Recent Advances in NASICON Type Oxide Electrolytes For Solid State Sodium Ion Rechargeable Batteries

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 31

Ionics

https://doi.org/10.1007/s11581-022-04765-3

REVIEW

Recent advances in NASICON‑type oxide electrolytes for solid‑state


sodium‑ion rechargeable batteries
Kushal Singh1 · Anjan Chakraborty1 · Raghunayakula Thirupathi1 · Shobit Omar1,2

Received: 12 June 2022 / Revised: 25 August 2022 / Accepted: 18 September 2022


© The Author(s), under exclusive licence to Springer-Verlag GmbH Germany, part of Springer Nature 2022

Abstract
Solid-state batteries have shown the potential to resolve the safety and durability issues associated with traditional liquid
electrolyte-based batteries. This article reviews the current developments of NASICON-type solid electrolytes for Na-ion
solid-state batteries. These ceramic-based oxides possess a 3D open-framework structure allowing for the fast diffusion of
large sodium ions. To date, the conductivity value as high as ~ 5 mS·cm−1 at 25 °C is reported for these materials, which needs
to be further improved. The requirement of high-temperature sintering (> 1200 °C), anisotropic thermal expansion, impurity
phase formation, and large interfacial impedance are other challenges of NASICON-type electrolytes. This article summa-
rizes various fundamental aspects governing the sodium-ion conduction in these oxides. Particular emphasis is given to the
strategies employed in recent investigations to improve the properties and alleviate the associated issues in designing stable
solid-state sodium-ion rechargeable batteries. This will also establish the groundwork for future research in these materials.

Keywords  Sodium ions · Inorganic solid electrolytes · Ionic conductivity · All-solid-state batteries · Interfaces

Introduction first commercialization by SONY Energytec in 1991 for the


handheld video recorder. They possess high specific energy
The ever-increasing global energy demand has resulted in (~ 250 Wh/Kg), superior operating voltage (∼3.6 V), longer
not only a rapid depletion of already scarce fossil resources, cycle life (400–1200 cycles or 3–4 years), and a faster charg-
but also a severe environmental crisis [1, 2]. In recent years, ing rate [10–17]. Although they are attractive, the short-
tremendous efforts have been made to evolve technologi- age of high-quality global lithium mineral reserves is a key
cal solutions to effectively utilize renewable energy sources impediment to their broad usage in utility-scale storage
such as solar, wind, and hydro. Since energy supply from systems [18]. Besides inadequate availability, high Li cost
these sources can be intermittent, storage of excess energy has driven the research to explore into other technologies.
is critical for realizing a clean and sustainable future [3, 4]. Na-ion batteries (NIBs) could become a low-cost storage
Alternative energy storage solutions are also vital because of alternative for future electric grids as intercalation chemistry
the recent surge in our demand for electric vehicles, which in both sodium and lithium is similar.
can mitigate our reliance on fossil fuels in automobile sec- Even though research publications related to Li-ion bat-
tor [5, 6]. Even though there exist multiple ways through teries are much more than Na-ion batteries (Fig. 1a), NIBs
which energy can be stored, battery technology has shown have gained considerable scientific interest in recent years.
immense promise to play a massive role in the energy stor- An exponential growth has been recorded in the research
age sector over the next decade [1, 7–9]. Li-ion batteries associated with NIBs (Fig. 1b). Although larger and heavier
(LIBs) have dominated consumer electronics since their sodium exhibits slower diffusion, its widespread availability
and low cost make NIBs suitable for large-scale applications
* Shobit Omar [18–24].
somar@iitk.ac.in One of the primary concerns for the practical applica-
tions of NIBs using non-aqueous liquid electrolytes arises
1
Materials Science and Engineering, Indian Institute from battery safety regardless of the competitive ionic con-
of Technology, Kanpur, Uttar Pradesh 208016, India
ductivity (~ 10 mS·cm−1) [25]. The organic solvents utilised
2
Sustainable Energy Engineering, Indian Institute in these electrolytes are flammable and thermally unstable,
of Technology, Kanpur, Uttar Pradesh, 208016, India

13
Vol.:(0123456789)
Ionics

Fig. 1  (a) Number of research publications related to Na- and Li-ion batteries in recent years. (b) Percentage increase in research publications in
the area of Na- and Li-batteries since 2011. (Source: Scopus, Google scholar)

posing major safety concerns, especially when the batter- electrolytes, i.e., NASICON, garnet, perovskite, LISICON,
ies are overcharged, internally shorted, or subjected to high LiPON, ­Li3N, sulfides, argyrodite, anti-perovskite, etc.,
temperatures [26]. Several approaches have been studied to are used for making solid-state LIBs [41]. However, for
modify the electrolyte and improve the safety and stabil- Na solid-state systems, only NASICON, β//-alumina, and
ity of cells. Solid-state electrolytes (SEs) appear to be the sulfides are used as ceramic solid electrolytes [25].
most attractive option because of their excellent thermal Na-ion conducting SEs can be broadly classified into
and chemical stabilities [27–30]. Furthermore, SEs have a three categories, viz. (1) polymer-based electrolytes, (2)
broader electrochemical potential window (up to 8.8 V vs sulfide-based electrolytes, and (3) oxide-based electro-
Na/Na+ till now) than the liquid organic electrolytes (< 5 V lytes. It is noteworthy that the material to be employed
vs Na/Na+), allowing them to combine with high-voltage as an electrolyte should have a conductivity of at least
cathodes and render high specific energy to the batteries 1 mS·cm −1 for any realistic design of solid-state NIBs.
[30]. SEs also simplify the battery design by eliminating Each type of the above-mentioned electrolytes has its
the need for a separator to prevent physical contact between own sets of advantages and challenges [42–44]. Although
the cathode and anode. Nonetheless, the solid-state NIBs polymer-based electrolytes are more flexible and have
are still in the early stages of development, largely due to lower contact resistance between the electrode and elec-
the unavailability of solid electrolytes with high sodium-ion trolyte, their room-temperature (RT) ionic conductivity
conductivity and poor interfacial compatibility between the (~ ­10 −6–0.1 mS·cm −1) is inadequate [25, 45]. Sulfide-
electrodes and SE [27, 28, 31–37]. based electrolytes managed to achieve moderate to high
The cell chemistry of the solid-state NIBs differs from ionic conductivity (~ 0.1–1 mS·cm−1) and form a favour-
that of the LIBs. Anode materials mainly comprise metallic able solid–solid interface with electrodes. However, their
lithium for Li solid-state system, whereas metallic sodium poor stability under ambient air poses a major fabrica-
is used for Na solid-state system. Materials for the positive tion challenge [25, 46]. Oxide-based sodium β//-alumina
electrode are primarily compounds containing the 3d tran- ­( Na 2O-5Al 2O 3) was the first commercially used supe-
sition metal cations, namely Co, Mn, Fe, or Ni, as redox- rionic conductor employed in Na-S and Zebra batter-
active elements. For Li solid-state system, layered oxides ies operating at ~ 250–300 °C [25, 30, 32, 46, 47]. The
such as ­LiCoO2, ­Li[Ni1-x-yMnxCoy]O2, olivine-structured requirement of high sintering temperature (> 1500 °C)
­LiFePO4, spinel ­LiMn2O4, and NASICON-type ­Li3V2(PO4)3 and the two-dimensional ionic conduction make this
with capacity values mostly in the range of 140–280 mAh/g ceramic unsuitable for large-scale storage applications
are used as cathode material [16, 38]. The direct analogues [25, 45, 46]. The oxides possessing the skeleton struc-
of these materials, i.e., ­NaCoO2, ­NaFePO4, ­NaMn2O4, and ture of NASICON (Na Super-Ionic Conductor) have
­Na3V2(PO4)3, are the preferred choice for Na solid-state bat- been demonstrated to achieve > 1 mS·cm −1 at RT [33].
teries with capacity values lying in the range of 100–250 The high conduction is attributed to the three-dimension-
mAh/g [21, 39, 40]. Several types of solid inorganic ally connected interstitial sites, which can accommodate

13
Ionics

large-sized sodium ions and provide paths for fast 3D NASICON overview
ionic diffusion inside the structure. Furthermore, the
strong covalent framework imparts good thermal and Chemistry and crystal structure
chemical stabilities to the structure [48–54]. Recognizing
the potential of NASICON-structured oxides, research- In general, the chemical composition possessing the NASI-
ers all over the world have shown keen interest in devel- CON structure can be represented as ­Na1+xZr2SixP3-xO12
oping them as SE for the solid-state NIBs operating at (0 ≤ x ≤ 3) [60]. It consists of a rigid three-dimensional
RT. Nevertheless, the high ionic impedance offered by network where each tetrahedron of either S ­ iO4 or P ­ O4 is
the NASICON electrolyte and the interfacial reaction connected to four ­ZrO6 octahedra by corner shared oxy-
between the electrode and electrolyte, which blocks the gen ions [61]. The pathways for the migration of Na ions
large-scale transfer of ions across the interface and sig- are formed by interconnecting channels that connect one
nificantly affect the cycling performance of solid-state interstitial site to another through a structural bottleneck.
batteries (SSBs), remain the primary challenges. NASICON term can also be used to describe a fam-
The solid-state NIBs employing NASICON ceramic ily of compounds with the general chemical formula of
electrolyte differ from Li-based SSBs. The melting point ­AMP3O12, where A denotes the sites occupied by cations
of the metal anode plays a crucial role in cell reaction which can be monovalent (such as ­K+, ­Na+, ­Li+), divalent
kinetics. Since the Na metal possesses a lower melting (such as C ­ o2+, ­Ca2+, ­Mg2+, ­Cu2+, ­Ba2+), trivalent (such
point ( TNa
m
 = 98 °C) than the Li metal ( TLi
m
 = 181 °C), it can as ­Y , ­A l 3+) or tetravalent (such as Z
3+
­ r 4+, ­G e 4+; ­H f 4+)
be converted into a molten state at a lower temperature, [25, 51]. On the other hand, M corresponds to the sites
resulting in enhanced kinetics [55]. Furthermore, molten occupied by cations that are divalent (such as M ­ n 2+,
metal can spread homogeneously on the ceramic NASI- ­Cd , ­Ni , ­Co , ­Zn ), tetravalent (such as ­Zr , ­Si4+,
2+ 2+ 2+ 2+ 4+

CON electrolyte, suppressing the dendrite formation along ­Ti4+), or pentavalent (such as S ­ b5+, ­Nb5+, ­V5+) [51, 60].
the grain boundaries [56]. Besides, sodium is relatively Due to this broad compositional space, NASICON-type
softer than lithium, which makes handling and process- oxides can be used as a solid electrolyte in both Li and Na
ing less challenging. The solid electrolyte is also more SSBs. However, the Li-ion conductivity in NASICON-
thermodynamically stable in direct contact with sodium type ­Li1.3Al0.3Ti1.7(PO4)3 solid electrolyte is much infe-
as Na is less reducing than Li. However, a large volume rior to that of N ­ a+ ionic conductivity in NASICON-type
change occurs in the electrode during charge–discharge ­Na3Zr2Si2PO12 [62, 63]. This is attributed to the higher
cycling of NIBs because of the bigger size of the sodium mobility of N ­ a+ compared to L ­ i + inside the NASICON
ion, which may lead to poor cyclability. Thus, retaining a structure, as both the ions possess similar conduction
stable electrode–electrolyte interface in Na-SSBs is more pathways. The lower Lewis acidity of N ­ a+ compared to
challenging than in Li-SSBs. +
­Li causes a weaker electrostatic attraction force between
A few review articles on NASICON-type materials ­Na + and skeletal ­O 2− ions, enabling a facile transfer of
for SIBs have been previously published. Rajagopalan sodium ions [64]. On the other hand, the smaller ­Li+ ions
et al. [57] reviewed the NASICON-structured electrode form a stronger bond with ­O2− resulting in lower Li-ion
and electrolyte materials, focusing mainly on the mecha- conductivity.
nisms for ionic conduction inside the crystal lattice. Rao When NASICON composition is represented as
et al. [58] looked into the synthesis and doping strategies ­NaA /MP 3O 12, sodium is either completely or partially
employed to enhance the Na-ion conductivity of NASI- filled in ­A / sites to preserve the bulk charge neutrality.
CON-based solid electrolytes. Similarly, Yang et al. [59] The occupancy of ­A/-sites governs the concentration of
presented the basic characteristics of NZSP and summa- the mobile Na-ions and their interactions with the immo-
rized the approaches used to improve the grain and grain bile skeleton atoms. These interactions or crystalline fields
boundary conductivities. However, in these studies, the can preferentially stabilize the sodium-ion at specific sites
microstructure, thermal and mechanical properties crucial resulting in a high activation energy barrier for migra-
for fabricating solid-state batteries were not discussed. The tion within the crystal lattice. Although a wide range of
current review article presents an overview of NASICON- chemical compositions exists, not all of them crystallize
type oxides encompassing their chemistry, phase forma- into the NASICON phase when synthesized. In a recent
tion, microstructure, conduction mechanism, etc. It also work, Ouyang et al. [65] developed the phase stability
highlights the strategies utilized to improve sodium-ion rules of NASICON-type materials with a broad compo-
conductivity, the challenges associated with fabrication, sitional space. A high throughput DFT calculations were
and the approaches used to overcome them. Other issues performed on 3881 possible NASICON compounds with
are also discussed, which will aid in directing future the chemical formula of ­Na x M y M′ 2−y (AO 4 ) z (BO 4 ) 3−z .
research on these materials for solid-state NIBs.

13
Ionics

While the Na content in the tested compositions ranged sites are six-fold coordinated with the neighboring oxygens
from 0.0 to 4.0 (in steps of 0.5), the average M/M’ charge of ­SiO4/PO4 tetrahedra. There exists one Na(1) and three
state was increased from + 2 to + 5 (in steps of 0.5). A two- Na(2) sites per formula unit. The low-symmetry monoclinic
dimensional descriptor was formulated, which can predict phase possesses another different Na site, i.e., Na(3) along
NASICON compositions that are likely to be synthesized. with Na(1) and Na(2) sites (shown in Fig. 2b) [33, 68]. This
It was reported that ~ 83.4% of the total compositions eval- additional Na(3) sites are three-fold coordinated with the
uated in this study are unlikely to be synthesized with the adjacent ­ZrO6 octahedra [33, 69]. Thus, the monoclinic
NASICON structure. phase consists of one Na(1), one Na(2) and two Na(3) sites
At room temperature, the crystal structure of per formula unit.
­Na1+xZr2P3-xSixO12 (0 ≤ x ≤ 3) remained monoclinic (C2/c) Figure 3a shows the X-ray diffraction (XRD) profiles of
in the range of 1.8 ≤ x ≤ 2.2, whereas the rhombohedral ­Na3Zr2Si2PO12 (NZSP) collected at various temperatures
phase ( R3c ) exists for all other values of x [51]. Several [70]. The monoclinic phase existing at RT in NZSP gradu-
monoclinic-rhombohedral phase transition studies demon- ally transforms to the high-symmetry rhombohedral phase
strate that in the compositions with the above given x range, on raising the temperature. It is worth noting that most of
the high-temperature stable rhombohedral phase loses sym- the XRD peaks in both phases lie very close to each other
metry upon cooling below 160 °C, resulting in the mono- and therefore overlap. However, the distinction between the
clinic distortion [66, 67]. Deng et al. [67] studied the phase two phases can be visualized from the small regions of the
behavior of rhombohedral N ­ a1+xZr2SixP3-xO12 electrolytes by XRD pattern, as shown in the enlarged view in Fig. 3b. With
constructing a phase diagram employing DFT calculations, increasing temperature, the diffraction peaks at 19.3° and
cluster expansion formalism and large-size Monte Carlo 30.5°, which correspond to the monoclinic phase, gradually
simulations. It was revealed for the first time that the two fade away, indicating the transformation to the rhombohedral
biphasic regions exist in the Na compositions of 0 < x < 3 phase.
and at temperatures below 650 K. This is in contrast to pre-
viously reported experimental results, which indicate that Conduction pathways
a single phase exists in the whole compositional range and
temperature greater than 450 K. Moreover, four monophasic Understanding the ionic transport mechanism is essential
regions are also found in the phase diagram [67]. However, for the rational design of NASICON-type ­NaA/MP3O12.
the phase separation in NZSP has experimentally not been As stated earlier, ­Na+ ions occupy two crystallographically
detected so far. The disagreement in the results is attributed distinct interstitial sites inside the rhombohedral structure.
to the poor kinetic, which prevents the redistribution of Si In contrast, the monoclinic phase, which is formed by the
and P and inhibits the phase separation. rotational distortion of the rhombohedral structure, has
Figure 2a illustrates that the ­Na+ ions occupy two crystal- three distinct locations for Na-ion migration inside the
lographically inequivalent interstitial sites in the rhombohe- stationary skeleton [33, 68]. During the rhombohedral to
dral phase, namely Na(1) and Na(2), present at the Wycoff monoclinic phase transition, the ­ZrO6 and (Si/P)O4 frame-
positions of 6b and 18e, respectively. The Na(1) and Na(2) work remain unaffected and only the Na(2) sites of the

Fig. 2  Schematic of NASICON-
type (a) rhombohedral ( R3c )
and (b) monoclinic (C2/c)
crystal structures

13
Ionics

Fig. 3  (a) XRD patterns of NZSP taken at various temperatures from RT to 200 °C. The enlarged view of the two regions in the XRD patterns is
also shown (b). (Reproduced with permission from ref. no. [70])

rhombohedral phase split into two positions, i.e., Na(2) and In the rhombohedral structure (shown in Fig. 4a), each
Na(3) in the monoclinic phase [71]. It is noteworthy that Na(2) site is locally surrounded by two Na(1) sites, whereas
the Na-ion diffusion in both the structures occurs by pass- each Na(1) site is surrounded by six Na(2) sites [74]. Increas-
ing through a bottleneck triangular area formed by the oxy- ing the distance between Na(1) and Na(2) increases the acti-
gens of the neighbouring tetrahedral and octahedral units vation energy required for ionic conduction. However, if the
(shown in Fig. 4). In rhombohedral structure, Na(1)-Na(2)- distance is considerably low, the activation energy rises as
Na(1) is the energetically preferred pathway for Na-ion dif- well. As a result, the Na(1)-Na(2) distance should be optimal
fusion (Fig. 4a). Nonetheless, both the Na(1)-Na(2)-Na(1) for fast ion hopping, which can be attained by tuning the lattice
and Na(1)-Na(3)-Na(1) pathways are involved in ionic con- size [33]. Na ions avoid hopping between two Na(2) sites in the
duction inside the monoclinic phase (Fig. 4b). Thus, ­Na+ rhombohedral symmetry as one oxygen ion resides between
ion jumps from Na(1) to Na(2) or Na(3) (approximate gap the two Na(2) sites. Furthermore, the Na(1)-Na(1) and Na(2)-
of 3 Å between sites depending on material composition), Na(2) distances, which are reported to be 6.48 Å and 4.98 Å,
resulting in 3D tunnelling of Na-ions through intercon- respectively, in ­Sc3+ substituted NZSP, are significantly greater
nected interstitial sites [25, 33, 68, 73]. than the distance between Na(1) and Na(2) sites (3.46 Å) [74].

