Next Article in Journal
Corrosion Properties of DLC Film in Weak Acid and Alkali Solutions
Next Article in Special Issue
Room Temperature Synthesis of Branched ZnO Nanowires Array with Tunable Morphology
Previous Article in Journal
Nanomaterials Applied in Coatings: Synthesis, Structures, Properties, and Applications
Previous Article in Special Issue
Regulating Monolayer Aligned Silver Nanowire Coatings for Energy-Saving Windows
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bi2O3-Assisted Sintering of Na3Zr2Si2PO12 Electrolyte for Solid-State Sodium Metal Batteries

1
Tianjin Key Laboratory for Photoelectric Materials and Devices, School of Materials Science and Engineering, Tianjin University of Technology, Tianjin 300384, China
2
School of Science, Tianjin University of Technology, Tianjin 300384, China
*
Authors to whom correspondence should be addressed.
Coatings 2022, 12(11), 1774; https://doi.org/10.3390/coatings12111774
Submission received: 8 October 2022 / Revised: 13 November 2022 / Accepted: 14 November 2022 / Published: 20 November 2022
(This article belongs to the Collection Feature Papers of Coatings for Energy Applications)

Abstract

:
Solid-state sodium metal batteries using non-flammable solid-state electrolytes are recognized as next-generation energy storage technology in view of their merits of high safety and low cost. However, the lower ion conductivity (below the application requirements of 10−3 S cm−1) and interface issues that exist in electrolytes/electrodes for most solid-state electrolytes hinder their practical application. In this paper, NASICON-type Na3Zr2Si2PO12 (NZSP) electrolytes with enhanced ion conductivity are synthesized by the Bi2O3-assisted sintering method. The influence of the Bi2O3 sintering agent content on the crystalline phase, microstructure, density and ion conductivity as well as the electrochemical performances applied in batteries for the obtained NZSP electrolytes are investigated in detail. With the presence of Bi2O3, the formed Na3Bi(PO4)2 impurity increased the Si/P ratio in the NASICON structure with higher Na+ occupancy, then enhanced the ionic conductivity to a level of 1.27 × 10−3 S cm−1. Unfortunately, the Bi2O3-assisted sintered NZSP shows a degradation in the cycling stability when applied to solid-state sodium batteries because of the decreased interfacial stability with Na anodes. The formation of a Bi-Na alloy during cycling might be conducive to Na dendrite growth in electrolytes, degrading the cycling performance. This work presents a facial method to improve the ion conductivity of NASICON-type electrolytes and gives insight into the interface issues of solid-state sodium metal batteries.

1. Introduction

Sodium-ion batteries are recognized as promising next-generation energy storage technology due to their merits of low cost and an abundance of sodium resources compared to traditional lithium ion batteries [1,2]. Similar to conventional lithium-ion batteries, sodium-ion batteries use organic liquid electrolytes to show potential safety hazards such as combustion and explosion. Alternatively, solid-state sodium metal batteries employing solid-state electrolytes and sodium metal anodes can effectively solute the safety issues originating from organic liquid electrolytes and highly improve the energy density of devices [3,4]. In view of their excellent structural and chemical stability, high ionic conductivity (>10−4 S cm−1) and wide electrochemical window (>5 V), NASICON-type solid electrolytes have received abundant research attention in recent years [5,6]. As a typical NASICON structure, Na3Zr2Si2PO12 (NZSP) was firstly exploited in 1976 by Hong and Goodenough and reported widely owing to its relative higher ionic conductivity [7,8]. Nevertheless, the ionic conductivity of NZSP electrolyte is still lower than that of the liquid electrolyte organic (10−3–10−2 S cm−1) [9]. In addition, the poor contact between rigid NZSP electrolytes and electrodes results in large interfacial impedance and the growth of dendrites as well as other interfacial issues. Therefore, improving the ionic conductivity of NZSP electrolytes and effectively solving the interface problem are the top priorities for the practical application of NZSP in solid-state metal sodium batteries.
The ionic conductivity of a NASICON electrolyte is determined by several factors, such as the number and distribution of Na ions, phase structure as well as the relative density. The most commonly used strategy to improve the ion conductivity of a NZSP electrolyte is the partial substitution of Zr4+ in a Na3Zr2Si2PO12 structure by variable heterogeneous ions [10]. For example, Lu et al. reported that the mobility of Na+ at the Na3 site could be promoted by doping 5 at% Ca2+, leading to an enhanced ionic conductivity of 1.67 × 10−3 S cm−1 [11]. Yang et al. employed Mg2+ as the substitution ion for Zr4+ by the means of non-equivalent substitution to generate a deficiency in positive charge and then to increase the concentration of Na+, resulting in higher ion conductivity at 1.33 × 10−3 S cm−1 [12]. Zhang et al. introduced La3+ into a Zr4+ site to form a new Na3La(PO4)2 phase, which not only improved the Na+ concentration in the Na3Zr2Si2PO12 (3.4 × 10−3 S cm−1) but also increased the relative density [13]. Miao R et al. reported the influence of a Bi2O3 additive on the electrochemical performance of Na3.1Y0.1Zr1.9Si2PO12 inorganic solid electrolytes [14]. We summarize the ionic conductivities of NASICON electrolytes with different ionic dopants in Table 1.
In this study, the Bi2O3 with a low melting point (825 °C) was introduced into the sintering process of a NASICON (Na3Zr2Si2PO12) solid electrolyte so as to create liquid flux at high temperatures and improve the relative density of the electrolyte pellet. After the introduction of Bi2O3, the new phase of Na3Bi(PO4)2 was generated at the grain boundary, which altered the Si/P ratio of Na3Zr2Si2PO12 phase and further enhanced the Na+ occupancy. The optimized ionic conductivity of Na3Zr2Si2PO12-Bi is 1.27 × 10−3 S cm−1, which is much higher than that of pure Na3Zr2Si2PO12 (2.84 × 10−4 S cm−1). However, the direct contact of Na with Na3Zr2Si2PO12-Bi triggers the formation of a NaxBi alloy at the metal/electrolyte interface, which accelerates the dendrite growth of Na inside the electrolyte pellets. Such Na penetration into the electrolyte prompts the decreased electrochemical performance of the symmetrical battery and the solid sodium metal full battery. We believe the electrochemical performance of Na3Zr2Si2PO12-Bi-based solid sodium metal batteries can be further enhanced by a physical obstruction between Na metal and Na3Zr2Si2PO12-Bi electrolyte.

