 20 h of ball milling, the emergence of a tetragonal phase was discerned and subsequently verified through the standard pattern of ICSD 47291 and another pertinent prior reference [27]. The insets in Figure 1a display photo images of NSS-BM-0 h, NSS-BM-5 h, NSS-BM-10 h, NSS-BM-15 h, and NSS-BM-20 h, respectively. In Figure 1b, particular X-ray diffraction (XRD) analysis was shown in an expanded form of 15° to 20° to observe the peaks between 17° to 18°. One of the tetragonal peaks emerged at 17.5° after 10 h of ball milling, evolving from the initial peaks at 15.9° (Na2Sb) and 17.8° (sulfur) presented in the precursor. This indicated structural modifications and potential phase transformations induced by the extended milling process. Further increasing the ball milling time to 20 h resulted in splitting the single peak into two narrow, smaller peaks. These observations provided insights into the structural changes and crystallinity of the material as a result of different ball milling durations. Figure 1c displays the Rietveld refinement analysis of NSS after 20 h. The result showed that, after 20 h ball milling, two phases were presented, with the main phase being tetragonal Na3SbS4 (P-421c) at 80.6%, with 26.9, 31.2, and 44.7 degrees 2θ peak from cubic NaSbS2 (Fm-3m) at 19.4%. The Rp and Rwp values obtained from the refinement were 14.16% and 18.73%, respectively. In Figure 1d, the grain size of Na3SbS4-0 h that was subjected to 20 h of ball milling was calculated using the Scherer equation [35]. The grain sizes of NSS-BM-0 h, NSS-BM-5 h, NSS-BM-10 h, and NSS-BM-20 h were determined to be 220, 218, 203, and 196 nm, respectively. The results indicated that the crystallite size of NSS decreased from 220 to 196 nm when increasing the milling time from 0 h to 20 h. Supplementary Figure S2a displays the X-ray diffraction (XRD) patterns and corresponding photographs of samples subjected to 20 and 25 h of ball milling. An interesting observation is the emergence of distinct 2θ peaks around 16.6° in BM-25 h, indicating the presence of NaSbS2. This suggests a transformation or phase transition within the material due to the extended ball milling process. This alteration may be attributed to enhanced structural rearrangements and increased defects induced by prolonged milling, leading to the formation of NaSbS2.
Figure 2a–d display SEM images of NSS-BM-0 h, NSS-BM-10 h, NSS-BM-15 h, and NSS-BM-20 h, respectively. In Figure 2, the SEM image shows a comparison between the precursor (NSS-BM-0 h) and the specific particle sizes obtained in the range of 1 µm at each milling interval. After undergoing ball milling for a duration of 10 h, the resultant particle size was measured to be within the range of 49–26 µm. Similarly, upon extending the milling time to 15 h, the particle size exhibited a distribution of 10–20 µm, and further increasing the milling time to 20 h led to a particle size range of 1–10 µm. The unmistakable inference drawn from these results is that the ball milling technique significantly and distinctly diminished the grain sizes of the particles, albeit in an irregular and unpredictable manner.
In Figure 3a, XRD patterns of NSS-BM-H250, NSS-BM-H300, and NSS-BM-H350 are shown alongside the standard Na3SbS4 pattern (#47291, ICSD). Insets in Figure 3a display sample images. Samples exhibited a higher intensity and a tetragonal phase with rising sintering temperature, and were darkened in color. The XRD analysis confirms tetragonal peaks for NSS-BM sintered at 250 °C, 300 °C, and 350 °C, which align with the standard pattern. Figure 3b charts show the NSS-BM grain size evolution via the Scherer equation as a function of sintering temperature. Crystallite size increases (250 °C: 273 nm, 300 °C: 297 nm) and then decreases (350 °C: 208 nm). NSS-BM-250, milled for 20 h and sintered at 250 °C, is the largest in size (273 nm) compared to NSS-20 h (196 nm). Supplementary Figure S3a showcases the X-ray diffraction (XRD) patterns along with accompanying photographs of NSS-BM20h-H150, 200, and 250 samples. Notably, the XRD pattern of NSS-BM20h-H150 exhibits several minor impurity peaks. This occurrence can be attributed to the lower sintering temperature (150 °C), which may not provide sufficient energy to promote complete phase transformation and crystallization. Consequently, residual impurities persist in the sample, leading to the observed minor peaks in the XRD pattern.