Fig. 4  Diffusion pathway in
NASICON-type (a) rhombohe-
dral and (b) monoclinic crystal
structures

13
Ionics

these elementary diffusion steps, three preferable conduc-


tion pathways in the x, y and z directions are examined. It is
found that in all the pathways, sodium ions are required to
pass through the puckered hexagonal ring with sides hav-
ing alternate tetrahedral and octahedral edges. Furthermore,
the activation energy involved for sodium-ion diffusion in
x, y and z directions are 285 meV, 315 meV, and 260 meV,
respectively, indicating a high three-dimensional conductiv-
ity inside the NZSP lattice.
In a recent work, Zhang et al. [77] used ab initio molecu-
lar dynamics simulations and climbing image nudged elastic
model calculations and reported a much lower diffusion bar-
rier for the correlated migration compared to the single-ion
direct hopping mechanism in the monoclinic NZSP. The
obtained results suggest that instead of a direct hopping
mechanism, correlated diffusion is the preferred conduction
mechanism in NZSP. Furthermore, it is predicted that this
Fig. 5  Arrhenius plot for the ionic conductivity of rhombohedral and
monoclinic phase of pure N ­ a3Zr2Si2PO12 obtained through molecular
type of correlated conduction can be enhanced by increasing
dynamics simulations. (Reproduced with permission from ref. no. the ­Na+ ion concentration.
[72])
Factors influencing the sodium‑ion conductivity

The sodium ions hop along the Na(1) → Na(2) → Na(1) chan- The magnitude of ionic conductivity in NASICON mate-
nels and pass through the two bottleneck triangles, ­T1 and ­T2 rial is primarily determined by the concentration of mobile
in the rhombohedral structure, depicted in Fig. 4a. ions present in the local symmetrical structure, the number
The monoclinic unit-cell of NASICON contains four of available hopping sites, and the energy barrier for diffu-
Na(1) sites, four Na(2) sites, and eight Na(3) sites [60]. The sion [45, 78]. Besides, the % bulk density of the specimen,
Na(2) sites are fully occupied, while Na(3) sites are nearly phase stability, and the vacancy distribution throughout the
half empty. Each Na(1) site in the local environment is sur- structure are all important controlling parameters to improve
rounded by two Na(2) sites and four Na(3) sites. However, conductivity. Increasing the concentration of charge carriers
partially occupied Na(3) sites are strongly displaced towards increases conductivity but only to a certain extent. At very
Na(1) sites [71]. A complete occupation of Na(2) sites tends high Na content, the conductivity declines mainly because
to resist the ionic movement. Bond valence analysis has been there are insufficient vacant sites for allowing the migration
performed in order to comprehend and predict sodium con- of sodium ions. Thus, the concentration of available vacan-
duction through different possible conduction path inside cies and occupied Na-sites should be optimal to accomplish
the monoclinic structure. By estimating the valence value fast sodium-ion conduction within the lattice. Based on the
maxima (VUmax) of various possible pathways [75], it was conductivity survey performed on a wide range of NASI-
concluded that there exist two non-equivalent Na(1)-Na(3) CON compounds, Guin et al. [69] suggested that the maxi-
migration paths. Both of them exhibit much lower valence mum conductivity is attained in the compositions that have
values (VUmax is 0.984) compared to Na(1)-Na(2) pathway Na content in the range of 3.2–3.5 per formula unit. The
(VUmax is 1.28). As a result, Na-ions diffuse easily along maximum Na ions present in a single unit cell is 4. Similarly,
with both the Na(1)-Na(3) paths without reaching a valence Zhang et al. [77] reported the highest conductivity in NZSP
sum higher than or equal to 1 at any stage. The valence bond at an optimum Na content of 3.3–3.55 per formula unit.
analysis also explains the exceptionally high ionic conductiv- The sodium-ion conductivity is also dependent on the
ity observed in the monoclinic NASICON structure [75, 76]. crystalline phase present in the NASICON compounds at
Bui et al. [61] studied the conduction mechanisms in RT. Since the rhombohedral phase is more symmetric than
NZSP exhibiting the monoclinic structure using the density the distorted monoclinic phase, it exhibits lower activa-
functional theoretical calculations. Four elementary diffu- tion energy for sodium-ion diffusion. Ran et al. [72] per-
sion steps are identified, of which three are inner-chain pro- formed molecular dynamics simulations to study the ionic
cesses, while the remaining one is an inter-chain process. conductivity in NZSP possessing pure rhombohedral and
In the inner-chain diffusion, the sodium ions migrate inside monoclinic phases. The ionic conductivity of the rhombo-
the chain. On the other hand, in the inter-chain diffusion, hedral phase is predicted to be one order of magnitude supe-
the sodium ions also jump across the chains. Combining rior to that of the monoclinic phase (as shown in Fig. 5).

13
Ionics

Furthermore, the activation energy for the rhombohedral such as tetragonal ­Na3PO4 and ­Na2ZrSi2O7 are also com-
phase (0.23 eV) is found to be much lower compared to the monly detected at grain boundaries.
monoclinic phase (0.33 eV). The magnitude of ionic conductivity inside the grains and
One of the strategies to improve the ionic conductivity is across the grain boundaries differs within the polycrystal-
to enlarge the area of the smallest oxygen triangle, T ­ 1, which line ceramic electrolyte. Several conductivity studies have
serves as a bottleneck for ionic conduction in NASICON revealed that the grain boundaries offer a significant por-
electrolytes [79]. The distance between the Na-ion and the tion of the total resistivity in NASICON-type oxides at RT
nearest oxygen ions is also critical [33]. In addition, the host [87]. The blocking nature of grain boundaries is attributed
(M) and dopant ­(M/) cations in ­AMM/P3O12 composition to the formation of insulating secondary phases, generation
should have the ideal size to achieve high ionic conductiv- of micro-cracks, and poor grain to grain contact because of
ity [69]. The partial substitution with acceptor cations in the cuboidal shaped grains in the microstructure.
NASICON structure plays an important role as doping not
only modifies the bottleneck area (depending on the bond Powder synthesis and bulk sample fabrication
length and ionic radius) but also boosts the concentration of
mobile Na-ion in the NASICON structure because of charge Powders of NASICON oxides are extensively synthesized
imbalance [80–82]. using two methods, i.e., sol–gel and solid-state reaction pro-
The arrangement of S ­ i4+ and P­ 5+ in the NZSP composi- cesses [90, 95, 98–101]. However, in a few studies, other
tions also impacts the sodium-ion transportation inside the processes such as hydrothermal [102, 103], co-precipitation
NASICON crystal structure. Roy et al. [68, 83] illustrate this [104], mechanochemical synthesis [105], and combustion
effect by performing molecular dynamics simulations on the synthesis [106] are also employed for the powder synthesis.
pure NZSP compound. Three framework structures of NZSP In the solid-state reaction method, the stoichiometric pro-
are considered with different Si/P ordering. The structure in portions of required powder reactants are thoroughly mixed
which ­Si4+ ions are uniformly distributed, well-constructed and ground into fine particles using wet ball milling. The
percolation channels of sodium ions are seen. On the other powder mixtures are then calcined to attain the NASICON
hand, islands of ­Na+-ions exist in the framework structure phase. The as-calcined powder is pressed into bulk green
with ­Si4+ rich planes. The low ionic conductivity observed ceramics which are subjected to high-temperature sintering
­ i4+ rich planes is ascribed to the modu-
in the structure with S (> 1200 °C) to achieve high densification. Even though the
lation of the electrostatic potential energy landscape, which solid-state reaction is a simple and low-cost method, the
originates because of the higher electrostatic repulsion of high-temperature treatment is a primary concern as it may
­Na+ with P ­ 5+ compared to S ­ i4+. Thus, the distribution of cause the volatilization of high vapour pressure elements,
Si/P in NZSP could be one of the reasons for observing such as Na and P and formation of undesirable secondary
conductivity variation in NZSP specimens when prepared phases [89, 100, 107–111].
under different processing conditions. On the other hand, the sol–gel synthesis method offers the
The sintering temperature and the holding time are two possibility of low-temperature processing, better morphologi-
critical parameters that determine the % bulk density and cal control, and superior chemical homogeneity within the
microstructure of any sintered specimen. In NASICON-type specimen [84, 105, 108]. However, the high cost incurred in
oxides, sintering at elevated temperatures (> 1200 °C) results this method and the processing complexity are the main chal-
in the formation of an electrically insulating secondary phase lenges, especially for large-scale production. In this process-
of monoclinic-ZrO2 at the grain boundaries [84–87]. The ing route, precursors of Zr and Si are initially hydrolyzed,
formation of zirconia is associated with the volatilization of followed by the addition of P and Na salts. The colloidal
the Na and P elements from the bulk samples on subjecting solution is then slowly heated to form a gel network which
them to high sintering temperature. In several investigations, is subsequently subjected to temperatures in the range of 700
excess amounts of P and Na are added during the powder to 1000 °C for the crystallization of a desirable NASICON
synthesis process to compensate for the cation volatilization phase [112, 113]. Bulk samples are pressed using the nano-
and consequently suppress the ­ZrO2 formation [81, 88, 89]. sized powder and are sintered at relatively low temperatures
In addition, multiple materials processing approaches, such between 1000 to 1250 °C [113, 114]. The processing param-
as sol–gel powder synthesis [34, 90], the addition of sinter- eters such as calcination temperature, sintering temperature,
ing aids ­(Na3BO3 [91–93], ­Bi2O3 [94], ­Na2SiO3 [70], ­Na2O, and pH of the sol are vital in stabilizing various phases in the
etc.), and various additives (Antimony tin oxide [95], Indium bulk specimen. The sintering temperature cannot be too high
tin oxide [96], etc.) to the starting mixture, and the use of since this could cause impurities such as Z ­ rO2 and N
­ a3PO4 to
advanced sintering techniques [50, 88, 97], are adopted to develop in the microstructure. In most cases, excess P and Na
attain high densification without subjecting the specimens to are added to compensate for the evaporation of these elements
the elevated sintering temperatures. Other secondary phases during the synthesis process [81, 89].

13
Ionics

Even though wet-chemistry approaches for synthesizing furnace at 800–1000 °C to produce the NASICON phase.
nano-sized powders are advantageous for producing dense The SA-SSR yields powder with good chemical homoge-
specimens at low temperatures, these processes require neity, low Z­ rO2 secondary phase and ultrafine particles
expensive precursors and challenging to scale up. As a of 0.1–0.2 μm.
result, alternative materials processing techniques are desir- The strategy of adding sintering aid in fabricating
able for lowering the sintering temperature while avoiding dense bulk NASICON ceramics has also proven to be
the formation of impurity phases, which have an adverse effective. It is reported that several sintering aids such as
effect on sodium-ion conduction. These solutions should ­Na3BO3 [91, 93], ­B i 2O3 [94], ­6 0Na 2O-10Nb2O5-30P2O5
also be cost-effective and adaptable to large-scale material [116] not only lower the sintering temperature but also
production. considerably improve the ionic conductivity. For exam-
In recent work by Naqash et al. [115], a novel method ple, NZSP ceramic sample with 9.1 wt.% ­Na 3BO 3 sin-
based on solution-assisted solid-state reaction (SA-SSR) tered at 700 °C exhibits three orders of magnitude supe-
is developed for the powder synthesis of NASICON rior conductivity (0.1 mS·cm−1) at RT than the specimen
compounds. This process is gaining popularity as it is without ­Na 3BO 3 [93]. Moreover, several advanced sin-
convenient, utilizes low-cost raw materials (water or acid tering routes, such as microwave sintering [97], spark
soluble salts), and is simple to scale up for large-scale plasma sintering [88], and cold sintering processes [50,
production. Figure 6 shows the schematic depicting the 94, 117, 118] have been explored to achieve the required
steps involved in the powder synthesis of NZSP. The stoi- densification at a relatively low sintering temperature
chiometric amounts of nitrates of Na and Zr are initially [50, 119]. Among these, the cold sintering process has
dissolved in water. Si(OC2H5)4 (TEOS), which serves as become increasingly attractive for the sintering of NASI-
a Si source, is then added to the solution, followed by a CON electrolytes as it can achieve high densification
little amount of nitric acid for the hydrolysis of TEOS. at a very low temperature (as low as 140 °C) [50]. This
Subsequently, a stoichiometric amount of ­NH4H2PO4 is recently developed technique utilizes various sintering
added, and stirred for 30 min to generate a homogenous aids (also called transient liquid) such as water, ace-
precipitate. The precipitate-containing solution is then tic acid, ­H NO 3, NaOH, KOH, etc. [119] Nevertheless,
slowly heated to 100 °C for the gentle removal of water. this method requires further investigations to gain deep
The resulting solid product is subsequently calcined in a insight into the actual sintering mechanism.

Fig. 6  Schematic depicting the steps involved in the solution-assisted solid-state reaction method for the powder synthesis of NASICON-type
oxides. The process is described in ref. no. [115]

13
Ionics

Microstructure of sintered samples the particles, and uniform cuboidal-shaped grains [128].
Further, this approach can reduce the sintering temperature
The microstructure has a substantial influence on the final and holding time and enhance the carrier concentration
properties of NASICON electrolytes. Good sintering, neg- [129]. Figure 7 shows the cuboidal-shaped grains typi-
ligible porosity, and the absence of micro-cracks favor high cally observed in the microstructure of these NASICON
ionic conductivity [99, 120]. The microstructure of the compositions.
NZSP specimen sintered at 1200 °C reveals the presence An optimum amount of various additives (such as NaF,
of several macropores and voids, which can be attributed Antimony tin oxide, Indium tin oxide, etc.) are added
to the partial melting of the NASICON phase [121]. The in NZSP to achieve dense microstructure, better neck-
liquid phase flows between the solid particles leaving behind ing between the grains, and uniform grain size [95, 96,
macropores [106]. Increasing the sintering temperature and 130]. Suzuki et al. [93] found that adding N ­ a 3BO 3 into
holding time favour the grain growth and densification of the NZSP improves the microstructure by promoting neck
the sample, which improves the conductivity up to a certain formation and grain growth, leading to high densifica-
limit [88, 99, 122]. The excessive sintering temperature or tion. The fractured surface of the ceramic without NBO
prolonged sintering duration results in Na/P deficient NZSP is shown in a secondary electron (SE) image and a back-
and promotes the liquid phase formation and the unwanted scattered electron (BSE) image in Fig. 8a and b, respec-
secondary phases ­(ZrO2, ­Na3PO4, etc.) near the grain bound- tively. The porous microstructure with nano-sized grains
aries [84, 122, 123]. As a result, there has always been a (100–300 nm) suggests little sintering occurring at 700 °C.
trade-off between the dense microstructure and the presence The BSE image revealed a sparse ceramic microstructure,
of impurity phases. indicating that necks developed to some extent between
Secondary phases such as Z ­ rO2, ­ZrSiO4, and S
­ iO2 can NASICON grains. If the presence of the Z ­ rO2 impurity
also be present in polycrystalline NASICON electrolytes, phase is ignored, the NBO-free ceramic body exhibits a
which may be detrimental to the mechanical integrity of relative theoretical density of 61% (see Fig. 8b). On the
SSBs [124]. Since ­ZrO2 is more thermodynamically stable, other hand, the ceramic body sintered with 9.1 wt% NBO
­ZrO2 crystallization occurs near the grain boundary region possesses a dense microstructure, as shown in Fig. 8c, d).
before the NASICON phase [125]. According to an effec- Despite the presence of tiny pores, the neck development
tive percolation theory, a small amount of the ­ZrO2 impu- and the grain growth progressed significantly compared to
rity phase in the grain boundary has a minimal influence on the NBO free sample. The distribution of ­ZrO2 and NBO
ionic conductivity since it acts as porosity. However, a large phases is shown in the BSE micrograph (Fig. 8d).
amount of Z ­ rO2 formation could lead to a different stoichi- Several sophisticated sintering processes are employed
ometry caused by the Zr-deficiency in the NASICON phase to fabricate the bulk samples of NZSP electrolyte with
[115, 126]. As a result, an inhomogeneous distribution of dense and fine-grained microstructure. Leng et al. [50]
Na ions exists within the grains and grain boundaries, which used a cold sintering technique and optimized the process-
has a detrimental effect on the intergranular migration of Na ing parameters to fabricate bulk specimens of Mg-doped
ions [81, 119, 127]. NASICON electrolyte (­ Na3.256Mg0.128Zr1.872Si2PO12). Fig-
The grain boundaries in the NZSP specimen can be ure 9 compares the microstructure of a specimen cold-
altered by using suitable dopants, varying the dopant con- sintered at 140 °C for 1 h under 780 MPa with a speci-
centration, and controlling the reactivity with the host ele- men dry-pressed at 780 MPa for 1 h. The cold sintered
ment [58]. Partial substitution with rare-earth elements specimen exhibits higher densification, better sintering
(such as L­ a3+, ­Y3+, ­Ce4+, ­Gd3+, etc.) in NZSP could result morphologies such as particle packing and neck forma-
in smaller size grains (~ 2–4 μm), greater necking between tion than the dry pressed sample.