2. Materials and Methods

The Na3Zr2Si2PO12-xBi (Na3+xZr2-xMxSi2PO12, x = 0, 0.1, 0.2, 0.3 and 0.4) was synthesized by conventional high-temperature solid-state reaction. Stoichiometric amounts of Na2CO3, NH4H2PO4, ZrO2, SiO2 and Bi2O3 were mixed by high-energy ball milling for 1 h at a rotating speed of 1400 rpm and then calcinated at 1100 °C for 12 h. The 15%-excessive Na2CO3 and NH4H2PO4 were added to compensate for the volatilization of Na and P. After ball milling, the obtained mixture was further ball-milled for 1 h. The resultant dry powder was sieved through a 200-mesh sieve, then mixed with PVA binder and pressed into green pellets with a diameter of 16 mm and a thickness of 1 mm. The green pellets were pre-calcinated at 650 °C for 2 h to remove the PVA binder and finally calcinated at 1200 °C for 12 h to obtain dense electrolyte pellets.
The phase composition of Na3Zr2Si2PO12-xBi was confirmed by X-ray diffractometer (ARL Equinox 3000, Thermo Fisher, MA, USA). The microstructure was observed by scanning electron microscope (Quanta FEG, FEI, NY, USA) equipped with energy dispersive X-ray spectroscopy. The relative density of electrolyte pellets was measured by electronic densitometer (YD-100ES, YiEnDa, China) based on the Archimedes principle. The ionic and electronic conductivities were tested on electrochemical station (CHI 760E). The blocking electrode was constructed by sputtering Au on two sides of electrolyte pellets. The ionic conductivity was determined by electrochemical impedance spectroscopy (EIS, CHI 760E, ChenHua, Shanghai, China) technique ranging from 100 kHz to 0.1 Hz with an amplitude of 5 mV under open circuit potential. The electronic conductivity was determined by the DC polarization method at a constant voltage of 0.1–1.8 V for 1800 s at room temperature.
Sodium foils with a diameter of 8 mm were attached to both sides of the polished electrolyte sheet and then repetitively rolled to modify the interfacial contact. The Na/Na3Zr2Si2PO12/Na and Na/Na3Zr2Si2PO12-xBi/Na symmetric batteries were assembled into CR2032 coin cell to investigate the Na plating/stripping behaviors in an argon-filled glove box at room temperature. Critical current measurements were performed by the galvanostatic cycling at current densities stepping from 0.1–0.6 mA cm−2 (step: 0.05 mA cm−2, time: 0.5 h) at room temperature. The CR2032 full battery was assembled using sodium foil as the anode, NZSP pellet as the electrolyte and NVP as the cathode material in an argon-filled glove box at room temperature. Commercial Na3V2(PO4)3 (Hubei Energy Technology Co., Ltd.), acetylene black and poly(vinyldifluoride) (PVDF) with a weight ratio of 8:1:1 were mixed in N-methyl pyrolidone (NMP) to form a homogeneous slurry, ergo coated on aluminum foil by doctor blade and finally dried at 120 °C for 24 h in vacuum oven. The mass loading of NVP was ~1 mg cm−2. Ten μL of liquid electrolytes (1 M NaClO4 in EC/DMC (1:1) + 5% FEC) was added between cathode and electrolyte pellet to establish wetted interface.