Rietveld refinement in Figure 2c shows a cubic phase decrease with a higher sintering temperature. The 20 h ball-milled sample sintered at 250 °C has the highest cubic phase content. This intricate interplay between ball milling, sintering, and crystallographic phases influences material properties. Figure 3d shows the crystal structure of Na3SbS4-BM-H250, which possesses a tetragonal crystal system and is a member of the space group P-421c, with Rp at 14.16% and Rwp at 18.73%. The lattice constants were determined to be a = 7.18392(0) Å and c = 7.26153(2) Å.
The Na1 site in 4d, the Na2 site in 2b, the Sb1 site in 2a, and the S site in 8e are identified as the four unique sites in the crystal structure. Na ions are surrounded by the SbS43- tetrahedral unit. The crystallographic details are listed in Table 1.
Figure 4a–d display the SEM images of NSS-BM-20 h, NSS-BM-H250, NSS-BM-H300, and NSS-BM-H350, respectively. The surface morphology of NSS-BM-20 h, shown in Figure 4a, revealed an irregular form in the range of 1 µm. Subsequently, sintering the sample at different temperatures induced substantial changes in its morphology. With increasing sintering temperatures, there was a notable increase in the neck formation between particles, resulting in the particle merging at NSS-BM-H250 (10–16 µm), 300 °C (13–45 µm), and obtaining a dense morphology at 350 °C. Moreover, the impact of the sintering temperature on the surface morphology and crystal structure of Na3SbS4 (NSS) is noteworthy. This influence is evidenced by the substantial crystal size augmentation observed in NSS-BM-H250, which can be attributed to the distinct grain sizes in NSS-BM-20 h (ranging from 1 to 10 µm) and NSS-BM-H250 (ranging from 10 to 16 µm). The sintering temperature acts as a critical factor in determining these variations. As the sintering temperature increases, the particles experience enhanced diffusion, leading to stronger bonding between adjacent particles. This phenomenon promotes particle coalescence and densification of the material.
The SEM/EDX mapping of NSS-BM-20 h (a)–(d) and NSS-BM-20h–H250 (e)–(h) is displayed in Figure 5. The elements Na, Sb, and S were evenly distributed across the sample, according to the four sections of the mapping. The SEM/EDX mapping, which was thoroughly approached, explains the synthesis of NSS samples. Processing was likely carried out under ideal conditions, resulting in a highly pure and homogenous product, as seen by the elements’ remarkably uniform appearance.
Figure 6a–h show the XPS spectra of the NSS materials with the BM (NSS-BM-20 h) and BM processes, which were then sintered at different temperatures (NSS-BM-20h-H250, NSS-BM-20h-H300, NSS-BM-20h-H350). The XPS results revealed two single signals for SbS43−, with no shift observed in the Sb 3d3 spectra of NSS-BM-20 h (Figure 6a), NSS-BM-20h-H250 (Figure 6c), NSS-BM-20h-H300 (Figure 6e), and NSS-BM-20h-H350 (Figure 6g). The increase in the O 1s signal with decreasing temperature was due to the removal of the adsorbed species and an increase in the stoichiometric ratio of Na/Sb in the sample. The S 2p spectra of NSS-BM-20 h in Figure 6b and NSS-BM-20h-H250 in Figure 6d, NSS-BM-20h-H300 in Figure 6f, and NSS-BM-20h-H350 in Figure 6h showed a Sb-Na-S signal, which was identified as a single doublet signal. The S 2p spectra, on the other hand, revealed no impurity signals, indicating that the samples were highly pure. The slightly lower signal observed in NSS-BM-20h-H250, compared to NSS-BM-20 h, could be due to the formation of surface defects or changes in the oxidation state of the sulfur species, which was confirmed by a previous study [23].
Figure 7a shows the ionic conductivity of NSS BM-20 h, NSS BM-20h-H250, 300, and 350 °C at room temperature (RT). The corresponding resistance values are provided in Table 2. Figure 7a and Table 2 show the BM-20 h total grain boundary resistance increasing from 176 Ω to 202 Ω in BM-20h-H250, indicating a decrease of bulk resistance (Rbulk), grain boundary resistance (Rgb), and total resistance (Rtotal) of the samples, and an increase in sintering temperature, resulting in a decrease in ionic conductivity. The inset in Figure 7a displays the corresponding equivalent circuit. However, the BM-20h-H250 sample showed the highest ionic conductivity (3.1 × 10−4 S cm−1). Interestingly, it is identical to a previous study, in which they used direct Na3SbS4.9H2O as a starting material [27].