Fig. 7  SEM micrograph of (a)


­Na3.4Sc0.4Zr1.6(SiO4)2(PO4)
(Reproduced with permission
form ref. no. [33]) and (b)
­Na3.33Ce0.02Sc0.33Zr1.65Si2PO12
(Reproduced with permission
form ref. no. [98])

13
Ionics

Fig. 8  SE and BSE images


of the (a) fractured surface
and (b) ion-milled surface of
NZSP, respectively. SE and
BSE images are also given for
the similar surfaces (c) and (d)
of NZSP with 9.1 wt.% NBO,
respectively. (Reproduced with
permission from ref. no. [93])

Fig. 9  SEM micrographs
of the fractured surfaces of
­Na3.256Mg0.128Zr1.872Si2PO12
specimens (a, c) dry-pressed at
25 °C and (b,d) cold sintered
at 140 °C, respectively. Both
specimens were pressed under
the pressure of 780 MPa for
1 h. (Reprinted with permission
from ref. no. [50])

Both the relative density and ionic conductivity of studied the microstructure of Sc-NZSP and reported that
NASICON electrolytes are improved in the final stage of NASICON-type materials often produce micro-cracks near
the sintering process due to the considerable grain growth. grain boundaries while cooling the material after sintering.
However, micro-cracking inside the large grains and along Although such types of micro-cracks pose an insignificant
the grain boundaries can occur to relieve the residual influence on the bulk density due to their small volume,
stresses formed due to the anisotropic thermal expansion they are detrimental to the conductivity of the sample as
in the NASICON lattice [33, 130]. Ma et al. [33] have these cracks block the transport of Na ions [33].

13
Ionics

Recent progress However, a silica-rich glassy phase exists along the grain
boundaries in the pellets prepared using the calcined
Goodenough and Hong [131] discovered ­Na + ion dif- colloidal powders. In another study, Lalère et al. [148]
fusion in the NASICON skeleton structure present in found a traceable quantity of Z ­ rO2 near the grain bound-
­Na 1+x Zr 2 Si x P 3-x O 12 system. Subsequently, this three- ary region in NZSP synthesized using the sol–gel method,
dimensional rigid framework has gained tremendous atten- which emphasized the importance of optimal processing
tion from the scientific community and become a centre of steps. As a result, the NZSP specimens exhibit a reason-
interest to explore various properties. This is also possible able ionic conductivity of 1.5 mS·cm −1 only at 200 °C.
because of the compositional flexibility and high tolerance Porkodi et al. [149] used an alternative synthesis approach
shown by the NASICON crystal structure for substituting based on molecular precursor for the preparation of NASI-
various cations. As the structure accepts both isovalent CON oxides. This method produces nano-sized precursors,
and aliovalent cation substitution, a wide range of NASI- which may be responsible for phase purity, good densifica-
CON compositions can be synthesised. The present sec- tion, and high ionic conductivity in the resultant materials.
tion describes the recent efforts undertaken to increase Naqash et al. [115] developed a unique solution-assisted
the sodium-ion conductivity of NZSP by doping it with solid-state reaction method for the powder synthesis of
various elements. Moreover, the strategies that have been NZSP. This method promotes cation homogeneity, sup-
employed recently to address the concerns of high sinter- press ­Z rO 2 formation, and produce large-sized grains.
ing temperature, anisotropic thermal expansion, mechani- Thus, a reasonably good ionic conductivity of 1 mS.cm−1
cal stability, and large interfacial resistance of NASICON is found for the NZSP sample at RT. This simple and
electrolytes are also discussed. cost-effective method has also been applied to fabricate
specimens of other NASICON compositions, for exam-
ple, ­Na3+xZr2Si2+xP1−xO12 [146] and ­Na3+xScxZr2-xSi2PO12
Ionic conductivity [33]. Figure 10 shows the total and bulk conductivities
measured for N ­ a 3+xZr 2Si 2+xP 1−xO 12 (where 0 ≤ x ≤ 0.6)
Previously researchers have extensively tried synthesizing samples prepared using the solution-assisted solid-state
single-phase NASICON electrolytes, albeit with little suc- reaction method [146]. A significantly high sodium-ion
cess [84, 86–88]. The primary concern has always been to conductivity of 5.2 mS.cm −1 at RT is reported in the
suppress the formation of secondary phases. An optimal ­Na 3.4Zr 2Si 2.4P 0.6O 12 composition. In recent work, Liu
processing route is crucial in obtaining phase pure solid et al. [145] investigated N­ b5+ doping for the first time into
NASICON electrolytes with improved electrical properties ­Na3.4Zr2Si2.4P0.6O12 by synthesising the powders using the
[132, 133]. The simple processing steps, such as type of SA-SSR method. A remarkably high ionic conductivity
solvent media, milling time, milling speed, and calcination of 5.51 mS.cm−1 is found for N ­ a3.3Zr1.9Nb0.1Si2.4P0.6O12
5+
temperature, effectively regulate the phase formation and composition. Adding ­N b reduces the Na content in
the chemical stoichiometry inside the material. Thus, all ­Na3.4Zr2Si2.4P0.6O12, and an optimal composition for maxi-
parameters are essential and impact the ionic conductivity mum conductivity is seen when the Na amount is 3.3 per
of the solid electrolyte [99, 134]. The majority of research formula unit. This NASICON composition has a favour-
on NZSP has been on the partial elemental substitution of able ratio of sodium ion concentration to vacancy concen-
NASICON structure with heteroatoms such as Zn, Sc, Ce, tration inside the crystal structure, allowing for efficient
Zr, La, etc. [33, 34, 54, 66, 84, 108, 129, 135] The ionic transfer of sodium ions.
conductivity of several NZSP-related compositions, which Several advanced sintering techniques have been
are recently reported and have shown high magnitude, are used for the fabrication of NZSP electrolyte with dense
given in Table  1. At room temperature, the total ionic and fine-grained microstructure. They have fascinat-
conductivity in most of these compounds is in the range ing advantages of moderate sintering temperature,
of ~ 2–4 mS.cm−1. However, in a few NASICON composi- shorter sintering time, and greater yields. Leng et  al.
tions, the magnitude of > 5 mS.cm−1 is also recorded [141, [50, 94] employed a cold sintering technique to fabri-
145, 146]. cate bulk specimens of Mg-doped NASICON electrolyte
The sol–gel method has been widely employed for ­( Na 3.256Mg 0.128Zr 1.872Si 2PO 12). A conductivity of 0.041
the synthesis of NASICON [100, 101]. Zhou et al. [147] mS·cm −1 is reported for the specimen cold sintered at
compared the microstructure of NASICON electrolyte 140 °C and 780 MPa for 1 h. Annealing the same sample
prepared by both the colloidal and sol–gel methods. The at 800 °C for 6 h significantly enhanced the ionic conduc-
sintered pellets fabricated from the calcined sol–gel pow- tivity to 0.5 mS·cm−1, which is five times superior to the
der possess a single-phase (homogenous) microstructure. dry-pressed specimen conventionally sintered at 1000 °C
(0.1 mS·cm−1).

13
Table 1  Compositions, fabrication methods, and ionic conductivity of recently reported NASICON-type electrolytes
Compositions Method/sintering conditions Ionic conduc- Salient features Crystal structure Electrochemical References
tivity at 25 °C window vs Na/

13
(mS·cm−1) Na + (V)

Na3.1Zr1.95Mg0.05Si2PO12 Solid-state reaction method 1.33 • High grain boundary conductivity M 6 [136]
Sintering: 1225 °C (~ 1 mS.cm−1 at RT)
• > 95% relative theoretical density
Na3.256Mg0.0625Zr1.872Si2PO12 Solid-state reaction method 2.7 • Formation of less insulating sec- M + R - [85]
Sintering: 1300 °C, 24 h ondary phase of doped-Na3PO4 is
formed, which reduces the effective
grain boundary resistance
Na3.1Mg0.05Zr1.95Si2PO12 Solid-state reaction method 3.5 • Dense microstructure with well M 4.5 [137]
Sintering: 1260 °C, 16 h crystallized grains is formed
• Enlargement of bottleneck trian-
gular area on doping with alkaline
earth cations
Na3.4Mg0.1Zr1.9Si2.2P0.8O12 Solid-state reaction method 3.6 • Stable cycling of symmetric cell M 6 [138]
Sintering: 1250 °C with Na as electrodes at 1 mA.cm−2
current density
• Excellent dendrites suppression
capability
Na3.4Sc0.4Zr1.6Si2PO12 Solution-assisted solid-state method 4.0 • Chemical homogeneity M + R 6 [33]
Sintering: 1265 °C, 5 h • Cost-effective for large-scale
production
Na3.125Zr1.75Sc0.125Ge0.125Si2PO12 Solid-state reaction method 4.64 • Sc doping improves the stability of R - [72]
Sintering: 1100 °C, 12 h the rhombohedral phase at RT
• Ge doping reduces the monoclinic
to rhombohedral phase transition
temperature
Na3.4Sc2Si0.4P2.6O12 Solid-state reaction method 2.4 • Increase in charge carrier density on R - [139]
Sintering: 1300 °C, 10 h Si substitution
• Enhanced correlated migration in
Si-doped samples
• Enlargement of the oxygen bot-
tleneck triangle T1
Na3.2Hf1.9Ca0.1Si2PO12 Solid-state reaction method 1.07 • ­Ca2+ doping increase both the M 8.8 [30]
Sintering: 1250 °C, 16 h charge carrier concentration and
specimen density
• Mixed ionic electronic conducting
phase is constructed on the surface
of the electrolyte, which ensures fast
charge transfer between Na and SE
Na3.1Zr1.9Ga0.1Si2PO12 Solid-state reaction method Sintering: 1.06 • Enlargement of bottleneck area M 5 [80]
1250 °C,16 h • Negligible electronic conductivity
Ionics
Table 1  (continued)
Ionics

Compositions Method/sintering conditions Ionic conduc- Salient features Crystal structure Electrochemical References
tivity at 25 °C window vs Na/
(mS·cm−1) Na + (V)

Na3.2Zr1.9Ca0.1Si2PO12 Sol–gel method 1.67 • A robust solid-state battery is M 7 [90]


Sintering: 1250 °C, 5 h developed by integrating Na metal
with monolithic Ca-doped NZSP
electrolyte
• Very low interfacial resistance of
the symmetrical cells with stable
sodium stripping/plating cycles at
0.3 mA.cm−2 for 600 h
Na3Zr1.8Zn0.2Si2PO11.8 Solid-state reaction method 1.44 • Lower valence state of the dopant M - [140]
Sintering: 1225 °C, 4 h is beneficial in achieving high bulk
conductivity mainly because of the
less electrostatic interaction with
­Na+
• Excessive doping reduces the con-
ductivity because of cell distortion
Na3.4Zr1.9Zn0.1Si2.2P0.8O12 Solid-state reaction method 5.27 • ­Zn2+ doping enlarged the size of the M + R 6 [141]
Sintering: 1250 °C bottleneck triangle and also raised
the concentration of Na-ions
• Solid-state cell based on SE with
polydopamine coating at the cath-
ode side exhibits a high capacity at
a C/2-rate
Na3.3Zr1.7La0.3Si2PO12 Sol–gel method 1.34 • ­La3+ ions exist in the form of phos- M + R - [142]
Sintering: 1250 °C, 12 h phate impurities, i.e. ­Na3La(PO4)2
• Si/P ratio contributes significantly
towards the conductivity enhance-
ment
Na3.3Zr1.7La0.3Si2PO12 Sol–gel method 3.4 • ­La3+ doping in NZSP generates new - - [34]
Sintering: 1200 °C, 24 h phases of ­Na3La(PO4)2, ­La2O3, and
­LaPO4 resulting in the formation of
composite electrolyte
• The ionic liquid was used on the
cathode side of the SE in SSBs to
serve as a wetting agent and enable
fast charge transfer kinetics
Na3.3Zr1.9La0.1Si2PO12 Solid-state reaction method 1.1 • Doping with rare earth elements - - [143]
Sintering: 1150 °C increases the content and mobility
of the Na- ions inside the lattice
NZSP + 5 wt % Antimony tin oxide Solid-state reaction method 1.43 • 5 wt.% ATO addition significantly M 4.6 [95]
(ATO) Sintering: 1100 °C, 6 h enhanced the densification process

13
Table 1  (continued)
Compositions Method/sintering conditions Ionic conduc- Salient features Crystal structure Electrochemical References
tivity at 25 °C window vs Na/

13
(mS·cm−1) Na + (V)

Na3.2Zr2Si2.2P0.8O12 -0.5NaF Solid-state reaction method 3.6 • NaF assisted grain boundary M + R 4.2 [130]
Sintering: 1150 °C, 24 h modification and improved the ionic
conductivity
Na3Zr2Si2PO12 Solid-state reaction method 1.2 • Na-air cell is fabricated using M - [144]
Sintering: 1275 °C, 15 h NASICON ceramic as a separator
• Cell exhibits an electrochemical
discharge of 600 mAh.g−1, 30% less
than the theoretical value
Na3Zr2Si2PO12 Solid-state reaction method with 1.13 • Only excess Na is added rather than M - [99]
excess sodium both Na and P
Sintering:1100 °C, 12 h
Na3Zr2Si2PO12 Solution-assisted solid-state method 1.0 • The powder processing method is M - [115]
Sintering: 1250 °C, 5 h convenient, low cost and has the
potential for large-scale production
Na3Zr2Si2PO12 Solid-state reaction method 1.8 • The spark plasma sintering method M - [88]
Sintering: 1200 °C was used to fabricate samples
• Highly dense samples were pre-
pared at a relatively low sintering
temperature of 1100 °C
• The formation of the m-ZrO2
impurity phase is suppressed by the
addition of excess Na and P
Na3.3Zr1.9Nb0.1Si2.4P0.6O12 Solution-assisted solid-state reaction 5.5 • The ionic conductivity initially M - [145]
Sintering:1260 °C, 5 h increases and then decreases with
the addition of ­Nb5+
Na3.4Zr2Si2.4P0.6O12 Solution-assisted solid-state reaction 5 • High densification at relatively low M + R - [146]
Sintering:1260 °C, 5 h temperature
• High ionic conductivity because
of the optimal ratio of the vacant
and occupied sodium sites in the
structure
Na3.3Zr1.65Sc0.33Ce0.02Si2PO12 Solid-state reaction method 2.44 • Absence of impurity phase and M - [98]
Sintering: 1250 °C, 5 h microcracks
• Enlarged bottleneck triangular area
Ionics
Ionics

grains and an ionic conductivity of 0.251 mS·cm−1 at RT.