3. Results and Discussion

Figure 1a displays the XRD pattern of solid electrolytes of Na3Zr2Si2PO12, which is sintered at 1200 °C. The diffraction peaks of Na3Zr2Si2PO12 can be owned as NASICON structures (PDF #84-1200), as shown in Figure 1b. A small amount of ZrO2 impurity can be found because of the excessive volatilization of Na and P during the high-temperature sintering process [13], while for Na3Zr2Si2PO12-xBi (x = 0, 0.1, 0.2, 0.3, and 0.4) samples, it is obvious that the Bi2O3 addition can facilitate the crystallization of Na3Zr2Si2PO12 and decrease the formation of ZrO2 impurities. With enhanced Bi2O3 contents, especially for x > 0.3, the intensity of the ZrO2 impurities gradually enhances. Moreover, the introduction of Bi2O3 induces the generation of the Na3Bi(PO4)2 phase, as confirmed by the characteristic peaks located at 21.5°, 25.5°, 32.2° and 33.6° (PDF #81-0336). Because the formation of Na3Bi(PO4)2 phase is cause by the solid-state reaction of Bi2O3, Na2CO3 and NH4H2PO4, the generation of the Na3Bi(PO4)2 phase can increase the Si/P ratio of the NASICON structure and enhance the Na+ occupancy, which may contribute to enhanced ionic conductivity. To investigate the origin of Na3Bi(PO4)2 formation, the solid-state reaction of Na2CO3, NH4H2PO4 and Bi2O3 was proceeded at 800 °C. As shown in Figure 1c, the XRD pattern can be indexed as the pure Na3Bi(PO4)2 phase (PDF #81-0036) [28]. Therefore, before the phase formation of Na3Zr2Si2PO12, parts of Bi2O3 were converted into Na3Bi(PO4)2, and the remaining Bi2O3 melted at a high temperature and was substituted for the Zr4+ site of Na3Zr2Si2PO12.
According to the Archimedes principle, the relative densities of Na3Zr2Si2PO12-xBi (x = 0, 0.1, 0.2, 0.3, 0.4) are 2.89, 3.07, 3.10, 3.08 and 3.02 g cm−3, corresponding to the relative densities of 89.2, 94.8, 95.7, 95.1 and 93.2 %, respectively (Table 1). To investigate the effects of the Bi2O3 addition on the morphology of Na3Zr2Si2PO12 electrolytes, scanning electron microscopy (SEM) has been performed on the Na3Zr2Si2PO12-xBi (x = 0, 0.1, 0.2, 0.3, 0.4) samples, as shown in Figure 2. The grain size of the initial Na3Zr2Si2PO12 sample is large and variable, and there are plenty of holes between the grains. The relative density of a Na3Zr2Si2PO12 pellet is 89.2 %. When x = 0.1 and 0.2, the grain size of Na3Zr2Si2PO12 particles becomes smaller and uniform, and there are no obvious gaps or holes observed. The highest relative density for a Na3Zr2Si2PO12-0.2Bi sample is 95.7%. As is known, Bi2O3 is a typical low melting point oxide (825 °C), which forms liquid flux at high temperatures and bonds the grain together, enabling enhanced crystallization and densification rates during sintering [28]. When x = 0.3 and 0.4, the grain size of Na3Zr2Si2PO12 particles further decreases, and thereupon, a small number of gaps and holes reappear between the grains, which weakens the density.
Figure 3a displays the EIS profiles of Na3Zr2Si2PO12-xBi (x = 0, 0.1, 0.2, 0.3, 0.4) at room temperature. The EIS curve consists of a semicircle in the high-frequency region and an inclined line in the low-frequency region [28]. The intersection of the semicircle’s destination and the inclined line (Z’-intercept) is the total resistance of an electrolyte pellet. After Bi2O3 introduction, the total resistance and the radius of the semicircle in the high-frequency region decrease obviously, demonstrating a decreased total resistance. Moreover, the total resistance is composed of grain resistance (Rg) and grain boundary resistance (Rgb) [29]. The Rg and Rgb of Na3Zr2Si2PO12-xBi samples are obtained by fitting the EIS curve (Figure 3b), and the corresponding results have been summarized in Table 2.