Figure 7b shows the Arrhenius plots of NSS BM-20h, NSS BM-20h-H250, 300, and 350 samples at 25, 40, 60, and 80 °C, respectively. The BM-20h-H250 sample exhibited the lowest activation energy (0.21), and its ionic conductivity was 0.31 m S cm−1 at 25 °C and 1.12 m S cm−1 at 80 °C. The results further confirmed that a low grain boundary leads to high ionic conductivity and low activation energy. The equivalent circuit of the system and the Arrhenius equation elucidated the activation energy, which was estimated using Equation (1), as follows:
σ = A   e x p E a / k T
where A is the pre-exponential factor, T is the absolute temperature, and k is the Boltzmann constant, i.e., the absolute temperature [36]. The comprehensive dataset presented in Table 2 furnishes substantiation for the derived conclusions and delineates the influential facets governing the ionic conductivity and activation energy of the NSS BM-20 h and NSS BM-20h-H250, 300, and 350 samples. These factors include bulk resistance (Rbulk), grain boundary resistance (Rgb), and total resistance (Rtotal).
Supplementary Figure S1a offers a clear depiction of the ionic conductivity plot, showcasing the effects of varying ball milling durations (10 h, 15 h) on NSS samples. As the duration of ball milling increases, there is a marked reduction in total resistance. This reduction reaches its zenith at the 20 h mark, resulting in a substantial boost in ionic conductivity. These enhancements can be attributed to the refinement in the crystal structure and the diminishment of particle size. Together, these elements synergize to amplify ion mobility and effectively suppress resistance. Supplementary Figure S1b complements this observation by presenting the activation energy corresponding to ball milling durations of 10 h, 15 h, and 20 h. It is worth noting that an extension in ball milling time leads to a reduction in activation energy, signifying an improvement in ionic conductivity. For a detailed breakdown of the ionic conductivity and activation energy values, please refer to Supplementary Table S1. In contrast, at the 25 h mark (Supplementary Figure S2), there is a notable surge in the total resistance from 202 Ω to 403 Ω, resulting in a significant drop in ionic conductivity from 2.77 to 1.32 × 10−4 S cm−1. This decline can be ascribed to the prolonged milling process, which encourages particle agglomeration and carries the potential for crystal damage. Consequently, this leads to an overall increase in total resistance and a corresponding decrease in ionic conductivity.
Furthermore, Supplementary Figure S3b imparts valuable insights into the impact of the sintering temperature on ionic conductivity, with a specific focus on NSS BM-20 h samples sintered at 150 °C, 200 °C, and 250 °C. A conspicuous trend emerges: as sintering temperature diminishes, so does ionic conductivity. This phenomenon finds its roots in the diminished thermal energy at lower temperatures, which effectively curtails ion mobility, consequently resulting in a discernible reduction in ionic conductivity. Detailed ionic conductivity values are provided in Supplementary Table S2.
Figure 8 portrays the air stability of the Na3SbS4-BM-H250 solid electrolyte. Remarkably, the results divulged the absence of any conspicuous reaction discernible within the H2S sensor during an extended period that exceeded 45 min, even in the presence of elevated humidity levels (56%) after air exposure at room temperature (RT). This captivating disclosure underscores the enigmatic reality that the sulfides within the Na3SbS4 solid electrolyte remain impervious to atmospheric circumstances, effectively impeding the genesis of H2S gas. This sustainable characteristic showcases the material’s praiseworthy endurance and bolsters its safety paradigm [19,37,38].
Figure 9a–d illustrate the electrochemical performance of the NFMO (Na2/3Fe1/2Mn1/2O2) cathode|Na3SbS4 -BM-H250 solid electrolyte|Na foil at RT. In Figure 9a, a comprehensive C-rate test was conducted, revealing specific capacities achieved at various current densities: 106 mAh/g at 0.01 A/g, 90 mAh/g at 0.02 A/g, 67 mAh/g at 0.05 A/g, and 98 mAh/g upon reverting to 0.01 A/g. This unveils the system’s substantial charge storage potential at lower current densities, with a capacity reduction at higher densities and some recovery upon returning to the initial density. Figure 9b depicts charge and discharge cycles at current densities of 0.01 A/g, 0.02 A/g, and 0.05 A/g. The specific discharge capacities are 106 mAh/g, 90 mAh/g, and 67 mAh/g, respectively. These outcomes elucidate the discharge behavior and its variation with different current densities. Complementing the findings in Figure 9c, Figure 9d shows the galvanostatic charge and discharge cycles at specific cycles (1 to 3, 5, 10, 20, 30). These voltage profiles visually represent the system’s electrochemical behavior and confirm its stability and consistency throughout the tested cycles. Overall, the remarkable cycle life performance, sustained discharge specific capacity, and high columbic efficiency observed in Figure 9c,d highlight the potential of the NFMO and Na3SbS4 solid electrolyte system for practical energy storage applications. The system’s ability to deliver reliable and efficient charge storage over an extended duration makes it a promising candidate for future advancements in energy storage technologies, aligning with the sustainability goals of modern energy solutions. This is particularly significant in the context of solid electrolytes, as evidenced by the comparison with other literature in Table 3, which highlights the performance of our work in relation to previous studies.

