Elsevier

Solid State Ionics

Volume 369, 15 October 2021, 115712
Solid State Ionics

Characterization of hot-pressed von Alpen type NASICON ceramic electrolytes

https://doi.org/10.1016/j.ssi.2021.115712Get rights and content

Highlights

  • Densified spray dried von Alpen NASICON powders using hot-pressing.

  • Smaller and more uniform spray dried particles and hot-pressing reduce glassy phase.

  • First experimental values of Poisson's ratio of NASICON, between 0.23 and 0.26.

  • First experimental values of shear modulus of NASICON, between 33.87 and 38.79 GPa.

  • Hot-pressed NASICON total conductivities ranged from 0.292 to 0.595 mS/cm.

Abstract

The Na-ion conducting solid electrolyte NASICON is a promising solid state electrolyte (SSE) for use in Na based electrochemical systems. The effects of particle morphology and synthesis of von Alpen type NASICON via hot-pressing were investigated. Spray dried particles were shown to improve mechanical and electrochemical properties. X-ray diffraction, scanning electron microscopy, impulse excitation measurements, Vickers hardness, and impedance spectroscopy were used to correlate relationships between phase purity, microstructure, mechanical properties, and electrochemical properties. Scanning electron microscopy showed the presence of a glass phase as between 19.08% and 28.23% area percentage and a ZrO2 phase between 0.46% and 3.51% area percentage in the hot-pressed samples. Impulse excitation showed elastic modulus varying from 83.65 to 97.65 GPa. The first values of Poisson's ratio and shear modulus of NASICON are reported as between 0.23 and 0.26 and 33.87 to 38.79 GPa, respectively. Vickers hardness and fracture toughness measurements showed Vickers hardness ranging from 4.30 to 4.68 GPa and fracture toughness ranging from 1.05 to 1.09 MPa m½. Impedance spectroscopy showed grain conductivities ranging from 1.023 to 1.287 mS/cm, with grain boundary resistance percentages between 73.22 and 47.5%. Total conductivities ranged from 0.292 to 0.596 mS/cm, with 0.596 mS/cm being achieved at room temperature for the spray dried powders with lower grain boundary resistance percentages.

Introduction

Li-ion conducting solid state electrolytes (SSE) have garnered much interest as a possible solution to limits faced by liquid electrolytes traditionally used in Li ion batteries. Much of this interest for Li based SSE stems from the improved safety solid electrolytes offer compared to non-aqueous liquid electrolytes as well as their ability to be used in higher energy density cells when paired with Li metal anodes [[1], [2], [3]]. While there have been a number of Li based SSE systems which have shown promise in enabling the use of Li metal anodes, serious concerns remain around the quantity of available Li and conflict associated with accessing remaining sources of available Li [[4], [5], [6], [7]].

Na based SSEs offer the same form of improvements to safety and energy density that Li based SSEs do compared to Li ion batteries, while having the added benefit of allowing for pairing with Na metal anodes [8]. Na metal has a higher propensity to creep than Li, which can result in higher charge rate capabilities when compared to Li [[9], [10], [11]]. While Li-based battery technology has significantly advanced in recent decades, the cost and availability of Li could be potential issues for application such as grid storage [12]. Sodium has a higher abundance than Li, is more affordable, and presents a possibility for fewer conflicts in sourcing [13].

Alongside the need for higher performance batteries for electric vehicles comes a need for improved grid energy storage. Flow cells are one format of electrochemical cells which show promise as grid energy storage systems, however a need remains for effective, low-cost single-ion conducting membranes to prevent cross contamination [14]. SSEs also show promise as membranes for flow cells, as there is much overlap between the engineering requirements for SSEs and flow cell membranes. One of the most promising Na based SSEs have been the Na super-ion conductor (NASICON) materials of the form NaxM2(XO4)3. NASICON materials, particularly variations of Na3Zr2Si2PO12 (NSZP), have long been investigated as solid electrolytes due to their high ionic conductivity [[15], [16], [17]]. These NASICON materials have been used commercially in molten sodium batteries, such as sodium sulfur and ZEBRA cells operating near 300 °C [18]. Hong originally described NASICON formulations based on Na1+xZr2SixP3-xO12 consisting of a solid-solution series between NaZr2P3O12 and Na4Zr2Si3O12 in 1976 [15]. Due to the presence of zirconia formation upon sintering, von Alpen suggested a zirconia-deficient form of NASICON with the general formula Na1+xZr2−x/3SixP3−xO12−2x/3 with end members of NaZr2P3O12 and Na4ZrSi3O10 [16]. Kuriakose et al. and Ahmad et al. analyzed the effects of sintering both Hong and von Alpen type NASICON samples, observing that von Alpen samples typically did not form the secondary phase zirconia upon sintering, while Hong samples did contain zirconia [19,20].