The total ionic conductivities of Na3Zr2Si2PO12-xBi (x = 0, 0.1, 0.2, 0.3, 0.4) are 2.84 × 104, 9.04 × 104, 1.27 × 103, 9.92 × 104 and 6.92 × 104, respectively. When x ≤ 0.2, the grain boundary conductivity grows significantly faster than the grain conductivity. When x = 0.2, the grain boundary conductivity (σgb = 3.32 × 103 S cm1) exceeds the grain conductivity (σb = 2.05 × 103 S cm1). The increasing of the grain conductivity originates from the enhanced density and the generation of Na3Bi(PO4)2 phase, which increases the Si/P ratio and Na+ occupancy of Na3Zr2Si2PO12 [30,31]. The increasing of the grain boundary conductivity is attributed to the reduction of gaps and pores between grains. when x ≥ 0.2, σg and σgb simultaneously decrease, and the σgb is smaller than σg. The reason why σgb is smaller than σg is the formation of insulating ZrO2 impurity, which hinders the Na+ transport between Na3Zr2Si2PO12 grains [32]. Figure 3c shows the i-t curves of Na3Zr2Si2PO12 and Na3Zr2Si2PO12-0.2Bi pellets under a fixed voltage of 1 V. The measured steady-state currents are 8.48 × 10−7 A for Na3Zr2Si2PO12 and 7.35 × 10−7 A for Na3Zr2Si2PO12-0.2Bi. Correspondingly, the electronic conductivities at room temperature are calculated to be 6.67 × 10−8 S cm−1 for Na3Zr2Si2PO12 and 5.23 × 10−8 S cm−1 for Na3Zr2Si2PO12-0.2Bi.
To investigate the effects of the Bi2O3 addition on the Na+ transport carrier of Na3Zr2Si2PO12 electrolytes, the EIS plots of Na3Zr2Si2PO12 and Na3Zr2Si2PO12-0.2Bi at different temperatures have been measured, as shown in Figure 4a,b,d,e. Figure 4c,f depicts the Arrhenius curve of Na3Zr2Si2PO12 and Na3Zr2Si2PO12-0.2Bi, respectively. The activation energy Ea of the Na3Zr2Si2PO12-0.2Bi sample at 228–373 K is 0.28 eV, which is much smaller than that of Na3Zr2Si2PO12 (0.31 eV), which is in line with the previously reported Ea value (0.26–0.40 eV) [32]. The reduced activation energy of the Na3Zr2Si2PO12-0.2Bi samples suggests that the addition of Bi2O3 is feasible for Na+ transport [30].
To investigate the Na plating/stripping behavior, the Na/Na3Zr2Si2PO12/Na and Na/Na3Zr2Si2PO12-0.2Bi/Na symmetrical batteries were assembled. Critical current density (CCD) is a critical parameter to evaluate the tolerance of Na dendrite growth [33]. As shown in Figure 5a,b, the CCD of the Na/Na3Zr2Si2PO12/Na symmetric cell is 0.35 mA cm−2, which is slightly higher than that of Na/Na3Zr2Si2PO12-0.2Bi/Na (0.30 mA cm−2). The decrease of CCD after Bi2O3 introduction originates from the spontaneous reaction of Na with Bi3+ of Na3Zr2Si2PO12-0.2Bi, which would induce the formation of a Na-Bi alloy and thereupon accelerate the dendrite growth [34,35,36]. After CCD tests, the symmetrical batteries were disassembled (Figure 5c). Conspicuously, the large area of the black region can be observed. The Na/Na3Zr2Si2PO12-0.2Bi/Na pellet was then ground into powder, and the corresponding XRD pattern can further confirm the existence of trace amount of Bi metal (Figure 5d).
Finally, we assembled solid sodium full batteries, using Na3Zr2Si2PO12 and Na3Zr2Si2PO12-0.2Bi as an electrolyte and commercial Na3V2(PO4)3 as a cathode. For Na3V2(PO4)3/Na3Zr2Si2PO12/Na and Na3V2(PO4)3/Na3Zr2Si2PO12-0.2Bi/Na, the initial discharging capacities were 96.8 mA h g−1 and 96.1 mA h g−1, with initial coulombic efficiencies of 91.1% and 94.1%, respectively (Figure 5e). After 100 cycles, Na3V2(PO4)3/Na3Zr2Si2PO12/Na and Na3V2(PO4)3/Na3Zr2Si2PO12-0.2Bi/Na maintain reversible capacities of 95.5 mA h g−1 and 92.4 mA h g−1 with capacity retentions of 98.66% and 96.15%, respectively. Although the addition of Bi2O3 into an Na3Zr2Si2PO12 electrolyte can increase the relative density/ionic conductivity and decrease the electronic conductivity, it does not manifest a positive effect on the electrochemical performance of symmetrical and full batteries due to the formation of Na-Bi alloys at the Na/electrolyte interface. Such Na-Bi alloys would promote the Na dendrite growth and further penetrate through the electrolyte. However, due to its benefits of enhanced ionic conductivity, decreased electronic conductivity and bulk density, we believe the extra interfacial modifying layer on a Na3Zr2Si2PO12-0.2Bi surface, which acts as a physical shield so as to avoid the direct contact of Na with Bi3+, can improve the electrochemical performance of Na3Zr2Si2PO12-0.2Bi-based solid sodium metal batteries.