4. Conclusions

In this study, we successfully synthesized the tetragonal phase of the Na3SbS4 (NSS) solid electrolyte using two distinct methodologies. The second method (BM-H) is particularly significant, which was executed at a temperature of 250 °C and yielded outcomes of remarkable importance. This approach has rendered a conspicuous ionic conductivity of 0.31 mS cm−1 at room temperature while exhibiting a remarkably diminutive activation energy of 0.21 eV. Moreover, it maintains stability in the air. The Na3SbS4 synthesized through the BM-H process at 250 °C has effectively functioned as a solid electrolyte, assuming the role of an intermediary amid the Na2/3Fe1/2Mn1/2O2 (NFMO) electrode and the sodium counter electrode. This arrangement showcased a significant discharge capacity of 104 mAh/g at a current density of 0.01 A/g. Furthermore, an impressive coulombic efficiency of 99.3% was achieved over 30 cycles, thereby highlighting the efficiency of charge transfer processes. The result indicates that Na3SbS4 stands as a preeminent solid electrolyte option for sodium-ion batteries, aligning with the sustainable energy storage solutions. Moving forward, it is recommended to focus on optimizing the BM-H (second method) synthesis process for large-scale production, fine-tuning the ball milling time to control grain size, and exploring alternative characterization techniques that are more aligned with the specific properties of Na3SbS4. Additionally, modifications to the electrode design should be considered for improved interface compatibility, and extended cycling tests are crucial for assessing long-term stability. Safety assessments and cost-effective production methods should also be explored, and collaborative research with experts in relevant fields can provide invaluable insights for further development and application of Na3SbS4 solid electrolytes.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/su152115662/s1.

Author Contributions

C.B.T. and C.-H.H. wrote this paper and analyzed the data; Y.A.G. and W.-R.L. conceived and designed the experiments; C.-T.H. analyzed the data; C.-T.H. and W.-R.L. supervised and revised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

All data are available within the article and in the supplementary materials archive.

Acknowledgments

The authors gratefully acknowledged to National Science of Technology Council (NSTC) project grant no. NSTC 112-2218-E-007-023, 111-2622-E-033-007, 111-2923-E-006-009, 112-2923-E-006-004 and 111-2221-E-033-004-MY3.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bron, P.; Johansson, S.; Zick, K.; Schmedt auf der Günne, J.; Dehnen, S.; Roling, B. Li10SnP2S12: An affordable lithium superionic conductor. J. Am. Chem. Soc. 2013, 135, 15694–15697. [Google Scholar] [CrossRef]
  2. Wang, Y.; Song, S.; Xu, C.; Hu, N.; Molenda, J.; Lu, L. Development of solid-state electrolytes for sodium-ion battery—A short review. Nano Mater. Sci. 2019, 1, 91–100. [Google Scholar] [CrossRef]
  3. Hayashi, A.; Noi, K.; Sakuda, A.; Tatsumisago, M. Superionic glass-ceramic electrolytes for room-temperature rechargeable sodium batteries. Nat. Commun. 2012, 3, 856. [Google Scholar] [CrossRef]
  4. Bo, S.H.; Wang, Y.; Kim, J.C.; Richards, W.D.; Ceder, G. Computational and experimental investigations of Na-ion conduction in cubic Na3PSe4. Chem. Mater. 2016, 28, 252–258. [Google Scholar] [CrossRef]
  5. Banerjee, A.; Park, K.H.; Heo, J.W.; Nam, Y.J.; Moon, C.K.; Oh, S.M.; Jung, Y.S. Na3SbS4: A solution processable sodium superionic conductor for all-solid-state sodium-ion batteries. Angew. Chem. 2016, 128, 9786–9790. [Google Scholar] [CrossRef]