The obtained conductivity value is comparable to that of
the sample formed by solid-state sintering performed at
1200 °C for 720 min.
Numerous attempts have been made over the years to
address the low sinterability issue of NZSP ceramics by
adding a variety of sintering aids. These additives facilitate
sintering at low temperatures, thereby preventing the vola-
tilization of Na and P and shortening the material processing
time. Noi et al. [92] investigated the liquid-phase sintering
of NZSP using N ­ a3BO3 (NBO) as a sintering aid for the
first time to reduce the NASICON sintering temperature
to 900 °C. The dense liquid phase sintered NZSP sample
containing 4.8 wt% NBO exhibits high conductivity of ~ 1
mS·cm−1 at RT with an activation energy of 28 kJ ­mol−1.
Fig. 10  Total (σtotal) and bulk (σb) conductivity of N
­ a3+xZr2Si2+xP1−xO12 The direct bonding of NZSP grains and the segregation of
as a function of x measured at 25 °C. (Reproduced with permission from NBO in particulate form are the main reasons for achieving
ref. no. [146])
high ionic conductivity in the liquid-phase sintered NZSP.
Utilising the liquid phase sintering approach, Oh et al. [70]
da Silva et  al. [119] examined the ionic conduc- fabricated dense NZSP specimens by adding sodium meta-
tivity in the bulk specimens of Sc-doped NZSP silicate as a sintering aid. The N­ a2SiO3 melt pool facilitates
­(Na3.4Sc0.4Zr1.6Si2PO12) prepared using different sintering the densification process and, on solidification, forms a Si-
methods. It was demonstrated that using the rapid sintering rich secondary phase enhancing the grain boundary conduc-
techniques, such as Field Assisted Sintering Technology tivity of NZSP. As illustrated in Fig. 11a, the melt pool also
(FAST) and Spark Plasma Sintering (SPS), sample densi- acts as a source for Na and Si, which diffuse inside the grains
fication (~ 99% of theoretical density) can be achieved at and augment the grain conductivity. Adding 5 wt.% N ­ a2SiO3
lower temperatures than using the conventional sintering significantly raises the relative density of NZSP from 81 to
method, with little variation in electrical properties. On the 93% after sintering at 1175 °C for 10 h and consequently
other hand, the cold sintering technique, which requires sin- improves the room temperature total ionic conductivity from
tering aids and high mechanical pressure, result in moderate 0.64 to 1.45 mS.cm−1 (shown in Fig. 11b).
densification (86.8% of theoretical density) but at tempera- Chen et al. [96] studied the effect of adding indium-tin
tures as low as 250 °C [119]. On annealing the cold-sintered oxide (ITO) on the microstructure and conductivity of NZSP.
specimen at higher temperatures, an increase in both the An optimized ITO addition (3 wt.%) results in the maximum
density and conductivity is noticed. The microwave sin- of both the grain and grain boundary conductivities in NZSP.
tering technique has also been employed to prepare dense The conductivity enhancement is a direct consequence of
NZSP ceramics at low temperatures [97]. The microwave the dense microstructure, uniform grain size, close contact
sintered NZSP specimen fabricated at 850 °C for 30 min between grains, and the reduction in electrostatic repulsion
possesses a high relative density of 96% with fine-sized between the ­In3+ and ­Na+. It is suggested that the excess ITO

Fig. 11  (a) Schematic of the


melt pool present during the
sintering of NZSP with 5 wt.%
­Na2SiO3. The arrow indicates
the diffusion of Na and Si from
the melt pool to the NASICON
phase. (b) The effect of N
­ a2SiO3
addition on the total ionic con-
ductivity and density of NZSP is
also shown. (Reproduced with
permission from ref. no. [70])

13
Ionics

present in the grain boundary favours high grain boundary The maximum grain and total ionic conductivity of 1.43
conductivity. Antimony-tin oxide (ATO) has also been used and 1.1 mS·cm−1, respectively, at 25 °C are reported for
as an additive to reduce the sintering temperature of NZSP the ­Na3.1Zr1.9La0.1Si2PO12 sample. Ortiz-Mosquera et al.
and boost its conductivity [95]. The NZSP ceramic with an [150] studied ­Na1+xGe2(SiO4)x(PO4)3−x (0 ≤ x ≤ 0.8) based
optimized 5 wt.% ATO sintered at 1100 °C for 6 h exhibits glass–ceramics exhibiting NASICON structure and possess-
a high conductivity of 1.43 mS·cm−1 at RT and good elec- ing a lowest activation energy for x = 0.6 composition. The
trochemical stability up to 4.6 V (vs. Na/Na+). The dense superior ionic conduction is a consequence of the bottleneck
microstructure with uniform and fine-sized grains is respon- size expansion caused by the partial replacement of a larger
sible for observing fast conduction in the grain boundaries of ­Si4+ ion (0.26 Å) for ­P5+ (0.17 Å).
these specimens. In recent work, Zhang et al. [34] modified The valence state and the amount of the substituting ele-
the grain boundaries by introducing ­La3+ in ­Na3Zr2Si2PO12, ment are also crucial parameters influencing the conductiv-
resulting in the formation of several secondary phases such ity of the doped-NZSP materials. Both of them govern the
as ­La2O3, ­LaPO4, etc. The ion transport along the grain electrostatic interactions between ­Na+ and the doping cation,
boundaries is suggested to be facilitated due to the space- which dictate the mobility of the sodium ion within the lat-
charge effect. tice. Chen et al. [140] studied the impact of the valence state
Several previous investigations have focused on enlarg- and the content of dopants ­(Nb5+, ­Ti4+, ­Y3+, and ­Zn2+) on
ing the triangular bottleneck area through which the sodium the crystal structure and conductivity of doped NZSP ceram-
ions migrate inside the NASICON skeleton structure. This ics. It is shown that the dopants with a low valence state (i.e.,
smallest cross-section area of the sodium-ion passageway ­Zn2+) are beneficial for bulk conductivity improvement pri-
imposes geometrical constraints and, thereby, determines marily because of the minimal coulombic repulsion between
the activation energy involved for the sodium ion conduc- the dopant ion and the Na-ion. The highest conductivity
tion inside the NASICON lattice. Park et al. [81] studied of 1.44 mS·cm−1 is measured in ­Na3Zr1.8Zn0.2Si2PO11.8.
the grain and grain boundary diffusion behaviour at low and An optimal doping amount for achieving maximum con-
high temperatures for the bare and 10 at.% Na excess NZSP. ductivity lies between 5 and 10 mol.%. Adding excessive
Based on both experimental and computational results, it is doping amount reduces the conductivity on account of sig-
shown that the grain boundary conductivity is the limiting nificant lattice distortion. Shen et al. [138] optimized the
factor of the total conductivity in NZSP at lower tempera- ­Na3.4Mg0.1Zr1.9Si2.2P0.8O12 composition by systematically
tures (< 100 °C). On the other hand, at higher temperatures replacing the P ­ 5+ with ­Si4+ and ­Zr4+ with ­Mg2+. The bulk
(> 100 °C), the total conductivity is dictated by the grain specimen of this composition fabricated via solid-state reac-
conductivity. In addition, the DFT calculations analysis tion route exhibits an ionic conductivity of 3.6 mS·cm−1 at
reveals that incorporating 10 at.% excess Na expands the RT, which is 17 times superior to the pristine NZSP. The
triangular bottleneck area and improves the thermodynamic excellent conductivity seen in this composition is due to
stability of the material. This leads to a significant enhance- higher sodium-ion concentration and the larger space avail-
ment in the ionic conductivity of NZSP (e.g., 120 mS·cm−1 able for ­Na+ migration.
at 300 °C) [81]. Using multiple computational studies, Zhang et al. [77]
Song et al. [137] observed that the cell volume and unit predicted the correlated migration mechanism inside NZSP,
cell parameters of NZSP are increased by partially substitut- and showed that increasing sodium-ion concentration acti-
ing ­Zr4+ with alkaline earth ions ­(Na3.1Zr1.95M0.05Si2PO12 vates the correlated migrations due to ­Na+-Na+ electrostatic
(M = Ca, Mg, Sr, Ba). Such substitution leads to a sig- repulsion. These excess sodium ions are forced to occupy
nificant increase in the bottleneck area of NZSP, although the high energy sites which lead to the reduction in energy
only a minor variation exists for different alkaline earth barrier for diffusion. Santhoshkumar et al. [139] employed
ions. Superior ionic conductivity of 3.5 mS·cm−1 at 25 °C several structural characterization techniques to investigate
is measured for Mg-doped NZSP, and it is ascribed to the the structure-conductivity correlations in scandium-based
formation of the largest bottleneck area (~ 6.522 Å2) com- NASICON compounds, namely N ­ a3+xSc2SixP3-xO12 (x = 0.0,
pared to the bare NZSP composition (5.223 Å2). Similarly, 0.2, 0.4, 0.8) prepared using the solid-state reaction method.
Ruan et al. [143] examined the enhancement in the ionic Maximum ionic conductivity of 2.4 mS·cm−1 at 30 °C is
conductivity of NZSP by partial doping with rare earth ele- reported for ­Na3.4Sc2Si0·4P2·6O12 NASICON material. This
ments ­(La3+, ­Nd3+, and Y ­ 3+) for Z
­ r4+. The Na-ion conduc- is attributed to the simultaneous increase in carrier density
tivity inside the lattice is improved because of the rise in and the ratio of vacant Na sites over occupied Na sites on Si-
sodium concentration to compensate for the charge imbal- doping, larger area of the oxygen triangle T1, and enhanced
ance created by acceptor cation substitution. In addition, correlated migration associated with a higher Na-ion con-
the larger size of the doping element has a positive effect on centration. A new type of Ca-doped Hf-based NASICON
the bottleneck area, resulting in conductivity improvement. electrolyte ­(Na3.2Hf1.9Ca0.1Si2PO12) has also been examined

13
Ionics

[30]. This composition possesses an ionic conductivity Ran et al. [72] have demonstrated the enhancement in the
of ~ 1.07 mS·cm−1 at 25 °C and a stable electrochemical grain and grain boundary conductivities through Ge and Sc
window up to 8.8 V (vs Na/Na+). The excellent conductiv- co-doping in NZSP. An exceptionally high conductivity of
ity found in this electrolyte is ascribed to the larger amount 4.64 mS·cm−1 is achieved at RT in the co-doped composi-
of mobile ­Na+ inside the lattice, higher densification due to tion of N
­ a3.125Zr1.75Sc0.125Ge0.125Si2PO12. It is reported that
Ca-doping, and improvement in grain boundary conductivity the ­Sc3+ doping stabilizes the rhombohedral phase at RT,
associated with the chemical composition changes near the while adding ­Ge4+ lowers the temperature at which the phase
grain boundary region. transform from monoclinic to rhombohedral.
Stabilization of high-temperature rhombohedral phase In summary, multiple strategies have been tested to
and lowering the rhombohedral-monoclinic phase transition enhance the sodium ion conductivity in NZSP electrolytes
temperature in NZSP have been tried previously, employing at room temperature. Most of them are focused on enlarging
various strategies. The rhombohedral phase is more symmet- the oxygen triangle bottleneck area, improving the sinter-
ric than monoclinic and thus exhibits higher Na-ion mobil- ability, increasing the charge carrier concentration, modi-
ity. The stability of this phase at a lower temperature may fying the grain boundaries, suppressing the formation of
augment the ionic conductivity of the NZSP-based com- secondary phases, stabilizing the high-temperature rhom-
pounds. Park et al. [151] observed that the phase transition bohedral phase, etc. The processing approach to synthe-
temperature plays a vital role in improving the sodium ion size nano-sized powder with good chemical homogeneity
conductivity in NASICON. An enhanced volume fraction using the cost-effective solution-assisted solid-state reaction
of high-temperature stable rhombohedral phase is achieved method appears to be most promising. A few studies have
at a lower temperature of 150 °C by partial substitution also reported achieving high conductivity of > 5 mS.cm−1
of Ge in NZSP ­(Na3Zr2-δGeδSi2PO12, where δ = 0.1, 0.2). at RT in specimens of different NASICON compositions
The Ge-doped specimens exhibit an ionic conductivity of fabricated using this method. However, more detailed work
14 mS·cm−1 at 150 °C, almost two times superior to that is required to elucidate the correlations between the process-
of the bare NZSP sample. A more recent study suggested ing parameters, chemistry, and conductivity in these oxides.
that adding NaF into the precursor induces phase transition
inside the grains of NZSP from monoclinic to rhombohedral Thermal expansion and microcracks
phase and simultaneously modifies the grain boundaries by
forming a “binder-like” glassy phase [130]. This simple and The existence of microcracks within the microstructure has
low-cost approach enhances the ionic conductivity of NZSP a detrimental effect on the conductivity and mechanical
from 0.45 to 1.7 mS·cm−1 (for ­Na3Zr2Si2PO12-0.7NaF). properties of the solid-state electrolyte. It is well reported
Most of the earlier investigations rely upon doping with that the micron-sized cracks in the NZSP-type materials
a single element for the ­Zr4+ site to improve the sodium- are related to the anisotropy of the axial thermal expansion
ion conductivity in NZSP. Recently, a few co-doping [153, 154]. The microcracking of ceramics occurs during the
approaches have also been adopted in NZSP, which take cooling cycle of the final sintering stage of the electrolyte
advantage of the synergy between two different dopant cati- fabrication. The thermal expansion coefficient (CTE) is a
ons to improve conductivity. Pal et al. [152] employed a fundamental material property that can be accurately meas-
co-doping scheme using ­Sc3+ and ­Yb3+ as co-dopants for ured in all the crystallographical directions using the high-
­Zr4+ to enhance the sodium-ion conductivity of NZSP. temperature XRD technique. Oota et al. [155] employed this
The ­Na3.33Zr1.67Sc0.29Yb0.04Si2PO12 composition consists technique to study the compositional dependence of crystal-
of both monoclinic and rhombohedral phases at RT. This lographic thermal expansion in N ­ a1+xZr2SixP3-xO12 (x = 0,
composition exhibits a conductivity of 1.62 mS·cm−1 at RT, 0.5, 1.0, 1.5, 2.0, 2.5, 3.0). Figure 12 shows a large thermal
which is more than 2.5 times higher than that of NZSP (0.61 expansion anisotropy between the a and c lattice parameters
mS·cm−1). The rhombohedral phase stabilization, higher for this compositional series. As the x value increases, the
mobile sodium content, presence of microstructure with thin slope of the curve of the a-axis changes from negative to
and micro-cracks free grain boundaries, and the suppres- zero. On the other hand, the CTE value of the c-axis for all
sion of m-ZrO2 impurity phase formation are the main rea- the compositions is positive. The coefficient value initially
sons for achieving high ionic conductivity in the co-doped declines with the rise in x value, and after reaching minima
samples. In a similar work, Thirupathi et al. [98] found that at x = 2, it starts increasing. Nevertheless, the overall lat-
co-doping of 1  mol.% C ­ e4+ in N­ a3.33Zr1.67Sc0.33Si2PO12 tice volume thermal expansion is positive for all the tested
improves the sodium conductivity from 0.96 of 2.44 compositions. Similarly, Kaus et al. [156] reported a strong
mS.cm−1. The increase in bottleneck triangular passage anisotropic thermal expansion behaviour in NASICON-type
area caused by a slight C­ e4+ addition is the possible reason ­Na3.4Sc2(SiO4)0.4(PO4)2.6. This composition exhibits a sig-
for the observed high conductivity. In more recent work, nificant expansion of 2.1% along the crystallographic c-axis

13
Ionics

Fig. 12  Unit-cell parameters (a
and c) and unit-cell volume of
­Nal+xZr2SixP3-xO12 as a function
of temperature. Errors bars indi-
cates one standard deviation.
(Reproduced with permission
from ref. no. [155])

in the temperature regime of 25 to 600 °C. However, no conductivity of the sample by blocking the transport of
significant expansion (less than 0.1%) is observed along the Na ions. Guin et al. [159] investigated the solid-solutions
crystallographic a-axis. The change observed in the lattice of ­Na3+xSc2SixP3−xO12 with 0.05 < x < 0.8 fabricated using
constants is reversible, and on cooling from 600 °C to RT, the solid-state reaction method. The samples with a rela-
the original lattice parameters are obtained. tive density of more than 97% exhibit conductivity 3 to 4
Alamo et al. [157] studied the thermal expansion proper- times lower than the specimens having a density between
ties of the compositions present in the ­Na1+xZr2P3-xSixO12 90 to 96%. This is ascribed to the excessive grain growth in
(0 ≤ x ≤ 1) series. The results revealed the compositional the highly dense samples. The microcracking is observed
dependency and the lowest CTE value observed for x = 0.33. in the middle of the large-sized grains (~ 100 µm). The
Using the acoustic emission method, Srikanth et al. [153] thermal expansion anisotropy induces large residual stress
examined the microcracking induced by anisotropic ther- in the material on cooling, resulting in microcracks. How-
mal expansion in several low thermal expansion materials, ever, microcracking primarily occurs at grain boundaries
including ­NaZr2P3O12. A direct correlation between micro- for ceramic specimens possessing large grains and high
cracking and the anisotropic axial CTE is established in this thermal expansion anisotropy [160].
study by detecting the formation of microcracks via intense Multiple investigations reported that microcracking
acoustic emission. arises along the grain boundaries when the grains exceed
Ma et  al. [33] observed that NASICON-type com- a critical size in polycrystalline NASICON materials
pounds often produce micro-cracks near grain boundaries [161]. These cracks are generated due to the dissimilari-
while cooling the material after sintering. Microcrack- ties in contraction during the cooling cycle of the sin-
ing is attributed to the anisotropy in the axial thermal tering process. According to Cleveland and Bradt [160],
expansion [153, 158]. Even though these micro-cracks there exists a critical grain size (Gs) for the occurrence of
have a negligible effect on the density of the material microcracking, which can be calculated using the follow-
due to their smaller size, they significantly reduce the ing relationship.