NASICON materials also have demonstrated effectiveness at operating as membranes in aqueous flow cells [21]. However, the ability to densify and manufacture thin films of NASICON remains a challenge. The processing methods used to create NASICON have been shown to have drastic impacts on the microstructure, mechanical, and electrochemical properties of NASICON [[22], [23], [24]]. While much work has been done to investigate the effects of various NASICON synthesis methods [19,[23], [24], [25], [26], [27]] and various sintering conditions [20,24,[28], [29], [30]], comparatively few studies have been performed on the effects of hot pressing on NASICON NZSP samples. Perthuis observed that hot pressing Hong-type NASICON at 1250C resulted in highly dense (98% theoretical density) and conductive (1mS/cm at room temperature) NASICON samples, however the effects of hot pressing on phase purity, microstructural evolution, and mechanical properties were not described [23]. Yde-Andersen hot pressed varied NASICON compositions prepared from gel-derived raw materials and obtained high densities (~95–99% theoretical density) and varied conductivities (~0.30–2.94 mS/cm at room temperature), however the effects on mechanical properties and phase purity, including the presence of a glassy phase, were only minimally described and effects on microstructural evolution were not described [31]. Moreover, the presence of the glassy impurity phase has been believed to affect the NASICON bulk phase composition and therefore the bulk properties [20,25]. In addition, the glassy phase has been observed to be present at grain boundaries, which likely affects ionic transport, stability, and perhaps mechanical properties. Comparatively few studies have also investigated mechanical properties of NASICON samples. Nonemacher investigated the elastic modulus (72–88 GPa), hardness (5.6–7.6 GPa), and fracture toughness (1.30–1.58 MPa m0.5) of Al and Y substituted NASICON solid electrolytes [32]. Fracture strengths ranging from 55 to 212 MPa have been found using three point bend, four point bend, and diametral strength tests [24,28,31,33].

In this study, we hypothesize that a combination of optimized hot-pressing conditions and powder processing could enable a processing route to produce a dense NASICON with a reduced glass content compared to conventional sintering. Furthermore, there is very limited information on the mechanical properties of VA NASICON. Hot-pressing has previously been shown to produce ceramic samples enabling reliable testing of mechanical properties such as elastic modulus, hardness, and fracture toughness [[34], [35], [36], [37]]. Thus, we hypothesize that hot-pressing can be used to prepare dense NASICON samples that allow for increased understanding of the mechanical properties of VA NASICON. Powder x-ray diffraction (XRD), scanning electron microscopy (SEM), energy-dispersive x-ray spectroscopy (EDS), impulse excitation, Vickers hardness, indentation fracture toughness, and electrochemical impedance spectroscopy (EIS) were used to investigate the effects of hot-pressing von Alpen type NASICON powders. Von-Alpen type NASICON was selected due to its lower propensity to form zirconia as compared to Hong type NASICON, which leads to higher conductivity and fracture strength [38]. Von Alpen NASICON powder synthesis was varied to alter particle morphologies to see how this effects the microstructure and properties of the hot-pressed samples.

Section snippets

Materials synthesis and processing

Von Alpen (VA) type NASICON (Na3.1Zr1.55Si2.3P0.7O11) powders were prepared through the following process. Reagent grade Na3PO4·12H2O, Na2CO3, ZrO2 and SiO2 were used as starting materials. The starting materials were mixed into ethanol. The resulting mixture was ball milled (Pulverisette 5, Fritsch) at 400 rpm for 0.5 h. After ball milling it mixture was dried at 80 °C in an oven for 24 h. After drying, the resulting powder was calcined at 1100 °C for 10 h. The calcined powder was ball milled

Materials characterization

SEM images of NASICON powders were taken for the three different powders used prior to hot pressing. Fig. 1 shows the particle morphology for NS, S, and S900C powders. A significant difference in particle morphology can be observed between NS and S powders prior to hot pressing. The NS particles consisted of irregular and larger particles while the S powders had a more consistent spherical morphology, in line with [39]. After undergoing heat treatment, S900C samples did not undergo any

Conclusions

A RIHP technique was used to synthesize dense von Alpen NASICON-type materials using three different precursor powders. Changes in particle morphology resulted in differences in phase purity, microstructure, mechanical properties, and electrochemical properties. The formation of a secondary, glassy phase impurity along the grain boundaries was more prevalent in NS precursor powders with larger, more irregular particles. The heightened presence of the glassy phase served to decrease the elastic