  6. Zhu, J.; Wang, Y.; Li, S.; Howard, J.W.; Neuefeind, J.; Ren, Y.; Zhao, Y. Sodium ion transport mechanisms in antiperovskite electrolytes Na3OBr and Na4OI2: An in situ neutron diffraction study. Inorg. Chem. 2016, 55, 5993–5998. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, Y.; Wang, Q.; Liu, Z.; Zhou, Z.; Li, S.; Zhu, J.; Zhao, Y. Structural manipulation approaches towards enhanced sodium ionic conductivity in Na-rich antiperovskites. J. Power Source 2015, 293, 735–740. [Google Scholar] [CrossRef]
  8. Yao, Y.-F.Y.; Kummer, J.T. Ion exchange properties of and rates of ionic diffusion in beta-alumina. J. Inorg. Nucl. Chem. 1967, 29, 2453–2475. [Google Scholar]
  9. Beevers, C.; Ross, M. The Crystal Structure of “Beta Alumina” Na2O11Al2O3. Zeitschrift für Kristallographie. Cryst. Mater. 1937, 97, 59–66. [Google Scholar]
  10. Hong, H.Y.-P. Crystal structures and crystal chemistry in the system Na1+xZr2SixP3−xO12. Mater. Res. Bull. 1976, 11, 173–182. [Google Scholar] [CrossRef]
  11. Von Alpen, U.; Bell, M.F.; Höfer, H.H. Compositional dependence of the electrochemical and structural parameters in the NASICON system (Na1+xSixZr2P3−xO12). Solid State Ion. 1981, 3, 215–218. [Google Scholar] [CrossRef]
  12. Udovic, T.J.; Matsuo, M.; Unemoto, A.; Verdal, N.; Stavila, V.; Skripov, A.V.; Orimo, S.I. Sodium superionic conduction in Na2B12H12. Chem. Commun. 2014, 50, 3750–3752. [Google Scholar] [CrossRef] [PubMed]
  13. Santos, D.M.F.; Sequeira, C.A.C. Sodium Borohydride as a fuel for the future. Renew. Sustain. Energy Rev. 2011, 15, 3980–4001. [Google Scholar] [CrossRef]
  14. Spearing, D.R.; Stebbins, J.F.; Farnan, I. Diffusion and the dynamics of displacive phase transitions in cryolite (Na3AlF6) and chiolite (Na5Al3F14): Multi-nuclear NMR studies. Phys. Chem. Miner. 1994, 21, 373–386. [Google Scholar] [CrossRef]
  15. Tatsumisago, M.; Hayashi, A. Sulfide glass-ceramic electrolytes for all-solid-state lithium and sodium batteries. Int. J. Appl. Glass Sci. 2014, 5, 226–235. [Google Scholar] [CrossRef]
  16. Berbano, S.S.; Seo, I.; Bischoff, C.M.; Schuller, K.E.; Martin, S.W. Formation and structure of Na2S + P2S5 amorphous matrials prepared by melt-quenching and mechanical milling. J. Non-Cryst. Solids 2012, 358, 93–98. [Google Scholar] [CrossRef]
  17. Mizuno, F.; Hayashi, A.; Tadanaga, K.; Tatsumisago, M. New, highly ion-conductive crystals precipitated from Li2S-P2S5 glasses. Adv. Mater. 2005, 17, 918–921. [Google Scholar] [CrossRef]
  18. Zhang, L.; Zhang, D.; Yang, K.; Yan, X.; Wang, L.; Mi, J.; Li, Y. Vacancy-contained tetragonal Na3SbS4 superionic conductor. Adv. Sci. 2016, 3, 1600089. [Google Scholar] [CrossRef]
  19. Wang, H.; Chen, Y.; Hood, Z.D.; Sahu, G.; Pandian, A.S.; Keum, J.K.; Liang, C. An air-stable Na3SbS4 superionic conductor prepared by a rapid and economical synthetic procedure. Angew. Chem. 2016, 128, 8693–8697. [Google Scholar] [CrossRef]
  20. Wang, Z.; Zhang, L.; Shang, X.; Wang, W.; Yan, X.; Yu, C.; Wang, L.M. Enhanced electrochemical performance enabled by ionic-liquid-coated Na3SbS4 electrolyte encapsulated in flexible filtration membrane. Chem. Eng. J. 2022, 428, 132094. [Google Scholar] [CrossRef]
  21. Kim, T.W.; Park, K.H.; Choi, Y.E.; Lee, J.Y.; Jung, Y.S. Aqueous-solution synthesis of Na3SbS 4 solid electrolytes for all-solid-state Na-ion batteries. J. Mater. Chem. A 2018, 6, 840–844. [Google Scholar] [CrossRef]
  22. Rush, L.E., Jr.; Hood, Z.D.; Holzwarth, N.A.W. Unraveling the electrolyte properties of Na3SbS4 through computation and experiment. Phys. Rev. Mater. 2017, 1, 075405. [Google Scholar] [CrossRef]