13
Ionics

NZSP. Nevertheless, the intergranular and intragranular


cracks exist but only in large grains (shown in Fig. 13).
The thermal expansion curves of a few ­Na1+xZr2SixP3-xO12
are compared in Fig. 14. The thermal expansion coefficient
varies from negative to positive value with an increase
in Na content. The negative CTE value observed in
­NaZr2P3O12 is ascribed to the microcracking induced in
the sintered specimen by the large thermal expansion ani-
sotropy. The densely sintered sample of 2 wt.% MgO con-
taining ­NaZr2P3O12 shows a near-zero thermal expansion
coefficient as it possesses fine-sized grains, which suppress
the formation of microcracks. However, MgO addition
reduces the positive CTE value of the N ­ a3Zr2Si2PO12 spec-
imen from 3.9 × ­10–6 to 2.1 × ­10–6 /K, most likely because
of the higher bulk density of the MgO added specimen.
Pal et al. [152] studied the Yb and Sc co-doped NZSP
Fig. 13  Intragranular microcrack present in the densely sintered spec-
imen of N ­ aZr3P3O12 containing 2 wt.% MgO. The specimen was sin- samples fabricated using the conventional solid-state reaction
tered at 1200℃. (Reproduced with permission from ref no. [106]) method. No evidence of microcracks in the microstructure
has been reported in the sintered specimens. The presence
of small-sized grains (∼1.5 μm) in the microstructure is sug-
gested to be the main reason as the refined grains reduce the
effect of thermal expansion by absorbing the stresses in the
grain boundaries and mitigating the formation of microcracks.
Ortiz-Mosquera et  al. [150] studied the microstruc-
ture of the fractured surface of ­Na1+xGe2(SiO4)x(PO4)3−x
(0 ≤ x ≤ 0.8) fabricated using a glass–ceramic route. The
average grain size increases with Si content from 75 nm for
x = 0 to 22 μm for x = 0.8. The micrograph of the sample
with high Si content (x = 0.8) shows the existence of micro-
cracks. These microcracks are formed during the crystalliza-
tion process when the tensile stresses are generated because
of the difference in the CTE values of the crystalline phase
and the parent glass matrix.
Fig. 14  Thermal expansion behaviour of the sintered specimen of In conclusion, microcracking is a critical issue that
­Na1+xZr2SixP3-xO12 compounds: (a) x = 0, (b) x = 0 with 2 wt.% MgO, severely diminishes the ionic conductivity of NASICON-
(c) x = 1 (d) x = 2, (e) x = 2 with 2 wt.% MgO. (Reproduced with per- type oxides. High thermal expansion anisotropy present in
mission from ref. no. [106]) these materials is responsible for the formation of micro-
cracks during the cooling cycle of the sintering process.
14.4 × 𝛾f It is evident from the previous studies that the grain size,
Gs = 2 (1) chemical composition and additives play a crucial role in
E × Δ𝛼max × ΔT 2
suppressing microcracking along the grain boundaries. Vari-
where 𝛾f is the fracture surface energy for grain boundary ous dopants (such as M ­ g2+, ­Sc3+, etc.) have shown immense
microcracking, Δαmax is the maximum difference in CTE of potential in reducing microcracks by lowering grain growth
a and c crystal axes, ΔT is the temperature change to initiate and promoting densification. Nevertheless, further research
microcracking during cooling, and E is an elastic modulus. is required to systematically study the role of dopants on the
Using this relationship, Bradt [162] reported the value of GS thermal expansion coefficient, which controls the amount of
ranging from 2 to 14 μm for the NZSP materials. microcracking in NASICON materials.
Dhas et al. [106] studied the effect of MgO addition on
the microstructure of NZSP synthesized using the con- Mechanical properties and electrode–electrolyte
trolled combustion process. The 2 wt.% MgO addition interface
not only facilitates densification but also lowers the grain
boundary microcracking by suppressing grain growth in The degradation of the mechanical strength of a solid elec-
trolyte, especially under cyclic charging, is a critical factor

13
Ionics

affecting the lifetime of solid-state batteries. During the determine the fracture strength in brittle solids. The indenta-
charge–discharge cycles, intercalation/de-intercalation of tion has a considerable compressive hydrostatic component
ions inside the electrode structure induces an enormous which prevents crack propagation and avoids fracture. More
cyclic volume change in the electrode. This dynamic phe- recently, this technique has been applied to battery materials
nomenon directly influences the chemical, electrical and as well.
mechanical properties, which possibly leads to microcrack- Kalnaus et al. [167] reported the correlations between
ing, formation of structural defects, dendrites formation, ionic conductivity and mechanical properties in
etc., at the electrode/electrolyte interface and within the solid ­Na1+xMnx/2Zr2−x/2(PO4)3 (x = 0.5, 1.0, 1.5, and 2). A reduc-
electrolyte [30, 138]. Understanding the elastic property of a tion in both the elastic modulus and hardness is observed
SE is crucial as the electrolyte must accommodate internal with an increase in x value (as shown in Fig. 15a). This
stresses by deforming elastically to maintain a stable elec- decrease appears to be non-linear, with a significant drop
trode/electrolyte interface. (almost resembling a step function) as the x value increases
Young modulus and shear modulus are the two most from 1 to 1.5. It is of practical interest to note that Hill's
important parameters to study the electrolyte resistance equation, as illustrated by the dashed lines in Fig. 15a, can
against the deformation from external pressure. Young describe the link between elastic modulus and nanoinden-
modulus measures the stiffness of the solid material, and its tation hardness and composition. The reduction in hard-
value depends on the bonding and the crystal structure of ness is advantageous for material processing, as it implies
the material. NASICON electrolytes exhibit Young’s modu- reduced resistance to local plastic strain. Figure 15b shows
lus in the range of 72–88 GPa. Min [163] summarized the the effect of Na content on fracture toughness and conduc-
elastic properties of sodium-ion solid-state conductors for tivity. The fracture toughness (KIC) initially decreases and
next-generation NIBs using existing data and first-principle then increases with the rise in the sodium content, reaching a
calculations. NASICON-type oxides exhibit the highest elas- maximum value of 0.89 M ­ Pam1/2 at x = 2. The observed non-
tic properties and low anisotropic behaviour among vari- linear relationship is attributed to the formation of a glassy
ous studied materials. According to the stability criterion phase at an intermediate Na concentration. The maximum
of Monroe and Newman [164], the shear modulus of the ionic conductivity (∼2.85 × ­10–2 mS·cm−1) present in this
SEs must be twice or greater than that of sodium metal (3.3 compositional series coincides with the maximum fracture
GPa) to prevent dendrite penetration inside the electrolyte toughness. When this material is utilised as a sodium-ion
[165]. The shear modulus of NASICON ceramic electrolytes conducting phase in composite electrodes, the combination
is 44.3 GPa, which is much higher than that of Na-metal. of high fracture toughness and low hardness in the opti-
Fracture toughness is another crucial mechanical prop- mal composition of x = 1.5 promotes the creation of stable
erty of SEs as it is a measure of the resistance to fracture interfaces.
that could occur during the cell operation. It governs the Valle et al. [168] investigated the mechanical properties
crack growth, a succession that leads to the short-circuiting of hot-pressed von Alpen-type NASICON electrolyte of
of SSBs [166]. Nano-indentation is a useful technique to composition ­Na3.1Zr1.55Si2.3P0.7O11 for NIBs. The fracture

Fig. 15  Correlations between the ionic conductivity, mechanical in (a) while the fracture toughness (KIC) and room temperature con-
properties, and composition of ­Na1+xMnx/2Zr2−x/2(PO4)3 with x vary- ductivity are shown in (b). (Reproduced with permission from ref.
ing between 0.5 and 2. The elastic modulus and hardness are given no [167].)

13
Ionics

toughness of specimens fabricated using spray-dried powder of sodium metal on the NASICON electrolyte, resulting in
ranges from 1.05 to 1.09 ­MPam½. Furthermore, the elastic better electrolyte-anode interfacial contact.
modulus in these samples is reported to be 97.65 GPa. The The high resistivity of the electrode and SE interface is
obtained values are found to be dependent on the microstruc- also linked to chemical and electrochemical compatibility
tural aspects, viz., grain size, the fraction and distribution of problems at the interface. Experimental investigations have
glass phase at grain boundaries and ­ZrO2 content. been unsuccessful in ascertaining the exact mechanisms by
Nonemacher et  al. [166] studied the micromechani- which the high resistance develops at the interface. This is
cal properties of Al/Y substituted NASICON with the primarily because of the difficulty accessing a thin interface
compositions of ­Na1+2xAlxYxZr2-2x(PO4)3 (NAYZPx) and of two solid materials where the reaction occurs. DFT calcu-
­Na3+2xAlxYxZr2-2x(SiO4)2(PO4) (NAYZSiPx). The NAYZPx lations have been found to be an effective tool for conducting
compositions crystallizes with rhombohedral symmetry, systematic studies on the chemical and electrochemical com-
while NAYZSiPx transforms from monoclinic to rhombo- patibility at the SE and electrode interface. Using this com-
hedral phase depending on the doping level and temperature. putation method, Zhu et al. [172, 173] studied the chemical
The NAYZPx exhibits elastic modulus and hardness val- and electrochemical stabilities of SE and electrode interfaces
ues in the range of ~ 72–82 GPa and ~ 4.8–5.8 GPa, respec- in solid-state Li-ion batteries. It was shown that most solid
tively. On the other hand, the elastic modulus and hardness electrolytes, including NASICON, are not thermodynami-
in NAYZSiPx lie in the range of ~ 72–88 GPa and ~ 5.6–7.6 cally stable. Interfacial reactions can occur either during
GPa, respectively. The fracture toughness values for both the heat processing or during the charge–discharge cycling of
compositional series are in the range of 1.30–1.58 MPa ­m½. a battery. The interface decomposition to form interphase
Yde-Andersen et al. [127] measured the average fracture layers is thermodynamically favourable. However, the slug-
strength of ­Na 2.94Zr 1.49P 0.8Si 2.2O 10.85 using the three- gish kinetics of decomposition and poor electronic conduc-
point bending method. The specimens sintered at 1200 °C tivity of decomposition interphases are predicted to be the
exhibit an average fracture strength of 162 MPa. Accord- reasons for observing good stability experimentally. It was
ing to another study, the average flexural strength of fully further suggested that the decomposition interphases with
dense NASICON bars tested using the four-point flexural high electronic insulation and good ionic conductivity could
method is 98 MPa, which is lower than that of β″-alumina be advantageous in obtaining high interface stability [173].
(250–300 MPa) [169]. Applying artificial interfacial coating layers with such prop-
The realization of solid-state NIBs is mostly hindered by erties is proposed as an effective strategy for stabilizing the
low ionic conductivity of SEs and large electrode–electro- interface.
lyte interfacial resistance [46, 170]. In general, the effective In another computational investigation, Lacivita et al.
contact area between the electrode and the SE is insufficient [174] performed the DFT calculations to study the electro-
compared to the electrode and liquid electrolyte. Thus, a chemical stability window of sodium-ion conducting solid-
good interfacial contact between the SE and electrode are state electrolytes, the kinetic voltage limits for Na extraction
required to minimize resistance and develop a high-capacity and the chemical reactivity with various cathode materials
sodium-ion battery. Employing a metallic anode can signifi- commonly used in NIBs. All NASICON materials with
cantly raise the energy density of a battery cell. Although the ­[PO4]3− exhibit excellent anodic stability, resisting oxida-
metallic anode can be stripped and plated reversibly, several tion up to 5 V. However, they all tend to decompose at low
challenges exist when using a solid electrolyte. Uchida et al. potential. Moreover, no NASICON composition is thermo-
[171] have successfully improved the interfacial contact and dynamically stable in contact with Na metal, which reduces
lowered the resistance to 14 Ω ­cm2 at room temperature by the polyanion groups, i.e. ­[SiO4]4− and ­[PO4]3−. The pure
applying a simple mechanical compression load of 30 MPa phosphate containing NASICONs are found less stable than
on a Na/NASICON assembly. The activation energy asso- those having ­[SiO4]4−. Furthermore, some chemical reactiv-
ciated with the charge transfer at Na/NASICON interface, ity is found only for Si-containing NASICON compositions
estimated by plotting the interfacial resistance versus 1/T, with partially de-sodiated electrodes.
is found to be 0.6 eV which is far greater than that of Na/β′′- In a recent experimental study, Goodenough and co-
alumina (0.28 eV). This higher activation energy magnitude workers [56] revealed that the insertion of a dry ­Na+ con-
is attributed to the interphase layer formed by pressing the ductive thin polymer film or the interlayer formed from an
Na metal on the NASICON surface at room temperature. in situ reaction with molten Na could dramatically reduce
The outer layer of NASICON is postulated to react with the electrode/SE interfacial resistance. Superior wetting of
Na metal at room temperature, creating a ­Na+ conducting sodium metal on the NASICON electrolyte is responsible for
interphase layer. This thin passivation layer also serves as the improved interface as it prevents the formation of den-
a barrier for further reaction of NASICON with Na metal. drites near the anode and their rapid penetration inside the
Besides, the presence of this interlayer improves the wetting SE along the grain boundaries during the charging process.

13
Ionics

The schematic of the dendrite formation and penetration [175] demonstrate a reduction of the interfacial resistance
inside the inorganic electrolyte when the Na metal wetting of Na|NASICON from 1658 to 101 Ω·cm 2 by lowering
is poor is depicted in Fig. 16. The poor physical contact will the surface tension of Na metal on forming composites
result in non-uniform coating and the rapid short-circuit of with amorphous S ­ iO 2. A uniform distribution of S­ iO 2
the cell. On the other hand, when the ceramic surface is particles within the composite improved the wettability
modified by the presence of an artificial interlayer and is of sodium. The Nyquist plot collected on the symmetric
easily wetted by Na metal, the unwanted dendrite nuclea- cell using Na metal or Na-SiO2 as electrodes is shown in
tion is suppressed owing to the uniform N ­ a+ flux across the Fig. 17a. The reduction of interfacial resistance is clearly
interface. It is important to note that in this case the grain visible for cell based on Na-SiO2 electrodes. Further, the
boundaries of the ceramic electrolyte are not directly in con- solid-state Na-SiO2|NASICON|Na-SiO2 symmetric cell
tact with the sodium anode. A significant improvement in exhibits a stable Na stripping/plating and can withstand
the cycling performance and high efficiency of the solid- current density up to 500 μA/cm 2 (Fig. 17b) and cycling
state sodium battery can be achieved using both the above- test for more than 135 h (Fig. 17d). On the other hand, the
mentioned strategies. In more recent work, Tian et al. [30] Na|NASICON|Na symmetric cell rapidly degrades in less
reported a stable electrochemical performance of solid-state than 10 h at 100 μA/cm2 (shown in Fig. 17c). The Na-SiO2
NIBs for 2400 h at RT with the current density as high as composites facilitate intimate physical contact with the
2 mA.cm−2. In this work, a mixed ionic-electronic interface NASICON electrolyte, resulting in a dramatic reduction in
of ­Na2O/NaxSn was in situ constructed on the surface of the interfacial contact and significant enhancement in the
­Na3.2Hf1.9Ca0.1Si2PO12. The interlayer provides close physi- electrochemical stability of solid–solid interface.
cal contact between the SE and electrode and homogenous In a recent study, Lu et al. [90] reported a high-perfor-
deposition/dissolution of sodium at the interface. The full mance monolithic solid-state NIB in which molten sodium
cell based on modified electrolyte and ­Na3V2(PO4)3 and Na ­ nO2 decorated porous NASI-
metal is infiltrated into the S
as electrodes exhibits a capacity of 103.1 mAh.g−1 at 0.5 C CON electrolyte. The SE consists of a dense layer sand-
even after 300 cycles. wiched between two porous layers of thickness ~ 100 μm.
Improving the wettability of sodium on NASICON- The sodiophilic nature of ­S nO 2 enables sodium melt
based electrolytes by making composites of sodium with to easily wet the inorganic electrolyte. This interfacial
some other material is another approach reported in the resistance is significantly reduced with a stable cycling
literature to achieve stable operation in SSBs. Fu et al.

Fig. 16  Schematic of the
sodium metal/solid electrolyte
interface where the sodium
metal possesses (a) a poor
wetting ability and (b) a good
wetting ability on ceramic
pellet. An artificial interlayer is
formed when the ceramic pellet
is easily wetted by sodium metal
(Reproduced with permission
from ref. no. [56])

13
Ionics

Fig. 17  Electrochemical
performance of symmetric
cells fabricated using Na metal
or Na-SiO2 as electrodes and
­Na3.2Zr1.9Mg0.1Si2PO12 as SE.
(a) EIS spectra, (b) Na strip-
ping/plating at current densities
ranging from 10 to 500 μA/cm2,
and (c) cycling tests at current
density 100 μA/cm2 of both the
cells are compared. The cycling
performance of symmetric
cell with Na-SiO2 electrode at
current densities of 0.1 mA/cm2
and 0.2 mA/cm.2 is also shown
(d). All the measurements were
conducted at room temperature.
(Reproduced with permission
from ref. no. [175])

performance is observed at different current densities from significant challenge for developing all-solid-state sodium
0.1 to 0.3 mA ­cm−2. batteries.
Unlike the Na metal anode, the battery cathode is com- In comparison to cathode/SE chemical compatibility,
posite in nature and generally comprises electroactive more emphasis has been placed on improving cathode/SE
material, SE, binders, and an electronically conductive interfacial interaction. Mechanically mixing the particles of
agent, making sodium-ion exchange process between the SE with active material during the electrode preparation pro-
cathode and the electrolyte rather complicated. Layered cess is a typical approach employed to improve the cathode/
oxides [176, 177], tunnelled structured oxides [178, 179], electrolyte interaction. Recently, wet infiltration has been
and polyanionic compounds [180–182] are some of the used to construct the composite electrode structure [183].
most commonly used electroactive materials for cathode, The intimate contact between the solid electrolyte and elec-
which are also similar to NASICON-based SEs in terms trode is achieved by infiltrating the precursor solution of
of stiffness. As a result of the low wettability of these the electroactive material into the porous side of the dense-
materials, the interfacial contact between the cathode and porous bilayer NZSP electrolyte. An effective ion transfer is
inorganic electrolyte is inadequate, resulting in a lack of attained with minimum stresses resulting from the volume
effective ion transport channels and high interface imped- change of the electroactive material during the charge–dis-
ance. Moreover, the volume change associated with the charge cycling.
electroactive particles during the charge/discharge process In addition, the introduction of various liquid solvents
further damages the interfacial contact. Thus, improving (such as organic liquid electrolyte, ionic liquid, etc.) between
the physical connection between the cathode and SE is a SE and electrode has also been reported for augmenting the