4. Conclusions

In this work, the Na3Zr2Si2PO12 electrolyte is prepared by a Bi2O3-assisted high-temperature solid-state reaction. At high temperatures, liquid Bi2O3 flux can prompt a relative dense microstructure of Na3Zr2Si2PO12 by bonding the grain. Meanwhile, the presence of Na3Bi(PO4)2 impurities can enhance the Si/P ratio of Na3Zr2Si2PO12 and the Na+ occupancy, thus increasing the ionic conductivity of the Na3Zr2Si2PO12 electrolyte. The optimized Na3Zr2Si2PO12-0.2Bi electrolyte manifests an ionic conductivity of 1.27 × 10−3 S cm−1 at room temperature and a low activation energy of 0.28 eV. Despite these, due to the spontaneous reaction of Na and Bi3+ when assembling solid-state symmetrical and full batteries, the Bi2O3 introduction cannot further enhance the electrochemical performance.

Author Contributions

Conceptualization, S.C. and W.M.; methodology, X.X. and Y.Z.; writing—review and editing, Z.M. and D.W.; supervision, J.C. and C.D.; All authors have read and agreed to the published version of the manuscript.

Funding

This research was financial supported by the National Natural Science Foundation of China (Grant No. 52202282) and Scientific Developing Foundation of Tianjin Education Commission (Grant No. 2018ZD09).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yabuuchi, N.; Kubota, K.; Dahbi, M.; Komaba, S. Research Development on Sodium-Ion Batteries. Chem. Rev. 2014, 114, 11636–11682. [Google Scholar] [CrossRef] [PubMed]
  2. Chen, S.; Wu, C.; Shen, L.; Zhu, C.; Huang, Y.; Xi, K.; Maier, J.; Yu, Y. Challenges and Perspectives for NASICON-Type Electrode Materials for Advanced Sodium-Ion Batteries. Adv. Mater. 2017, 29, 1700431. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, G.; Sun, X.; Yu, X.; Weng, W.; Yang, J.; Zhou, D.; Xiao, R.; Chen, L.; Yao, X. Na10SnSb2S12: A nanosized air-stable solid electrolyte for all-solid-state sodium batteries. Chem. Eng. J. 2021, 420, 127692. [Google Scholar] [CrossRef]
  4. Wang, Y.; Richards, W.D.; Ong, S.P.; Miara, L.J.; Kim, J.C.; Mo, Y.; Ceder, G. Design principles for solid-state lithium superionic conductors. Nat. Mater. 2015, 14, 1026–1031. [Google Scholar] [CrossRef]
  5. Gandi, S.S.; Gandi, S.; Madduluri, V.R.; Katari, N.K.; Dutta, D.P.; Ravuri, B.R. Na3+x[CrxTi2−x(PO4)3] glass-ceramic electrolyte: Ionic conductivity and structural correlations for different heat treating temperatures and time schedules. Ionics 2019, 25, 4179–4188. [Google Scholar] [CrossRef]
  6. Leng, H.; Nie, J.; Luo, J. Combining cold sintering and Bi2O3-Activated liquid-phase sintering to fabricate high-conductivity Mg-doped NASICON at reduced temperatures. J. Mater. 2019, 5, 237–246. [Google Scholar] [CrossRef]
  7. 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]
  8. Goodenough, J.B.; Hong, H.Y.P.; Kafalas, J.A. Fast Na+-ion transport in skeleton structures. Mater. Res. Bull. 1976, 11, 203–220. [Google Scholar] [CrossRef]
  9. Kato, Y.; Hori, S.; Saito, T.; Suzuki, K.; Hirayama, M.; Mitsui, A.; Yonemura, M.; Iba, H.; Kanno, R. High-power all-solid-state batteries using sulfide superionic conductors. Nat. Energy 2016, 1, 16030. [Google Scholar] [CrossRef]
  10. Sundar, G.S.; Suman, G.; Kumar, K.N.; Dutta, D.P.; Rao, R.B. Investigation on the applicability of high Na-ion conducting Na3+x[ZrxSc2-x(PO4)3] glass-ceramic electrolyte for use in safer Na-ion batteries. J. Phys. Chem. Solids 2019, 126, 209–218. [Google Scholar] [CrossRef]
  11. Lu, Y.; Alonso, J.A.; Yi, Q.; Lu, L.; Wang, Z.L.; Sun, C. A High-Performance Monolithic Solid-State Sodium Battery with Ca2+ Doped Na3Zr2Si2PO12 Electrolyte. Adv. Energy Mater. 2019, 9, 1901205. [Google Scholar] [CrossRef]