  23. Zhang, S.; Zhao, Y.; Zhao, F.; Zhang, L.; Wang, C.; Li, X.; Sun, X. Gradiently sodiated alucone as an interfacial stabilizing strategy for solid-state Na metal batteries. Adv. Funct. Mater. 2020, 30, 2001118. [Google Scholar] [CrossRef]
  24. Liang, B.; Yu, L.; Wang, G.; Lin, C.; Gao, C.; Shen, X.; Jiao, Q. Physical and electrochemical behavior of affordable Na3SbS4 solid electrolyte at different heat treatment temperatures. Ceram. Int. 2022, 48, 30144–30150. [Google Scholar] [CrossRef]
  25. Zhang, D.; Cao, X.; Xu, D.; Wang, N.; Yu, C.; Hu, W.; Zhang, L. Synthesis of cubic Na3SbS4 solid electrolyte with enhanced ion transport for all-solid-state sodium-ion batteries. Electrochim. Acta 2018, 259, 100–109. [Google Scholar] [CrossRef]
  26. Wu, E.A.; Kompella, C.S.; Zhu, Z.; Lee, J.Z.; Lee, S.C.; Chu, I.H.; Meng, Y.S. New insights into the interphase between the Na metal anode and solid-state sulfide electrolytes: A joint experimental and computational study. ACS Appl. Mater. Interfaces 2018, 10, 10076–10086. [Google Scholar] [CrossRef]
  27. 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]
  28. Zhang, Q.; Zhang, C.; Hood, Z.D.; Chi, M.; Liang, C.; Jalarvo, N.H.; Wang, H. Abnormally low activation energy in cubic Na3SbS4 superionic conductors. Chem. Mater. 2020, 32, 2264–2271. [Google Scholar] [CrossRef]
  29. Wang, H.; Chen, Y.; Hood, Z.D.; Keum, J.K.; Pandian, A.S.; Chi, M.; Sunkara, M.K. Revealing the structural stability and Na-ion mobility of 3D superionic conductor Na3SbS4 at extremely low temperatures. ACS Appl. Energy Mater. 2018, 1, 7028–7034. [Google Scholar] [CrossRef]
  30. Wang, H.; Yu, M.; Wang, Y.; Feng, Z.; Wang, Y.; Lü, X.; Liang, C. In-situ investigation of pressure effect on structural evolution and conductivity of Na3SbS4 superionic conductor. J. Power Source 2018, 401, 111–116. [Google Scholar] [CrossRef]
  31. Chaudhuri, S.; Ghosh, A.; Chattopadhyay, S.K. Green synthetic approaches for medium ring–sized heterocycles of biological and pharmaceutical interest. In Green Synthetic Approaches for Biologically Relevant Heterocycles; Elsevier: Amsterdam, The Netherlands, 2021; pp. 617–653. [Google Scholar]
  32. El-Eskandarany, M.S. The history and necessity of mechanical alloying. Mech. Alloying 2015, 2, 13–47. [Google Scholar]
  33. Kang, S.J.L. Sintering: Densification, Grain Growth, and Microstructure; Elsevier: Amsterdam, The Netherlands, 2004. [Google Scholar]
  34. Pasagada, V.K.V.; Yang, N.; Xu, C. Electron beam sintering (EBS) process for Ultra-High Temperature Ceramics (UHTCs) and the comparison with traditional UHTC sintering and metal Electron Beam Melting (EBM) processes. Ceram. Int. 2022, 48, 10174–10186. [Google Scholar] [CrossRef]
  35. Sottmann, J.; Herrmann, M.; Vajeeston, P.; Hu, Y.; Ruud, A.; Drathen, C.; Wragg, D.S. How Crystallite Size Controls the Reaction Path in Nonaqueous Metal Ion Batteries: The Example of Sodium Bismuth Alloying. Chem. Mater. 2016, 28, 2750–2756. [Google Scholar] [CrossRef]
  36. Xu, R.C.; Xia, X.H.; Yao, Z.J.; Wang, X.L.; Gu, C.D.; Tu, J.P. Preparation of Li7P3S11 glass-ceramic electrolyte by dissolution- evaporation method for all-solid-state lithium-ion batteries. Electrochim. Acta 2016, 219, 235–240. [Google Scholar] [CrossRef]
  37. Tsuji, F.; Masuzawa, F.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]
  38. Nikodimos, Y.; Huang, C.J.; Taklu, B.W.; Su, W.N.; Hwang, B.J. Chemical stability of sulfide solid-state electrolytes: Stability toward humid air and compatibility with solvents and binders. Energy Environ. Sci. 2022, 15, 991–1033. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of Na3SbS4 powder with mechanochemical (BM) and BM processes followed by sintering process.