13
Ionics

interfacial contact [34, 184]. Combining the benefits of inor- metal on the NASICON surface has a strong influence on the
ganic ceramic electrolyte and organic polymer electrolyte, formation and growth of dendrites [56, 187] Since Na metal
a ceramic-polymer composite approach has also been used does not wet the inorganic oxide electrolyte surface and has
to fabricate SEs with improved ionic conductivity, good poor physical contact with SE, a non-uniform Na-ion flux
mechanical property and high chemical/thermal stability exists across the interface. The Na ions plate preferentially at
[185]. Using this approach, solid electrolyte can accommo- the grain boundaries of the ceramic electrolyte where local
date the cyclic volume changes occurring in the cathode ­Na+ flux is significant under the electrical field. This causes
during the charge–discharge process. Plcharski et al. [186] the nucleation of dendrites which grow along the grain
investigated the addition of NASICON powder to PEO-NaI boundaries and penetrate the NASICON-type electrolytes.
solid electrolytes and found that the conductivity of PEO- The addition of interfacial wetting agents and inserting a
based composite solid electrolytes containing NASICON buffer layer have been explored in the literature to resolve
is greatly increased when compared to pristine PEO-NaI this issue. These efforts have reduced the interfacial resist-
composites. Zhang et al. [53] integrated the polyethylene ance to some extent [30, 56, 188]. Nevertheless, the volume
oxide/NASICON composites to be used as a SE in NIBs. change associated with the solid electrode during the charg-
The composite electrolyte exhibits a conductivity of 2.4 ing-discharging cycles needs to be accommodated by SE for
mS.cm−1 at 80 °C and is stable up to 150 °C. The polymer maintaining the close connection with the solid electrode.
provides flexibility to the composite, and the ceramic phase Developing interlayers with multiple functionalities of high
renders the high conductivity. The solid-state battery based conductivity, good sodium wettability, and adequate flex-
on composite electrolyte, N ­ a3V2(PO4)3 and Na as electrodes ibility, which can provide intimate physical contact between
showed an initial reversible capacity of 106.1 mA h ­g−1 and the electrode and SE and reduce the interfacial resistance,
excellent cycling performance over 120 cycles. In a recent remains the work for the future.
investigation, Yang et al. [141] used a thin coating of poly- Another challenge associated with the NASICON based
dopamine on the cathode side of NASICON electrolyte to electrolyte is the high resistance offered by grain bound-
attain a highly reversible capacity in SSBs. aries. For reducing the grain boundary resistance, a high
In summary, the interfacial modification for the intimate sintering temperature (≥ 1200 °C) is essential, which could
physical contact between electrodes and SEs is imperative further lead to challenges in material processing (for exam-
to achieve low interfacial resistance and high cyclability in ple, secondary phase formation, evaporation of Na and P,
SSBs. Enhancing the wetting of sodium metal on NASICON microcracking, etc.). The complex and expensive techniques
electrolytes by introducing interfacial interlayers, forming for synthesizing phase pure NASICON electrolytes are not
composites, and giving heat treatment for uniform spread- favourable for large-scale production. As a result, a cost-
ing of molten Na onto the SE are a few strategies employed effective and simple synthetic approach must be developed
for Na anode/SE interface modification, resulting in high to realize the bulk production of high-performance solid
efficiency in SSBs. The challenge of high resistance in the electrolytes. Although, a recently developed solution-
cathode/SE interface is addressed by constructing composite assisted solid-state reaction method seems to resolve this
cathodes with nanoarchitecture, adding liquid electrolyte or issue, more work is required to understand the correlations
polymer at the interface, and forming composite electrolyte between the processing parameters, chemistry, and conduc-
of NASICON and polymer. tivity in these oxides.
Even though these materials have been studied over the
last 45 years, there is still ambiguity about the correlations
Challenges and perspectives between the crystal structure and ionic conductivity. Few
studies reported that the rhombohedral phase is more favour-
The NASICON-type oxides generally exhibit good thermal able for the conduction of sodium ions as it possesses higher
stability and fast 3D ionic conduction. Most studies have symmetry than the monoclinic phase [33, 72, 130]. While
focused on enhancing the sodium-ion conduction in these other investigations suggested that the monoclinic phase
oxides, and now the maximum ionic conductivity of ~ 5 has superior ionic conductivity owing to the larger unit-cell
mS·cm−1 at RT is reported in the literature [141, 145]. Nev- size and the lower energy barrier for Na ion conduction,
ertheless, the practical application of NASICON electrolytes which is caused by the high-energy excited Na-sites [85,
is often restricted by the poor interfacial contact between the 138, 141, 142]. Also, the maximum total ionic conductiv-
electrode/SE interface and the short circuit of the cell caused ity in ­Na1+xZr2SixP3-xO12 found in the x range of 1.8 and
by the penetration of dendrites when utilizing pure Na as 2.2, where the monoclinic phase is stable. A deeper insight
an anode. Low interfacial conductivity of less than ­10–2 into the Na conduction mechanism can assist in the rational
mS·cm−1 results in poor rate performance and short cycle design of new and improved NASICON compositions as an
life [25, 90, 99, 130]. Moreover, the poor wettability of Na alternative to the β-alumina.

13
Ionics

Fig. 18  Summary of strategies employed to improve the performance of NASICON-type oxides for sodium-ion batteries

In summary, improving the rate capability and cyclability tailor the properties of these oxides. Further, the ionic trans-
of solid-state sodium-ion batteries based on NASICON elec- port mechanism inside the structure and the effect of adding
trolyte requires high ionic conduction across the electrolyte various dopants and sintering aids on the conductivity and
and a good electrode/SE interface. Figure 18 summarizes microstructure have been comprehensively described. More-
the strategies that have been employed in the past to over- over, this review provides a perspective and future research
come the challenges associated with NASICON solid elec- opportunities on these oxides, which show immense promise
trolytes. Since previous approaches have been to a certain in rechargeable solid-state batteries.
extent successful in augmenting the conductivity in these
oxides, future research should be primarily focused on find- Acknowledgements  The authors gratefully acknowledge the support
from the Department of Science and Technology, Government of India
ing solutions to reduce the interfacial resistance between (grant no. EMR/2016/005438) toward the funding of this research.
the electrode and SE by maintaining close physical contact
during the charging-discharging process.

References
Conclusions
1. Larcher D, Tarascon J-M (2015) Towards greener and more
Na-ion conducting solid electrolytes are critical for realiz- sustainable batteries for electrical energy storage. Nat Chem
7(1):19–29. https://​doi.​org/​10.​1038/​nchem.​2085
ing the development of sodium-ion batteries in all-solid- 2. Barbir F, Veziroǧlu T, Plass H Jr (1990) Environmental damage
state designs. Among various materials, NASICON-type due to fossil fuels use. Int J Hydrogen Energy 15(10):739–749.
oxides have shown tremendous potential as they can pro- https://​doi.​org/​10.​1016/​0360-​3199(90)​90005-J
vide large interconnected channels for a fast migration of 3. Gür TM (2018) Review of electrical energy storage technologies,
materials and systems: challenges and prospects for large-scale
the large sodium ions. A significant number of experimen- grid storage. Energy Environ Sci 11(10):2696–2767. https://​doi.​
tal and theoretical studies have been published in the past, org/​10.​1039/​C8EE0​1419A
focussing on understanding, and improving several aspects 4. Dell RM, Rand DAJ (2001) Energy storage-a key technology
of these oxides. This review discusses the basic character- for global energy sustainability. J Power Sources 100(1–2):2–17.
https://​doi.​org/​10.​1016/​S0378-​7753(01)​00894-1
istics and highlights the various challenges associated with 5. Verma S, Mishra S, Gaur A, Chowdhury S, Mohapatra S,
NASICON-type electrolytes. It also summarizes several Dwivedi G, Verma P (2021) A comprehensive review on energy
recently explored strategies to overcome these issues and to

13
Ionics

storage in hybrid electric vehicle. J Traffic Transp Eng 8(5):621– review. Nano Mater Sci 1(2):91–100. https://​doi.​org/​10.​1016/j.​
637. https://​doi.​org/​10.​1016/j.​jtte.​2021.​09.​001 nanoms.​2019.​02.​007
6. Sanguesa JA, Torres-Sanz V, Garrido P, Martinez FJ, Marquez- 26. Yu Y, Che H, Yang X, Deng Y, Li L, Ma Z-F (2020) Non-flam-
Barja JM (2021) A review on electric vehicles: technologies and mable organic electrolyte for sodium-ion batteries. Electrochem
challenges. Smart Cities 4(1):372–404. https://​doi.​org/​10.​3390/​ Commun 110:106635. https://​doi.​org/​10.​1016/j.​elecom.​2019.​
smart​citie​s4010​022 106635
7. Bruce PG, Freunberger SA, Hardwick LJ, Tarascon J-M (2012) 27. Che H, Chen S, Xie Y, Wang H, Amine K, Liao X-Z, Ma Z-F
Li–O2 and Li–S batteries with high energy storage. Nat Mater (2017) Electrolyte design strategies and research progress for
11(1):19–29. https://​doi.​org/​10.​1038/​nmat3​191 room-temperature sodium-ion batteries. Energy Environ Sci
8. Yang Z, Zhang J, Kintner-Meyer MC, Lu X, Choi D, Lemmon 10(5):1075–1101. https://​doi.​org/​10.​1039/​C7EE0​0524E
JP, Liu J (2011) Electrochemical energy storage for green grid. 28. Kim JJ, Yoon K, Park I, Kang K (2017) Progress in the Devel-
Chem Rev 111(5):3577–3613. https://d​ oi.o​ rg/1​ 0.1​ 021/c​ r1002​ 90v opment of Sodium-Ion Solid Electrolytes. Small Methods
9. Choi JW, Aurbach D (2016)  Promise and reality of post-lithium- 1(10):1700219. https://​doi.​org/​10.​1002/​smtd.​20170​0219
ion batteries with high energy densities. Nat Rev Mater 1:16013. 29. Hou H, Xu Q, Pang Y, Li L, Wang J, Zhang C, Sun C (2017)
https://​doi.​org/​10.​1038/​natre​vmats.​2016.​13 Efficient storing energy harvested by triboelectric nanogenerators
10. Armand M, Tarascon J-M (2008) Building better batteries. using a safe and durable all‐solid‐state sodium‐ion battery. Adv
Nature 451:652–657. https://​doi.​org/​10.​1038/​45165​2a Sci 4(8):1700072. https://​doi.​org/​10.​1002/​advs.​20170​0072
11. Goodenough JB, Park K-S (2013) The Li-ion rechargeable bat- 30. Tian H, Liu S, Deng L, Wang L, Dai L (2021) New-type Hf-
tery: a perspective. J Am Chem Soc 135(4):1167–1176. https://​ based NASICON electrolyte for solid-state Na-ion batteries with
doi.​org/​10.​1021/​ja309​1438 superior long-cycling stability and rate capability. Energy Stor-
12. Sun Y-K, Chen Z, Noh H-J, Lee D-J, Jung H-G, Ren Y, Wang age Mater 39:232–238. https://​doi.​org/​10.​1016/j.​ensm.​2021.​04.​
S, Yoon CS, Myung S-T, Amine K (2012) Nanostructured 026
high-energy cathode materials for advanced lithium batteries. 31. Zhang C, Gamble S, Ainsworth D, Slawin AM, Andreev YG,
Nat Mater 11(11):942–947. https://​doi.​org/​10.​1038/​nmat3​435 Bruce PG (2009) Alkali metal crystalline polymer electrolytes.
13. Sun C, Liu J, Gong Y, Wilkinson DP, Zhang J (2017) Recent Nat Mater 8(7):580–584. https://​doi.​org/​10.​1038/​nmat2​474
advances in all-solid-state rechargeable lithium batteries. Nano 32. Hayashi A, Noi K, Sakuda A, Tatsumisago M (2012) Superi-
Energy 33:363–386. https://​doi.​org/​10.​1016/j.​nanoen.​2017.​01.​ onic glass-ceramic electrolytes for room-temperature recharge-
028 able sodium batteries. Nat Commun 3(1):1–5. https://​doi.​org/​10.​
14. Zhang W, Nie J, Li F, Wang ZL, Sun C (2018) A durable and safe 1038/​ncomm​s1843
solid-state lithium battery with a hybrid electrolyte membrane. 33. Ma Q, Guin M, Naqash S, Tsai C-L, Tietz F, Guillon O (2016)
Nano Energy 45:413–419. https://d​ oi.o​ rg/1​ 0.1​ 016/j.n​ anoen.2​ 018.​ Scandium-substituted ­Na3Zr2(SiO4)2(PO4) prepared by a solu-
01.​028 tion-assisted solid-state reaction method as sodium-ion conduc-
15. Pender JP, Jha G, Youn DH, Ziegler JM, Andoni I, Choi EJ, tors. Chem Mater 28(13):4821–4828. https://​doi.​org/​10.​1021/​
Heller A, Dunn BS, Weiss PS, Penner RM, Mullins CB (2020) acs.​chemm​ater.​6b020​59
Electrode degradation in lithium-ion batteries. ACS Nano 34. Zhang Z, Zhang Q, Shi J, Chu YS, Yu X, Xu K, Ge M, Yan
14(2):1243–1295. https://​doi.​org/​10.​1021/​acsna​no.​9b043​65 H, Li W, Gu L (2017) A Self‐Forming Composite Electrolyte
16. Nitta N, Wu F, Lee JT, Yushin G (2015) Li-ion battery materials: for Solid‐State Sodium Battery with Ultralong Cycle Life. Adv
present and future. Mater Today 18(5):252–264. https://​doi.​org/​ Energy Mater 7(4):1601196. https://d​ oi.o​ rg/1​ 0.1​ 002/a​ enm.2​ 0160​
10.​1016/j.​mattod.​2014.​10.​040 1196
17. Manthiram A (2017) An outlook on lithium-ion battery technol- 35. Duchêne L, Kühnel R-S, Stilp E, Reyes EC, Remhof A, Hage-
ogy. ACS Cent Sci 3(10):1063–1069. https://​doi.​org/​10.​1021/​ mann H, Battaglia C (2017) A stable 3 V all-solid-state sodium–
acsce​ntsci.​7b002​88 ion battery based on a closo-borate electrolyte. Energy Environ
18. U.S. Geological Survey (2020) Mineral commodity summaries Sci 10(12):2609–2615. https://​doi.​org/​10.​1039/​C7EE0​2420G
2020, Reston, VA (2020) 204. https://​doi.​org/​10.​3133/​mcs20​20 36. Noguchi Y, Kobayashi E, Plashnitsa LS, Okada S, Yamaki J-I
19. Hartmann P, Bender CL, Vračar M, Dürr AK, Garsuch A, Janek (2013) Fabrication and performances of all solid-state symmetric
J, Adelhelm P (2013) A rechargeable room-temperature sodium sodium battery based on NASICON-related compounds. Electro-
superoxide ­(NaO2) battery. Nat Mater 12(3):228–232. https://d​ oi.​ chim Acta 101:59–65. https://​doi.​org/​10.​1016/j.​elect ​acta.​2012.​
org/​10.​1038/​nmat3​486 11.​038
20. Yabuuchi N, Kubota K, Dahbi M, Komaba S (2014) Research 37. Wei T, Gong Y, Zhao X, Huang K (2014) An All‐Ceramic Solid‐
development on sodium-ion batteries. Chem Rev 114(23):11636– State Rechargeable ­Na+‐Battery Operated at Intermediate Tem-
11682. https://​doi.​org/​10.​1021/​cr500​192f peratures. Adv Funct Mater 24(34):5380–5384. https://​doi.​org/​
21. Xiang X, Zhang K, Chen J (2015) Recent advances and pros- 10.​1002/​adfm.​20140​0773
pects of cathode materials for sodium‐ion batteries. Adv Mater 38. Liu C, Massé R, Nan X, Cao G (2016) A promising cathode for
27(36):5343–5364. https://​doi.​org/​10.​1002/​adma.​20150​1527 Li-ion batteries: L ­ i3V2(PO4)3. Energy Storage Mater 4:15–58.
22. Luo W, Shen F, Bommier C, Zhu H, Ji X, Hu L (2016) Na-ion https://​doi.​org/​10.​1016/j.​ensm.​2016.​02.​002
battery anodes: materials and electrochemistry. Acc Chem Res 39. Ramesh A, Tripathi A, Balaya P (2022) Int. J. Appl.
49(2):231–240. https://​doi.​org/​10.​1021/​acs.​accou​nts.​5b004​82 Ceram. Technol. 19(2): 913
23. Su H, Jaffer S, Yu H (2016) Transition metal oxides for sodium- 40. Ni Q, Bai Y, Wu F, Wu C (2017) Adv. Sci. 4(3):1600275
ion batteries. Energy Storage Mater 5:116–131. https://​doi.​org/​ 41. Zheng F, Kotobuki M, Song S, Lai MO, Lu L (2018) Review on
10.​1016/j.​ensm.​2016.​06.​005 solid electrolytes for all-solid-state lithium-ion batteries. J Power
24. Hwang J-Y, Myung S-T, Sun Y-K (2017) Sodium-ion batteries: Sources 389:198–213. https://​doi.​org/​10.​1016/j.​jpows​our.​2018.​
present and future. Chem Soc Rev 46(12):3529–3614. https://​ 04.​022
doi.​org/​10.​1039/​C6CS0​0776G 42. Ngai KS, Ramesh S, Ramesh K, Juan JC (2016) A review of
25. Wang Y, Song S, Xu C, Hu N, Molenda J, Lu L (2019) Develop- polymer electrolytes: fundamental, approaches and appli-
ment of solid-state electrolytes for sodium-ion battery–A short cations. Ionics 22(8):1259–12779. https://​d oi.​o rg/​1 0.​1 007/​
s11581-​016-​1756-4