  12. Yang, J.; Wan, H.-L.; Zhang, Z.-H.; Liu, G.-Z.; Xu, X.-X.; Hu, Y.-S.; Yao, X.-Y. NASICON-structured Na3.1Zr1.95Mg0.05Si2PO12 solid electrolyte for solid-state sodium batteries. Rare Met. 2018, 37, 480–487. [Google Scholar] [CrossRef]
  13. Zhang, Z.; Zhang, Q.; Shi, J.; Chu, Y.S.; Yu, X.; Xu, K.; Ge, M.; Yan, H.; Li, W.; Gu, L.; et al. A Self-Forming Composite Electrolyte for Solid-State Sodium Battery with Ultralong Cycle Life. Adv. Energy Mater. 2017, 7, 1601196. [Google Scholar] [CrossRef]
  14. Miao, R.J.; Cao, X.G.; Wang, W.G.; Zhang, H.Y. Influence of Bi2O3 additive on the electrochemical performance of Na3.1Y0.1Zr1.9Si2PO12 inorganic solid electrolyte. Ceram. Int. 2021, 47, 17455–17462. [Google Scholar] [CrossRef]
  15. Chen, D.; Luo, F.; Zhou, W.; Zhu, D. Influence of Nb5+, Ti4+, Y3+ and Zn2+ doped Na3Zr2Si2PO12 solid electrolyte on its conductivity. J. Alloy. Compd. 2018, 757, 348–355. [Google Scholar] [CrossRef]
  16. Khakpour, Z. Influence of M: Ce4+, Gd3+ and Yb3+ substituted Na3+xZr2−xMxSi2PO12 solid NASICON electrolytes on sintering, microstructure and conductivity. Electrochim. Acta 2016, 196, 337–347. [Google Scholar] [CrossRef]
  17. Jolley, A.G.; Cohn, G.; Hitz, G.T.; Wachsman, E.D. Improving the ionic conductivity of NASICON through aliovalent cation substitution of Na3Zr2Si2PO12. Ionics 2015, 21, 3031–3038. [Google Scholar] [CrossRef]
  18. Huang, C.C.; Yang, G.M.; Yu, W.H.; Xue, C.; Zhai, Y.F.; Tang, W.P.; Hu, N.; Wen, Z.Y.; Lu, L.; Song, S.F. Gallium-substituted Nasicon Na3Zr2Si2PO12 solid electrolytes. J. Alloy. Compd. 2021, 855, 157501. [Google Scholar] [CrossRef]
  19. Ma, Q.L.; Guin, M.; Naqash, S.; Tsai, C.L.; Tietz, F.; Guillon, O. Scandium-Substituted Na3Zr2(SiO4)2(PO4) Prepared by a Solution Assisted Solid-State Reaction Method as Sodium-Ion Conductors. Chem. Mater. 2016, 28, 4821–4828. [Google Scholar] [CrossRef]
  20. Song, S.F.; Duong, H.M.; Korsunsky, A.M.; Hu, N.; Lu, L. A Na+ Superionic Conductor for Room-Temperature Sodium Batteries. Sci. Rep. 2016, 6, 1–10. [Google Scholar] [CrossRef]
  21. Wang, X.; Mei, W.; Chen, J.; Wang, D.; Mao, Z. Rare Earth Oxide-Assisted Sintered NASICON Electrolyte Composed of a Phosphate Grain Boundary Phase with Low Electronic Conductivity. ACS Appl. Energy Mater. 2022, 5, 777–783. [Google Scholar] [CrossRef]
  22. Thirupathi, R.; Omar, S. A Strategic Co-doping Approach Using Sc3+ and Ce4+ toward Enhanced Conductivity in NASICON-Type Na3Zr2Si2PO12. J. Phys. Chem. C 2021, 125, 27723–27735. [Google Scholar] [CrossRef]
  23. Heo, E.; Wang, J.E.; Yun, J.H.; Kim, J.-H.; Kim, D.J.; Kim, D.K. Improving Room Temperature Ionic Conductivity of Na3–xKxZr2Si2PO12 Solid-Electrolytes: Effects of Potassium Substitution. Inorg. Chem. 2021, 60, 11147–11153. [Google Scholar] [CrossRef] [PubMed]
  24. Ma, Q.L.; Tsai, C.L.; Wei, X.K.; Heggen, M.; Tietz, F.; Irvine, J.T.S. Room temperature demonstration of a sodium superionic conductor with grain conductivity in excess of 0.01 S cm−1 and its primary applications in symmetric battery cells. J. Mater. Chem. A 2019, 7, 7766–7776. [Google Scholar] [CrossRef] [Green Version]
  25. Shen, L.; Yang, J.; Liu, G.; Avdeev, M.; Yao, X. High ionic conductivity and dendrite-resistant NASICON solid electrolyte for all-solid-state sodium batteries. Mater. Today Energy 2021, 20, 100691. [Google Scholar] [CrossRef]
  26. Oh, J.S.; He, L.; Plewa, A.; Morita, M.; Zhao, Y.; Sakamoto, T.; Song, X.; Zhai, W.; Zeng, K.; Lu, L. Composite NASICON (Na3Zr2Si2PO12) Solid-State Electrolyte with Enhanced Na+ Ionic Conductivity: Effect of Liquid Phase Sintering. ACS Appl. Mater. Interfaces 2019, 11, 40125–40133. [Google Scholar] [CrossRef]
  27. Shao, Y.J.; Zhong, G.M.; Lu, Y.X.; Liu, L.L.; Zhao, C.L.; Zhang, Q.Q.; Hu, Y.S.; Yang, Y.; Chen, L.Q. A novel NASICON-based glass-ceramic composite electrolyte with enhanced Na-ion conductivity. Energy Storage Mater. 2019, 23, 514–521. [Google Scholar] [CrossRef]