Scheme 1. Synthesis of Na3SbS4 powder with mechanochemical (BM) and BM processes followed by sintering process.
Sustainability 15 15662 sch001
Figure 1. (a) X-ray diffraction patterns of BM-only Na3SbS4 with respective sample photos. (b) Magnified version of main peak from 15 to 20° at 2 θ. (c) Rietveld refinement profile of NSS-BM-20 h using Bragg positions as t-Na3SbS4 and c-NaSbS2. (d)The crystallite size of ball milling sample as an hours function.
Figure 1. (a) X-ray diffraction patterns of BM-only Na3SbS4 with respective sample photos. (b) Magnified version of main peak from 15 to 20° at 2 θ. (c) Rietveld refinement profile of NSS-BM-20 h using Bragg positions as t-Na3SbS4 and c-NaSbS2. (d)The crystallite size of ball milling sample as an hours function.
Sustainability 15 15662 g001
Figure 2. SEM images of (a) NSS precursor (NSS-BM-0 h), (b) NSS-BM-10 h, (c) NSS-BM-15 h, and (d) NSS-BM-20 h.
Figure 2. SEM images of (a) NSS precursor (NSS-BM-0 h), (b) NSS-BM-10 h, (c) NSS-BM-15 h, and (d) NSS-BM-20 h.
Sustainability 15 15662 g002
Figure 3. (a) X-ray diffraction patterns of BM process and then sintering Na3SbS4 (BM-H), with respective sample photos. (b) The crystallite size of NSS BM-H sample as a function of different temperatures. (c) Rietveld refinement profile of NSS-BM20h-H250 using Bragg positions as t-Na3SbS4 and c-NaSbS2. (d) Crystal structure of t-Na3SbS4-BM20h-H250.
Figure 3. (a) X-ray diffraction patterns of BM process and then sintering Na3SbS4 (BM-H), with respective sample photos. (b) The crystallite size of NSS BM-H sample as a function of different temperatures. (c) Rietveld refinement profile of NSS-BM20h-H250 using Bragg positions as t-Na3SbS4 and c-NaSbS2. (d) Crystal structure of t-Na3SbS4-BM20h-H250.
Sustainability 15 15662 g003
Figure 4. SEM images of (a) NSS-BM-20 h as precursor, (b) NSS-BM-H250, (c) NSS-BM-H300, and (d) NSS-BM-H350.
Figure 4. SEM images of (a) NSS-BM-20 h as precursor, (b) NSS-BM-H250, (c) NSS-BM-H300, and (d) NSS-BM-H350.
Sustainability 15 15662 g004
Figure 5. SEM/EDX mapping of (ad) NSS-BM-20 h and (eh) NSS-BM-20h–H250.
Figure 5. SEM/EDX mapping of (ad) NSS-BM-20 h and (eh) NSS-BM-20h–H250.
Sustainability 15 15662 g005
Figure 6. High-resolution XPS spectra of NSS with BM process, and BM process and then different temperatures: (a) Sb 3d3 spectra and (b) S 2p spectra of NSS-BM-20 h; (c) Sb 3d3 spectra and (d) S 2p spectra of NSS-BM-20h-H250; (e) Sb 3d3 spectra and (f) S 2p spectra of NSS-BM-20h-H300; (g) Sb 3d3 spectra and (h) S 2p spectra of NSS-BM-20h-H350.
Figure 6. High-resolution XPS spectra of NSS with BM process, and BM process and then different temperatures: (a) Sb 3d3 spectra and (b) S 2p spectra of NSS-BM-20 h; (c) Sb 3d3 spectra and (d) S 2p spectra of NSS-BM-20h-H250; (e) Sb 3d3 spectra and (f) S 2p spectra of NSS-BM-20h-H300; (g) Sb 3d3 spectra and (h) S 2p spectra of NSS-BM-20h-H350.
Sustainability 15 15662 g006
Figure 7. (a) Nyquist impedance plots at RT: equivalent circuit in the system and (b) Arrhenius plots at 25, 40, 60, and 80 °C of NSS-BM20 h and NSS BM-H250, 300, and 350 °C.