13
Ionics

43. Golodnitsky D, Strauss E, Peled E, Greenbaum S (2015) On 61. Bui KM, Dinh VA, Okada S, Ohno T (2016) Na-ion diffusion in
order and disorder in polymer electrolytes. J Electrochem Soc a NASICON-type solid electrolyte: a density functional study.
162(14):A2551–A2566. https://​doi.​org/​10.​1149/2.​01615​14jes Phys Chem Chem Phys 18(39):27226–27231. https://​doi.​org/​10.​
44. Fergus JW (2012) Ion transport in sodium ion conducting solid 1039/​C6CP0​5164B
electrolytes. Solid State Ionics 227:102–112. https://​doi.​org/​10.​ 62. Aono H, Sugimoto E, Sadaoka Y, Imanaka N, Adachi GY
1016/j.​ssi.​2012.​09.​019 (1989) Ionic Conductivity of the Lithium Titanium Phosphate
45. Zhao C, Liu L, Qi X, Lu Y, Wu F, Zhao J, Yu Y, Hu YS, Chen ­(Li1 + X M X ­Ti2 − X (PO 4) 3, M = Al, Sc, Y, and La) Systems. J Elec-
L (2018) Solid‐state sodium batteries. Adv Energy Mater trochem Soc 136(2):590. https://​doi.​org/​10.​1149/1.​20966​93
8(17):1703012. https://​doi.​org/​10.​1002/​aenm.​20170​3012 63. Khireddine H, Fabry P, Caneiro A, Bochu B (1997) Optimization
46. Lu Y, Li L, Zhang Q, Niu Z, Chen J (2018) Electrolyte and of NASICON composition for ­Na+ recognition. Sens Actuators B
interface engineering for solid-state sodium batteries. Joule 40(2):223–230. https://d​ oi.o​ rg/1​ 0.1​ 016/S
​ 0925-4​ 005(97)8​ 0266-3
2(9):1747–1770. https://​doi.​org/​10.​1016/j.​joule.​2018.​07.​028 64. Hayashi A, Masuzawa N, Yubuchi S, Tsuji F, Hotehama C,
47. Hueso KB, Armand M, Rojo T (2013) High temperature sodium Sakuda A, Tatsumisago M (2019) A sodium-ion sulfide solid
batteries: status, challenges and future trends. Energy Environ electrolyte with unprecedented conductivity at room tem-
Sci 6(3):734–749. https://​doi.​org/​10.​1039/​C3EE2​4086J perature. Nat Commun 10(1):5266. https://​doi.​org/​10.​1038/​
48. Rajagopalan R, Chen B, Zhang Z, Wu XL, Du Y, Huang s41467-​019-​13178-2
Y, Li B, Zong Y, Wang J, Nam GH (2017) Improved revers- 65. Ouyang B, Wang J, He T, Bartel CJ, Huo H, Wang Y, Lacivita
ibility of ­Fe3+/Fe4+ redox couple in sodium super ion conduc- V, Kim H, Ceder G (2021) Synthetic accessibility and stability
tor type N ­ a3Fe2(PO4)3 for sodium‐ion batteries. Adv Mater rules of NASICONs. Nat Commun 12(1):5752. https://​doi.​org/​
29(12):1605694. https://​doi.​org/​10.​1002/​adma.​20160​5694 10.​1038/​s41467-​021-​26006-3
49. Fu L, Xue X, Tang Y, Sun D, Xie H, Wang H (2018) Size control- 66. Jolley AG, Taylor DD, Schreiber NJ, Wachsman ED (2015)
ling and surface engineering enable N ­ aTi2(PO4)3/C outstanding Structural Investigation of Monoclinic‐Rhombohedral Phase
sodium storage properties. Electrochim Acta 289:21–28. https://​ Transition in ­Na3Zr2Si2PO12 and Doped NASICON. J Am Ceram
doi.​org/​10.​1016/j.​elect​acta.​2018.​09.​024 Soc 98(9):2902–2907. https://​doi.​org/​10.​1111/​jace.​13692
50. Leng H, Huang J, Nie J, Luo J (2018) Cold sintering and ionic 67. Deng Z, SaiGautam G, Kolli SK, Chotard J-N, Cheetham AK,
conductivities of N ­ a3.256Mg0.128Zr1.872Si2PO12 solid electrolytes. Masquelier C, Canepa P (2020) Phase Behavior in Rhombo-
J Power Sources 391:170–179. https://​doi.​org/​10.​1016/j.​jpows​ hedral NaSiCON Electrolytes and Electrodes. Chem Mater
our.​2018.​04.​067 32(18):7908–7920. https://d​ oi.o​ rg/1​ 0.1​ 021/a​ cs.c​ hemma​ ter.0​ c026​
51. Anantharamulu N, Rao KK, Rambabu G, Kumar BV, Radha V, 95
Vithal M (2011) A wide-ranging review on Nasicon type mate- 68. Roy S, Kumar PP (2013) Influence of Si/P ordering on N ­ a+ trans-
rials. J Mater Sci 46(9):2821–2837. https://​doi.​org/​10.​1007/​ port in NASICONs. Phys Chem Chem Phys 15(14):4965–4969.
s10853-​011-​5302-5 https://​doi.​org/​10.​1039/​C3CP4​3376E
52. Kim J-K, Lim YJ, Kim H, Cho G-B, Kim Y (2015) A hybrid 69. Guin M, Tietz F (2015) Survey of the transport properties of
solid electrolyte for flexible solid-state sodium batteries. Energy sodium superionic conductor materials for use in sodium batter-
Environ Sci 8(12):3589–3596. https://​doi.​org/​10.​1039/​C5EE0​ ies. J Power Sources 273:1056–1064. https://​doi.​org/​10.​1016/j.​
1941A jpows​our.​2014.​09.​137
53. Zhang Z, Zhang Q, Ren C, Luo F, Ma Q, Hu Y-S, Zhou Z, Li H, 70. Oh JAS, He L, Plewa A, Morita M, Zhao Y, Sakamoto T,
Huang X, Chen L (2016) A ceramic/polymer composite solid Song X, Zhai W, Zeng K, Lu L (2019) Composite NASICON
electrolyte for sodium batteries. J Mater Chem A 4(41):15823– ­(Na3Zr2Si2PO12) solid-state electrolyte with enhanced ­Na+ ionic
15828. https://​doi.​org/​10.​1039/​C6TA0​7590H conductivity: effect of liquid phase sintering. ACS Appl Mater
54. Jolley AG, Cohn G, Hitz GT, Wachsman ED (2015) Improving Interfaces 11(43):40125–40133. https://​doi.​org/​10.​1021/​acsami.​
the ionic conductivity of NASICON through aliovalent cation 9b149​86
substitution of N­ a3Zr2Si2PO12. Ionics 21(11):3031–3038. https://​ 71. Boilot JP, Collin G, Colomban P (1987) Crystal structure of the
doi.​org/​10.​1007/​s11581-​015-​1498-8 true nasicon: N ­ a3Zr2Si2PO12. Mater Res Bull 22(5):669–676.
55. Adelhelm P, Hartmann P, Bender CL, Busche M, Eufinger C, https://​doi.​org/​10.​1016/​0025-​5408(87)​90116-4
Janek J (2015) From lithium to sodium: cell chemistry of room 72. Ran L, Baktash A, Li M, Yin Y, Demir B, Lin T, Li M, Rana M,
temperature sodium–air and sodium–sulfur batteries. Beilstein J Gentle I, Wang L, Searles DJ, Knibbe R (2021) Sc, Ge co-doping
Nanotechnol 6:1016–1055. https://d​ oi.o​ rg/1​ 0.3​ 762/b​ jnano.6​ .1​ 05 NASICON boosts solid-state sodium ion batteries’ performance.
56. Zhou W, Li Y, Xin S, Goodenough JB (2017) Rechargeable Energy Storage Mater 40:282–291. https://​doi.​org/​10.​1016/j.​
sodium all-solid-state battery. ACS Cent Sci 3(1):52–57. https://​ ensm.​2021.​05.​017
doi.​org/​10.​1021/​acsce​ntsci.​6b003​21 73. Naqash S (2019) Sodium ion conducting ceramics for sodium ion
57. Rajagopalan R, Zhang Z, Tang Y, Jia C, Ji X, Wang (2021) batteries, Forschungszentrum Jülich GmbH, Institut für Energie-
Energy Storage Mater. 34:171 und Klimaforschung. Forschungszentrum Jülich GmbH, Zentral-
58. Rao YB, Bharathi KK, Patro L (2021) Review on the synthesis bibliothek, Verlag, Germany
and doping strategies in enhancing the Na ion conductivity of 74. Deng Y, Eames C, Nguyen LHB, Pecher O, Griffith KJ, Courty
­Na3Zr2Si2PO12 (NASICON) based solid electrolytes. Solid State M, Fleutot B, Chotard J-N, Grey CP, Islam MS, Masquelier C
Ionics 366:115671 (2018) Crystal structures, local atomic environments, and ion
59. Yang Z, Tang B, Xie Z, Zhou Z (2021) NASICON-Type Na3Zr- diffusion mechanisms of scandium-substituted sodium supe-
2Si2PO12 Solid-State Electrolytes for Sodium Batteries. ChemE- rionic conductor (NASICON) solid electrolytes. Chem Mater
lectroChem 8(6):1035–1047. https://​doi.​org/​10.​1002/c​ elc.​20200​ 30(8):2618–2630. https://​doi.​org/​10.​1021/​acs.​chemm​ater.​7b052​
1527 37
60. Hong H-P (1976) Crystal structures and crystal chemistry in 75. Mazza D (2001) Modeling ionic conductivity in Nasicon struc-
the system N ­ a1+xZr2SixP3−xO12. Mater Res Bull 11(2):173–182. tures. J Solid State Chem 156(1):154–160. https://​doi.​org/​10.​
https://​doi.​org/​10.​1016/​0025-​5408(76)​90073-8 1006/​jssc.​2000.​8975

13
Ionics

76. Baur WH, Dygas JR, Whitmore DH, Faber J (1986) Neutron 93. Suzuki K, Noi K, Hayashi A, Tatsumisago M (2018) Low tem-
powder diffraction study and ionic conductivity of N ­ a2Zr2SiP2O12 perature sintering of ­Na1+xZr2SixP3−xO12 by the addition of
and ­Na3Zr2Si2PO12. Solid State Ionics 18–19:935–943. https://​ ­Na3BO3. Scr Mater 145:67–70. https://​doi.​org/​10.​1016/j.​scrip​
doi.​org/​10.​1016/​0167-​2738(86)​90290-0 tamat.​2017.​10.​010
77. Zhang Z, Zou Z, Kaup K, Xiao R, Shi S, Avdeev M, Hu Y-S, 94. Leng H, Nie J, Luo J (2019) Combining cold sintering and
Wang D, He B, Li H, Huang X, Nazar LF, Chen L (2019) Cor- ­Bi2O3-Activated liquid-phase sintering to fabricate high-conduc-
related migration invokes higher N ­ a+-Ion conductivity in NaSI- tivity Mg-doped NASICON at reduced temperatures. J Materi-
CON-type solid electrolytes. Adv Energy Mater 9(42):1902373. omics 5(2):237–246. https://​doi.​org/​10.​1016/j.​jmat.​2019.​02.​005
https://​doi.​org/​10.​1002/​aenm.​20190​2373 95. Cao XG, Zhang XH, Tao T, Zhang HY (2020) Effects of
78. Kumar PP, Yashonath S (2006) Ionic conduction in the solid antimony tin oxide (ATO) additive on the properties of
state. J Chem Sci 118(1):135–154. https://​doi.​org/​10.​1007/​ ­Na3Zr2Si2PO12 ceramic electrolytes. Ceram Int 46(6):8405–8412.
BF027​08775 https://​doi.​org/​10.​1016/j.​ceram​int.​2019.​12.​074
79. Losilla ER, Aranda MAG, Bruque S, París MA, Sanz J, West 96. Chen D, Luo F, Gao L, Zhou W, Zhu D (2017) Influence of
AR (1998) Understanding Na mobility in NASICON materials: a Indium-Tin Oxide Additive on the Sintering Process and Con-
rietveld, 23Na and 31P MAS NMR, and impedance study. Chem ductivity of N ­ a3Zr2Si2PO12 Solid Electrolyte. J Electron Mater
Mater 10(2):665–673. https://​doi.​org/​10.​1021/​cm970​648j 46(11):6367–6372. https://​doi.​org/​10.​1007/​s11664-​017-​5674-7
80. Huang C, Yang G, Yu W, Xue C, Zhai Y, Tang W, Hu N, Wen Z, 97. Wang X, Liu Z, Tang Y, Chen J, Wang D, Mao Z (2021) Low
Lu L, Song S (2021) Gallium-substituted Nasicon N ­ a3Zr2Si2PO12 temperature and rapid microwave sintering of N ­ a3Zr2Si2PO12
solid electrolytes. J Alloys Compd 855(2):157501 solid electrolytes for Na-Ion batteries. J Power Sources
81. Park H, Jung K, Nezafati M, Kim C-S, Kang B, Appl ACS (2016) 481:228924. https://​doi.​org/​10.​1016/j.​jpows​our.​2020.​228924
Sodium ion diffusion in Nasicon ­(Na3Zr2Si2PO12) solid electro- 98. Thirupathi R, Omar S (2021) J Phys Chem C 125(50):27723
lytes: effects of excess sodium. Mater Interfaces 8(41):27814– 99. Narayanan S, Reid S, Butler S, Thangadurai V (2019) Sinter-
27824. https://​doi.​org/​10.​1021/​acsami.​6b099​92 ing temperature, excess sodium, and phosphorous depend-
82. Saito Y, Ado K, Asai T, Kageyama H, Nakamura O encies on morphology and ionic conductivity of NASICON
(1992) Ionic conductivity of NASICON-type conductors ­Na3Zr2Si2PO12. Solid State Ionics 331:22–29. https://​doi.​org/​
­Na1. 5M0. 5Zr1. 5(PO4)3 (M: ­Al3+, ­Ga3+, ­Cr3+, ­Sc3+, ­Fe3+, ­In3+, 10.​1016/j.​ssi.​2018.​12.​003
­Yb3+, ­Y3+). Solid State Ionics 58(3–4):327–331. https://​doi.​org/​ 100. Essoumhi A, Favotto C, Mansori M, Satre P (2004) Synthesis
10.​1016/​0167-​2738(92)​90136-D and characterization of a NASICON series with general formula
83. Roy S, Kumar PP (2013) Influence of Cationic ordering on ion ­Na2. 8Zr2− ySi1. 8− 4yP1. 2+ 4yO12 (0⩽ y⩽ 0.45). J Solid State Chem
transport in NASICONs: Molecular dynamics study. Solid State 177(12):4475–4481. https://​doi.​org/​10.​1016/j.​jssc.​2004.​09.​026
Ionics 253:217–222. https://​doi.​org/​10.​1016/j.​ssi.​2013.​09.​030 101. Martucci A, Sartori S, Guglielmi M, Di Vona ML, Licoccia S,
84. Fuentes R, Figueiredo F, Marques F, Franco J (2001) Influence Traversa E (2002) NMR and XRD study of the influence of the
of microstructure on the electrical properties of NASICON mate- P precursor in sol-gel synthesis of NASICON powders and films.
rials. Solid State Ionics 140(1–2):173–179. https://​doi.​org/​10.​ J Eur Ceram Soc 22(12):1995–2000. https://​doi.​org/​10.​1016/​
1016/​S0167-​2738(01)​00701-9 S0955-​2219(01)​00536-2
85. Samiee M, Radhakrishnan B, Rice Z, Deng Z, Meng YS, Ong SP, 102. Clearfield A, Subramanian MA, Wang W, Jerus P (1983) The use
Luo J (2017) Divalent-doped ­Na3Zr2Si2PO12 natrium superionic of hydrothermal procedures to synthesize NASICON and some
conductor: Improving the ionic conductivity via simultaneously comments on the stoichiometry of NASICON phases. Solid State
optimizing the phase and chemistry of the primary and secondary Ionics 9–10(2):895–902. https://​doi.​org/​10.​1016/​0167-​2738(83)​
phases. J Power Sources 347:229–237. https://​doi.​org/​10.​1016/j.​ 90108-X
jpows​our.​2017.​02.​042 103. Clearfield A, Jerus P, Cotman RN (1981) Hydrothermal and
86. Bogusz W, Krok F, Jakubowski W (1981) Bulk and grain bound- solid-state synthesis of sodium zirconium silicophosphates. Solid
ary electrical conductivities of NASICON. Solid State Ionics State Ionics 5:301–304. https://​doi.​org/​10.​1016/​0167-​2738(81)​
2(3):171–174. https://​doi.​org/​10.​1016/​0167-​2738(81)​90175-2 90252-6
87. Bohnke O, Ronchetti S, Mazza D (1999) Conductivity meas- 104. Ignaszak A, Pasierb P, Gajerski R, Komornicki S (2005) Synthe-
urements on nasicon and nasicon-modified materials. Solid sis and properties of Nasicon-type materials. Thermochim Acta
State Ionics 122(1–4):127–136. https://​doi.​org/​10.​1016/​S0167-​ 426(1):7–14. https://​doi.​org/​10.​1016/j.​tca.​2004.​07.​002
2738(99)​00062-4 105. Fuentes R, Figueiredo F, Soares M, Marques F (2005) Submicro-
88. Lee JS, Chang CM, Lee YI, Lee JH, Hong SH (2004) Spark metric NASICON ceramics with improved electrical conductivity
plasma sintering (SPS) of NASICON ceramics. J Am Ceram Soc obtained from mechanically activated precursors. J Eur Ceram Soc
87(2):305–307. https://d​ oi.o​ rg/1​ 0.1​ 111/j.1​ 551-2​ 916.2​ 004.0​ 0305 25(4):455–462. https://​doi.​org/​10.​1016/j.​jeurc​erams​oc.​2004.​02.​019
89. Yoldas BE, Lloyd I (1983) Nasicon formation by chemical 106. Dhas NA, Patil KC (1994) Controlled combustion synthesis and
polymerization. Mater Res Bull 18(10):1171–1177. https://​doi.​ properties of fine-particle NASICON materials. J Mater Chem
org/​10.​1016/​0025-​5408(83)​90019-3 4(3):491–497. https://​doi.​org/​10.​1039/​JM994​04004​91
90. Lu Y, Alonso JA, Yi Q, Lu L, Wang ZL, Sun C (2019) A 107. McEntire BJ, Bartlett R, Miller G, Gordon R (1983) Effect of
high‐performance monolithic solid‐state sodium battery with decomposition on the densification and properties of nasicon
­C a 2+ doped N ­ a 3Zr 2Si 2PO 12 electrolyte. Adv Energy Mater ceramic electrolytes. J Am Ceram Soc 66(10):738–742. https://​
9(28):1901205. https://​doi.​org/​10.​1002/​aenm.​20190​1205 doi.​org/​10.​1111/j.​1151-​2916.​1983.​tb105​41.x
91. Nishio A, Shirai N, Minami H, Izumi H, Inoishi A, Okada S 108. Bell NS, Edney C, Wheeler JS, Ingersoll D, Spoerke ED (2014)
(2021) Effect of N ­ a3BO3 Addition into N ­ a3V2(PO4)3 Single- The Influences of Excess Sodium on Low‐Temperature NaSI-
Phase All-Solid-State Batteries. Electrochemistry 21:244–249. CON Synthesis. J Am Ceram Soc 97(12):3744–3748. https://​
https://​doi.​org/​10.​5796/​elect​roche​mistry.​21-​00023 doi.​org/​10.​1111/​jace.​13167
92. Noi K, Suzuki K, Tanibata N, Hayashi A, Tatsumisago M 109. Gasmi N, Gharbi N, Zarrouk H, Barboux P, Morineau R, Livage
(2018) Liquid‐phase sintering of highly N ­ a+ ion conducting J (1995) Comparison of different synthesis methods for Nasicon
­Na3Zr2Si2PO12 ceramics using N ­ a3BO3 additive. J Am Ceram ceramics. J Sol-Gel Sci Technol 4(3):231–237. https://​doi.​org/​
Soc 101(3):1255–1265. https://​doi.​org/​10.​1111/​jace.​15288 10.​1007/​BF004​88378