  28. Diouri, M.; Sadel, A.; Zahir, M.; Drache, M.; Conflant, P.; Wignacourt, J.P.; Boivin, J.C. Na3Bi(PO4)2 type solid solutions; investigation of structural and electrical properties. J. Alloy. Compd. 1992, 188, 206–210. [Google Scholar] [CrossRef]
  29. Cooper, K.R. Progress Toward Accurate Through-Plane Ion Transport Resistance Measurement of Thin Solid Electrolytes. J. Electrochem. Soc. 2010, 157, B1731. [Google Scholar] [CrossRef] [Green Version]
  30. Wang, X.; Liu, Z.; Tang, Y.; Chen, J.; Wang, D.; Mao, Z. Low temperature and rapid microwave sintering of Na3Zr2Si2PO12 solid electrolytes for Na-Ion batteries. J. Power Sources 2021, 481, 228924. [Google Scholar] [CrossRef]
  31. Sun, F.; Xiang, Y.; Sun, Q.; Zhong, G.; Banis, M.N.; Li, W.; Liu, Y.; Luo, J.; Li, R.; Fu, R.; et al. Insight into Ion Diffusion Dynamics/Mechanisms and Electronic Structure of Highly Conductive Sodium-Rich Na3+xLaxZr2−xSi2PO12 (0 ≤ x ≤ 0.5) Solid-State Electrolytes. ACS Appl. Mater. Interfaces 2021, 13, 13132–13138. [Google Scholar] [CrossRef] [PubMed]
  32. Fuentes, R.O.; Figueiredo, F.; Marques, F.M.B.; Franco, J.I. Reaction of NASICON with water. Solid State Ion. 2001, 139, 309–314. [Google Scholar] [CrossRef]
  33. He, S.; Xu, Y.; Chen, Y.; Ma, X. Enhanced ionic conductivity of an F−-assisted Na3Zr2Si2PO12 solid electrolyte for solid-state sodium batteries. J. Mater. Chem. A 2020, 8, 12594–12602. [Google Scholar] [CrossRef]
  34. Lu, Y.; Zhao, C.-Z.; Yuan, H.; Cheng, X.-B.; Huang, J.-Q.; Zhang, Q. Critical Current Density in Solid-State Lithium Metal Batteries: Mechanism, Influences, and Strategies. Adv. Funct. Mater. 2021, 31, 2009925. [Google Scholar] [CrossRef]
  35. Fang, W.; Fan, L.; Zhang, Y.; Zhang, Q.; Yin, Y.; Zhang, N.; Sun, K. Synthesis of carbon coated Bi2O3 nanocomposite anode for sodium-ion batteries. Ceram. Int. 2017, 43, 8819–8823. [Google Scholar] [CrossRef]
  36. Xu, R.; Kinderman, R.; Van der Lugt, W. Electrical resistivities of liquid Na-Bi and Rb-Bi alloys. J. Phys. Condens. Matter 1991, 3, 127–133. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of (a) Na3Zr2Si2PO12, (b) Na3Zr2Si2PO12-xBi (x = 0, 0.1, 0.2, 0.3 and 0.4) and (c) Na3Bi(PO4)2.
Figure 1. XRD patterns of (a) Na3Zr2Si2PO12, (b) Na3Zr2Si2PO12-xBi (x = 0, 0.1, 0.2, 0.3 and 0.4) and (c) Na3Bi(PO4)2.
Coatings 12 01774 g001
Figure 2. Top-view SEM images of Na3Zr2Si2PO12 samples with different Bi3+ concentrations: (a) zero, (b) 0.1, (c) 0.2, (d) 0.3 and (e) 0.4.
Figure 2. Top-view SEM images of Na3Zr2Si2PO12 samples with different Bi3+ concentrations: (a) zero, (b) 0.1, (c) 0.2, (d) 0.3 and (e) 0.4.
Coatings 12 01774 g002
Figure 3. (a) Nyquist plots (b) Grain resistance, grain boundary resistance and total resistance of Na3Zr2Si2PO12-xBi (x = 0, 0.1, 0.2, 0.3, 0.4) samples at room temperature. (c) Polarization current−time curves at 1 V of Na3Zr2Si2PO12 and Na3Zr2Si2PO12-0.2Bi.
Figure 3. (a) Nyquist plots (b) Grain resistance, grain boundary resistance and total resistance of Na3Zr2Si2PO12-xBi (x = 0, 0.1, 0.2, 0.3, 0.4) samples at room temperature. (c) Polarization current−time curves at 1 V of Na3Zr2Si2PO12 and Na3Zr2Si2PO12-0.2Bi.
Coatings 12 01774 g003
Figure 4. (a,b) Low-temperature, room-temperature and high-temperature Nyquist plots of Na3Zr2Si2PO12 electrolyte. (c) Arrhenius plot of Na3Zr2Si2PO12 electrolyte. (d,e) Low-temperature, room-temperature and high-temperature Nyquist plots of Na3Zr2Si2PO12- electrolyte. (f) Arrhenius plot of Na3Zr2Si2PO12-0.2Bi electrolyte.