Figure 7. (a) Nyquist impedance plots at RT: equivalent circuit in the system and (b) Arrhenius plots at 25, 40, 60, and 80 °C of NSS-BM20 h and NSS BM-H250, 300, and 350 °C.
Sustainability 15 15662 g007
Figure 8. H2S concentration of Na3SbS4-BM-H250 hydrolysis in air as a function of time.
Figure 8. H2S concentration of Na3SbS4-BM-H250 hydrolysis in air as a function of time.
Sustainability 15 15662 g008
Figure 9. Cycling performance of Na2/3Fe1/2Mn1/2O2|Na3SbS4-BM-H250|Na at RT. (a) C-rate test, (b) galvanostatic charge and discharge cycle of different current densities, (c) cycle life test, and (d) galvanostatic charge and discharge cycle of current density at 0.01 A/g over 30 cycles.
Figure 9. Cycling performance of Na2/3Fe1/2Mn1/2O2|Na3SbS4-BM-H250|Na at RT. (a) C-rate test, (b) galvanostatic charge and discharge cycle of different current densities, (c) cycle life test, and (d) galvanostatic charge and discharge cycle of current density at 0.01 A/g over 30 cycles.
Sustainability 15 15662 g009
Table 1. Crystal structure information of Na3SbS4-BM20h-H250. Crystal system—tetragonal lattice parameters: A = 7.1839(20) Å. C = 7.2615(32) Å. Space group: P-421c (114). Volume, Z: V = 373.08(6) Å3, Z = 2.
Table 1. Crystal structure information of Na3SbS4-BM20h-H250. Crystal system—tetragonal lattice parameters: A = 7.1839(20) Å. C = 7.2615(32) Å. Space group: P-421c (114). Volume, Z: V = 373.08(6) Å3, Z = 2.
AtomsWyckoffOccupancyx/ay/az/c
Na14d101/21/2
Na22b1001/2
Sb2a1000
S8e10.293310.330800.68745
Rwp = 18.73%; Rp = 14.16%; χ2 = 1.29.
Table 2. Comparison of ionic conductivity and activation energy of Na3SbS4 with BM process, and BM process and then sintering at different temperatures.
Table 2. Comparison of ionic conductivity and activation energy of Na3SbS4 with BM process, and BM process and then sintering at different temperatures.
SamplesL
(cm)
S
(cm2)
Rb
(Ω)
Rgb
(Ω)
Rb + Rgb
(Ω)
σ25 °C
(S cm−1)
Ea
(eV)
NSS-BM-20 h0.0440.785431171.4202.42.77 × 10−40.22
NSS-BM-H2500.0430.785411.43164.7176.133.11 × 10−40.21
NSS-BM-H3000.0430.785413.38518.9532.281.03 × 10−40.23
NSS-BM-H3500.0390.785426.55439.2465.751.07 × 10−40.28
Table 3. Comparison of σRT and Ea of Na3SbS4 from our results with those from prior reporters.
Table 3. Comparison of σRT and Ea of Na3SbS4 from our results with those from prior reporters.
Solid ElectrolytesIonic Conductivity at RT (mS cm−1)Activation Energy (eV)References
Na3SbS40.310.21This work
Na3SbS40.540.074[24]
Na3SbS40.310.2[27]
t-Na3SbS40.60.224[28]
Na3SbS41.060.21[26]
t-Na3SbS430.25[18]
t-Na3SbS40.80.3[22]
Na3SbS41.050.22[29]
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

Thairiyarayar, C.B.; Huang, C.-H.; Gandomi, Y.A.; Hsieh, C.-T.; Liu, W.-R. Synthesis and Characterization of Na3SbS4 Solid Electrolytes via Mechanochemical and Sintered Solid-State Reactions: A Comparative Study. Sustainability 2023, 15, 15662. https://doi.org/10.3390/su152115662

AMA Style

Thairiyarayar CB, Huang C-H, Gandomi YA, Hsieh C-T, Liu W-R. Synthesis and Characterization of Na3SbS4 Solid Electrolytes via Mechanochemical and Sintered Solid-State Reactions: A Comparative Study. Sustainability. 2023; 15(21):15662. https://doi.org/10.3390/su152115662

Chicago/Turabian Style

Thairiyarayar, Celastin Bebina, Chia-Hung Huang, Yasser Ashraf Gandomi, Chien-Te Hsieh, and Wei-Ren Liu. 2023. "Synthesis and Characterization of Na3SbS4 Solid Electrolytes via Mechanochemical and Sintered Solid-State Reactions: A Comparative Study" Sustainability 15, no. 21: 15662. https://doi.org/10.3390/su152115662

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