Author statement

Jeff Sakamoto and Joey Valle: Conceptualization the study. Joey Valle conducted materials densification, characterization, and EIS measurements. Jeff Wolfenstine interpreted data and provided experience and expertise in writing the manuscript. Wooseok Go synthesized the NaSICON powder. Youngsik Kim supervised Wooseok Go's synthetic work. Claire Huang conducted the mechanical property measurement. Dhruv Tatke assisted in SEM characterization.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

The authors declare the following financial interests/personal relationships which may be considered as potential competing interests.

Acknowledgements

This work was supported as part of the Joint Center for Energy Storage Research, an Energy Innovation Hub funded by the U.S. Department of Energy, Office of Science, Basic Energy Sciences.This work was supported by the 2021 Research Fund (1.210041.01) of UNIST(Ulsan National Institute of Science & Technology)

References (62)

  • H. Perthuis et al.

    Well densified NASICON type ceramics, elaborated using sol-gel process and sintering at low temperatures

    Mat. Res. Bull.

    (1984)
  • R.S. Gordon et al.

    Fabrication and characterization of NASICON electrolytes

    Solid State Ionics

    (1981)
  • P. Colomban

    Orientational disorder, glass/crystal transition, and superionic conductivity in NASICON

    Solid State Ionics

    (1986)
  • S. Naqash et al.

    Na2Zr2(SiO4)2(PO4) prepared by a solution-assisted solid state reaction

    Solid State Ionics

    (2017)
  • R.O. Fuentes et al.

    Influence of microstructure on the electrical properties of NASICON materials

    Solid State Ionics

    (2001)
  • S. Yde-Andersen et al.

    Properties of Nasicon electrolytes prepared from alkoxide derived gels: ionic conductivity, durability in molten sodium and strength test data

    Solid State Ionics

    (1984)
  • J.F. Nonemacher et al.

    Micromechanical assessment of Al/Y-substituted NASICON solid electrolytes

    Ceram. Int.

    (2019)
  • J. Kim et al.

    Low temperature performance of sodium-nickel chloride batteries with NaSICON solid electrolyte

    J. Electroanal. Chem.

    (2015)
  • J. Wolfenstine et al.

    Electrical and mechanical properties of hot-pressed versus sintered LiTi2(PO4)3

    Solid State Ionics

    (2009)
  • S. Smith et al.

    Electrical, mechanical and chemical behavior of Li1.2Zr1.9Sr0.1(PO4)3

    Solid State Ionics

    (2017)
  • W.H. Baur et al.

    Neutron powder diffraction study and ionic conductivity of Na2Zr2SiP2O12 and Na3Zr2Si2PO12

    Solid State Ionics

    (1986)
  • S. Naqash et al.

    Impact of sodium excess on electrical conductivity of Na3Zr2Si2PO12 + x Na2O ceramics

    Solid State Ionics

    (2019)
  • G. Han et al.

    Correlating the effect of dopant type (Al, Ga, Ta) on the mechanical and electrical properties of hot-pressed Li-garnet electrolyte

    J. Am. Euro. Ceram. Soc.

    (2020)
  • J.-M. Tarascon et al.

    Issues and challenges facing rechargeable Li batteries

    Nature

    (2001)
  • A. Manthiram et al.

    Li battery chemistries enabled by solid-state electrolytes

    Nat. Rev. Mater.

    (2017)
  • P. Albertus et al.

    Status and challenges in enabling the Li metal electrode for high-energy and low-cost rechargeable batteries

    Nat. Energy

    (2018)
  • H. Romero et al.

    Mining development and environmental injustice in the Atacama Desert of Northern Chile

    Environ. Just.

    (2012)
  • D.B. Agusdinata et al.

    Socio-environmental impacts of Li mineral extraction: towards a research agenda

    Environ. Res. Lett.

    (2018)
  • C. Zhao et al.

    Solid-state sodium batteries

    Adv. Energy Mater.

    (2018)
  • C.D. Fincher et al.

    Elastic and plastic characteristics of sodium metal

    ACS Appl. Energy Mater.

    (2020)
  • M. Bay et al.

    Sodium plating from Na-β″-alumina ceramics at room temperature, paving the way for fast-charging all-solid state batteries

    Adv. Energy Mater.

    (2020)
  • Cited by (16)

    View all citing articles on Scopus
    View full text