Figure 4. (a,b) Low-temperature, room-temperature and high-temperature Nyquist plots of Na3Zr2Si2PO12 electrolyte. (c) Arrhenius plot of Na3Zr2Si2PO12 electrolyte. (d,e) Low-temperature, room-temperature and high-temperature Nyquist plots of Na3Zr2Si2PO12- electrolyte. (f) Arrhenius plot of Na3Zr2Si2PO12-0.2Bi electrolyte.
Coatings 12 01774 g004
Figure 5. Critical current density of (a) Na/Na3Zr2Si2PO12/Na and (b) Na/Na3Zr2Si2PO12-0.2Bi/Na symmetric cells at different current densities at room temperature. (c) Disassembled Na/Na3Zr2Si2PO12-0.2Bi/Na symmetrical battery and (d) corresponding XRD pattern of Na/Na3Zr2Si2PO12-0.2Bi/Na symmetrical battery after grinding. (e) Cycling performance of Na3V2(PO4)3/Na3Zr2Si2PO12/Na and Na3V2(PO4)3/Na3Zr2Si2PO12-0.2Bi/Na solid sodium full battery.
Figure 5. Critical current density of (a) Na/Na3Zr2Si2PO12/Na and (b) Na/Na3Zr2Si2PO12-0.2Bi/Na symmetric cells at different current densities at room temperature. (c) Disassembled Na/Na3Zr2Si2PO12-0.2Bi/Na symmetrical battery and (d) corresponding XRD pattern of Na/Na3Zr2Si2PO12-0.2Bi/Na symmetrical battery after grinding. (e) Cycling performance of Na3V2(PO4)3/Na3Zr2Si2PO12/Na and Na3V2(PO4)3/Na3Zr2Si2PO12-0.2Bi/Na solid sodium full battery.
Coatings 12 01774 g005
Table 1. The ionic conductivities of previously reported NASICON electrolytes with different dopants at room temperature.
Table 1. The ionic conductivities of previously reported NASICON electrolytes with different dopants at room temperature.
ComponentsIonic Conductivity (S cm−1)Reference
Na3Zr1.9Ca0.1Si2PO121.67 × 10−3[11]
Na3Zr1.7La0.3Si2PO123.4 × 10−3[13]
Na3.1Y0.1Zr1.9Si2PO12 + 1 wt.% Bi2O31.21 × 10−3[14]
Na3Zr1.9Nb0.1Si2PO122.11 × 10−4 [15]
Na3Zr1.9Ti0.1Si2PO125.97 × 10−4[15]
Na3Zr1.8Zn0.2Si2PO122.50 × 10−3[15]
Na3Zr1.9Yb0.1Si2PO121.7 × 10−4[16]
Na3Zr1.9Gd0.1Si2PO126.0 × 10−4[16]
Na3Zr1.9Ce0.1Si2PO129.0 × 10−4[16]
Na3Zr1.9Al0.1Si2PO124.39 × 10−4[17]
Na3Zr1.9Fe0.1Si2PO127.53 × 10−4[17]
Na3Zr1.9Co0.1Si2PO121.55 × 10−4[17]
Na3Zr1.9Ni0.1Si2PO126.18 × 10−4[17]
Na3Zr1.9Ga0.1Si2PO121.06 × 10−3[18]
Na3.4Zr1.6Sc0.4Si2PO124.0 × 10−3[19]
Na3.1Zr1.95Mg0.05Si2PO123.5 × 10−3[20]
Na3.2Zr1.8Sm0.2Si2PO121.36 × 10−3[21]
Na3.3Zr1.7Ho0.3Si2PO121.07 × 10−3[21]
Na3.33Ce0.02Sc0.33Zr1.65Si2PO122.44 × 10−3[22]
Na2.9K0.1Zr2Si2PO127.734 × 10−4[23]
Na3.4Zr2Si2.4P0.6O125.5 × 10−3[24]
Na3.4Zr1.9Mg0.1Si2.2P0.8O123.6 × 10−3[25]
Na3Zr2Si2PO12 + 5 wt.% Na2SiO31.45 × 10−3[26]
Na3.2Zr2Si2.2P0.8O12 + 0.5 NaF3.6 × 10−3[27]
Table 2. Density and conductivity of Na3Zr2Si2PO12-xBi samples (x = 0, 0.1, 0.2, 0.3 and 0.4) at room temperature.
Table 2. Density and conductivity of Na3Zr2Si2PO12-xBi samples (x = 0, 0.1, 0.2, 0.3 and 0.4) at room temperature.
NZSP-xBiρ
(g cm−3)
Rg
(Ω)
Rgb
(Ω)
σb
(S cm−1)
σgb
(S cm−1)
σtotal
(S cm−1)
x = 02.8995.0146.07.20 × 1044.69 × 1042.84 × 104
x = 0.13.0740.046.01.95 × 1031.69 × 1039.04 × 104
x = 0.23.1038.423.72.05 × 1033.32 × 1031.27 × 103
x = 0.33.0842.544.52.03 × 1031.94 × 1039.92 × 104
x = 0.43.0250.354.71.45 × 1031.33 × 1036.92 × 104
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Cen, S.; Mei, W.; Xing, X.; Zeng, Y.; Mao, Z.; Wang, D.; Chen, J.; Dong, C. Bi2O3-Assisted Sintering of Na3Zr2Si2PO12 Electrolyte for Solid-State Sodium Metal Batteries. Coatings 2022, 12, 1774. https://doi.org/10.3390/coatings12111774

AMA Style

Cen S, Mei W, Xing X, Zeng Y, Mao Z, Wang D, Chen J, Dong C. Bi2O3-Assisted Sintering of Na3Zr2Si2PO12 Electrolyte for Solid-State Sodium Metal Batteries. Coatings. 2022; 12(11):1774. https://doi.org/10.3390/coatings12111774

Chicago/Turabian Style

Cen, Shangxu, Wentao Mei, Xiangyuan Xing, Yiwei Zeng, Zhiyong Mao, Dajian Wang, Jingjing Chen, and Chenlong Dong. 2022. "Bi2O3-Assisted Sintering of Na3Zr2Si2PO12 Electrolyte for Solid-State Sodium Metal Batteries" Coatings 12, no. 11: 1774. https://doi.org/10.3390/coatings12111774

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