13
Ionics

110. Ahmad A, Wheat T, Kuriakose A, Canaday J, McDonald A gels: ionic conductivity, durability in molten sodium and strength
(1987) Dependence of the properties of Nasicons on their com- test data. Solid State Ionics 14(1):73–79. https://d​ oi.o​ rg/1​ 0.1​ 016/​
position and processing. Solid State Ionics 24(1):89–97. https://​ 0167-​2738(84)​90014-6
doi.​org/​10.​1016/​0167-​2738(87)​90070-1 128. Zhang Q, Liang F, Qu T, Yao Y, Ma W, Yang B, Dai Y (2018)
111. Warhus U, Maier J, Rabenau A (1988) Thermodynamics of Effect on ionic conductivity of N ­ a3+ xZr2-xMxSi2PO12 (M= Y, La)
NASICON ­(Na1+xZr2SixP3−xO12). J Solid State Chem 72(1):113– by doping rare-earth elements (2018) IOP Conference Series:
125. https://​doi.​org/​10.​1016/​0022-​4596(88)​90014-X Materials Science and Engineering, IOP Publishing, 423:012122
112. Zhang S, Quan B, Zhao Z, Zhao B, He Y, Chen W (2004) Prepa- 129. Khakpour Z (2016) Influence of M: ­Ce4+, ­Gd3+ and ­Yb3+ sub-
ration and characterization of NASICON with a new sol–gel pro- stituted ­Na3+xZr2-xMxSi2PO12 solid NASICON electrolytes on
cess. Mater Lett 58(1):226–229. https://​doi.​org/​10.​1016/​S0167-​ sintering, microstructure and conductivity. Electrochim Acta
577X(03)​00450-6 196:337–347. https://​doi.​org/​10.​1016/​2Fj.​elect​acta.​2016.​02.​199
113. Yadav P, Bhatnagar M (2012) Structural studies of NASICON 130. Shao Y, Zhong G, Lu Y, Liu L, Zhao C, Zhang Q, Hu Y-S, Yang
material of different compositions by sol–gel method. Ceram Y, Chen L (2019) A novel NASICON-based glass-ceramic com-
Int 38(2):1731–1735. https://​doi.​org/​10.​1016/j.​ceram​int.​2011.​ posite electrolyte with enhanced Na-ion conductivity. Energy
09.​022 Storage Mater 23:514–521. https://d​ oi.o​ rg/1​ 0.1​ 016/j.e​ nsm.2​ 019.​
114. Traversa E, Aono H, Sadaoka Y, Montanaro L (2000) Electrical 04.​009
properties of sol–gel processed NASICON having new compo- 131. Goodenough JB, Hong H-P, Kafalas J (1976) Fast N ­ a+-ion
sitions. Sens Actuators B 65(1–3):204–208. https://​doi.​org/​10.​ transport in skeleton structures. Mater Res Bull 11(2):203–220.
1016/​S0925-​4005(99)​00293-2 https://​doi.​org/​10.​1016/​0025-​5408(76)​90077-5
115. Naqash S, Ma Q, Tietz F, Guillon O (2017) N ­ a3Zr2(SiO4)2(PO4) 132. Banik A, Famprikis T, Ghidiu M, Ohno S, Kraft MA, Zeier WG
prepared by a solution-assisted solid-state reaction. Solid State (2021) On the underestimated influence of synthetic conditions
Ionics 302:83–91. https://​doi.​org/​10.​1016/j.​ssi.​2016.​11.​004 in solid ionic conductors. Chem Sci 12(18):6238–6263. https://​
116. Okubo K, Wang H, Hayashi K, Inada M, Enomoto N, Hasegawa doi.​org/​10.​1039/​D0SC0​6553F
G, Osawa T, Takamura H (2018) A dense NASICON sheet pre- 133. Yao Z, Zhu K, Zhu K, Zhang J, Li X, Chen J, Wang J, Yan K, Liu
pared by tape-casting and low temperature sintering. Electrochim J (2021) Co-Precipitation Synthesis and Electrochemical Proper-
Acta 278:176–181. https://​doi.​org/​10.​1016/j.​elect​acta.​2018.​05.​ ties of NASICON-Type L ­ i1. 3Al0.3Ti1.7(PO4)3 Solid Electrolytes.
020 J Mater Sci: Mater Electron 32:24834–24844. https://​doi.​org/​10.​
117. Grady ZM, Tsuji K, Ndayishimiye A, Hwan-Seo J, Randall 1007/​s10854-​021-​06943-x
CA (2020) Densification of a Solid-State NASICON Sodium- 134. Dai H, Xu W, Hu Z, Chen Y, Wei X, Yang B, Chen Z, Gu J,
Ion Electrolyte Below 400° C by Cold Sintering with a Fused Yang D, Xie F (2020) Effective approaches of improving the
Hydroxide Solvent. Appl ACS Energy Mater 3(5):4356–4366. performance of chalcogenide solid electrolytes for all-solid-state
https://​doi.​org/​10.​1021/​acsaem.​0c000​47 sodium-ion batteries. Front Energy Res 8:97. https://​doi.​org/​10.​
118. Grasso S, Biesuz M, Zoli L, Taveri G, Duff AI, Ke D, Jiang A, 3389/​fenrg.​2020.​00097
Reece MJ (2020) A review of cold sintering processes. Adv Appl 135. Xie B, Jiang D, Wu J, Feng T, Xia J, Nian H (2016) Effect of
Ceram 119(3):115–143. https://d​ oi.o​ rg/1​ 0.1​ 080/1​ 74367​ 53.2​ 019.​ substituting Ce for Zr on the electrical properties of NASICON
17068​25 materials. J Phys Chem Solids 88:104–108. https://​doi.​org/​10.​
119. da Silva JGP, Bram M, Laptev AM, Gonzalez-Julian J, Ma Q, 1016/j.​jpcs.​2015.​10.​003
Tietz F, Guillon O (2019) Sintering of a sodium based NASI- 136. Yang J, Wan H-L, Zhang Z-H, Liu G-Z, Xu X-X, Hu Y-S, Yao
CON electrolyte: a comparative study between cold, field X-Y (2018) NASICON-structured N ­ a3.1Zr1.95Mg0.05Si2PO12 solid
assisted and conventional sintering methods. J Eur Ceram Soc electrolyte for solid-state sodium batteries. Rare Met 37(6):480–
39(8):2697–2702. https://​doi.​org/​10.​1016/j.​jeurc​erams​oc.​2019.​ 487. https://​doi.​org/​10.​1007/​s12598-​018-​1020-3
03.​023 137. Song S, Duong HM, Korsunsky AM, Hu N, Lu L (2016) A N ­ a+
120. Kuriakose A, Wheat T, Ahmad A, Dirocco J (1984) Synthesis, superionic conductor for room-temperature sodium batteries. Sci
sintering, and microstructure of NASICONS. J Am Ceram Soc Rep 6(1):1–10. https://​doi.​org/​10.​1038/​srep3​2330
67(3):179–183. https://​doi.​org/​10.​1111/j.​1151-​2916.​1984.​tb197​ 138. Shen L, Yang J, Liu G, Avdeev M, Yao X (2021) High ionic con-
37.x ductivity and dendrite-resistant NASICON solid electrolyte for
121. Kwon OJ, Yoon DN (1980) The liquid phase sintering of W-Ni. all-solid-state sodium batteries. Mater Today Energy 20:100691
Sintering processes, Materials Science Research. Plenum Press, 139. Santhoshkumar B, Rao PL, Ramanathan K, Bera A, Yusuf S,
New York Hathwar VR, Pahari B (2021) Structure and ionic conductivity
122. Lee S-M, Lee S-T, Lee D-H, Lee S-H, Han S-S, Lim S-K (2015) of ­Na3+xSc2SixP3-xO12 (x= 0.0, 0.2, 0.4, 0.8) NASICON materi-
Effect of particle size on the density and ionic conductivity of als: A combined neutron diffraction, MAS NMR and impedance
­Na3Zr2Si2PO12 NASICON. J Ceram Process Res 16(1):49–53. study. Solid State Sci 111:106470
https://​doi.​org/​10.​36410/​jcpr.​2015.​16.1.​49 140. Chen D, Luo F, Zhou W, Zhu D (2018) Influence of ­Nb5+, ­Ti4+,
123. Choi S-D, Park J-W (1996) Preparation of NASICON Powder ­Y3+ and Z ­ n2+ doped N ­ a3Zr2Si2PO12 solid electrolyte on its
and Electrolyte. Sens Mater 8(8):505–511 conductivity. J Alloys Compd 757:348–355. https://​doi.​org/​10.​
124. Ceramatec, Inc. (1980) J Power Sources 5 (4):413 1016/j.​jallc​om.​2018.​05.​116
125. Fuentes R, Figueiredo F, Marques F, Franco J (2001) Process- 141. Yang J, Liu G, Avdeev M, Wan H, Han F, Shen L, Zou Z, Shi
ing and electrical properties of NASICON prepared from yttria- S, Hu Y-S, Wang C (2020) Ultrastable all-solid-state sodium
doped zirconia precursors. J Eur Ceram Soc 21(6):737–743. rechargeable batteries. ACS Energy Lett 5(9):2835–2841. https://​
https://​doi.​org/​10.​1016/​S0955-​2219(00)​00264-8 doi.​org/​10.​1021/​acsen​ergyl​ett.​0c014​32
126. Von Alpen U, Bell MF, Höfer HH (1981) Compositional depend- 142. Sun F, Xiang Y, Sun Q, Zhong G, Banis MN, Li W, Liu Y, Luo J,
ence of the electrochemical and structural parameters in the Nasi- Li R, Fu R (2021) Insight into Ion Diffusion Dynamics/Mecha-
con system (­ Na1+xSixZr2P3−xO12). Solid State Ionics 3–4:215– nisms and Electronic Structure of Highly Conductive Sodium-
218. https://​doi.​org/​10.​1016/​0167-​2738(81)​90085-0 Rich ­Na3+xLaxZr2–xSi2PO12 (0≤ x≤ 0.5) Solid-State Electrolytes.
127. Yde-Andersen S, Lundsgaard JS, Møller L, Engell J (1984) Prop- ACS Appl Mater Interfaces 13(11):13132–13138. https://d​ oi.o​ rg/​
erties of NASICON electrolytes prepared from alkoxide derived 10.​1021/​acsami.​0c218​82

13
Ionics

143. Ruan Y, Song S, Liu J, Liu P, Cheng B, Song X, Battaglia V 158. Pet’kov V, Asabina E, Shchelokov I (2013) Thermal expansion
(2017) Improved structural stability and ionic conductivity of of NASICON materials. Inorg Mater 49(5):502–506. https://​doi.​
­Na3Zr2Si2PO12 solid electrolyte by rare earth metal substitutions. org/​10.​1134/​S0020​16851​30501​17
Ceram Int 43(10):7810–7815. https://d​ oi.o​ rg/1​ 0.1​ 016/j.c​ erami​ nt.​ 159. Guin M, Tietz F, Guillon O (2016) New promising NASICON
2017.​03.​095 material as solid electrolyte for sodium-ion batteries: Correlation
144. Hayashi K, Shima K, Sugiyama F (2013) A mixed aqueous/ between composition, crystal structure and ionic conductivity of
aprotic sodium/air cell using a NASICON ceramic separator. ­Na3+ xSc2SixP3− xO12. Solid State Ionics 293:18–26. https://​doi.​
J Electrochem Soc 160(9):A1467–A1472. https://​doi.​org/​10.​ org/​10.​1016/j.​ssi.​2016.​06.​005
1149/2.​06730​9jes 160. Cleveland J, Bradt R (1978) Grain size/microcracking relations
145. Liu Y, Liu L, Peng J, Zhou X, Liang D, Zhao L, Su J, Zhang for pseudobrookite oxides. J Am Ceram Soc 61(11–12):478–481.
B, Li S, Zhang N, Ma Q, Tietz F (2022) A niobium-substituted https://​doi.​org/​10.​1111/j.​1151-​2916.​1978.​tb161​21.x
sodium superionic conductor with conductivity higher than 161. Jackman SD, Cutler RA (2012) Effect of microcracking on ionic
5.5 mS ­cm−1 prepared by solution-assisted solid-state reaction conductivity in LATP. J Power Sources 218:65–72. https://​doi.​
method. J Power Sources 518:230765. https://​doi.​org/​10.​1016/j.​ org/​10.​1016/j.​jpows​our.​2012.​06.​081
jpows​our.​2021.​230765 162. Bradt RC (1995) Low expansion materials. Ceram Trans 52:5–18
146. Ma Q, Tsai C-L, Wei X-K, Heggen M, Tietz F, Irvine JT (2019) 163. Min K (2021) High-throughput Ab initio investigation of the
Room temperature demonstration of a sodium superionic con- elastic properties of inorganic electrolytes for all-solid-state Na-
ductor with grain conductivity in excess of 0.01 S ­cm−1 and its ion batteries. J Electrochem Soc 168(3):030541. https://​doi.​org/​
primary applications in symmetric battery cells. J Mater Chem 10.​1149/​1945-​7111/​abf015
A 7(13):7766–7776. https://​doi.​org/​10.​1039/​C9TA0​0048H 164. Monroe C, Newman J (2005) The impact of elastic deformation
147. Zhou M, Ahmad A (2007) Synthesis, processing and characteri- on deposition kinetics at lithium/polymer interfaces. J Electro-
zation of nasicon solid electrolytes for C ­ O2 sensing applications. chem Soc 152(2):A396–A404. https://​doi.​org/​10.​1149/1.​18508​
Sens Actuators B 122(2):419–426. https://​doi.​org/​10.​1016/j.​snb.​ 54
2006.​06.​011 165. Tang B, Jaschin PW, Li X, Bo S-H, Zhou Z (2020) Mater Today
148. Lalère F, Leriche J-B, Courty M, Boulineau S, Viallet V, 41:200
Masquelier C, Seznec V (2014) An all-solid state NASICON 166. Nonemacher JF, Naqash S, Tietz F, Malzbender J (2019) Micro-
sodium battery operating at 200 °C. J Power Sources 247:975– mechanical assessment of Al/Y-substituted NASICON solid
980. https://​doi.​org/​10.​1016/j.​jpows​our.​2013.​09.​051 electrolytes. Ceram Int 45(17):21308–21314. https://​doi.​org/​10.​
149. Porkodi P, Yegnaraman V, Kamaraj P, Kalyanavalli V, Jeyakumar 1016/j.​ceram​int.​2019.​07.​114
D (2008) Synthesis of NASICON- A molecular precursor-based 167. Kalnaus S, Amin R, Parish C, Parejiya A, Essehli R, Westover
approach. Chem Mater 20(20):6410–6419. https://​doi.​org/​10.​ A, Tsai W-Y, Nanda J, Belharouak I (2021) Appl ACS Energy
1021/​cm800​208k Mater 4(10):11684
150. Ortiz-Mosquera JF, Nieto-Munoz AM, Rodrigues AC (2020) 168. Valle JM, Huang C, Tatke D, Wolfenstine J, Go W, Kim Y, Saka-
New ­Na1+ xGe2(SiO4)x(PO4)3–x NASICON Series with Improved moto J (2021) Characterization of hot-pressed von Alpen type
Grain and Grain Boundary Conductivities. ACS Appl Mater NASICON ceramic electrolytes. Solid State Ionics 369:115712
Interfaces 12(12):13914–13922. https://​doi.​org/​10.​1021/​acsami.​ 169. Kim J, Jo SH, Bhavaraju S, Eccleston A, Kang SO (2015) Low
9b230​65 temperature performance of sodium–nickel chloride batteries
151. Park H, Kang M, Park Y-C, Jung K, Kang B (2018) Improving with NaSICON solid electrolyte. J Electroanal Chem 759:201–
ionic conductivity of NASICON (­ Na3Zr2Si2PO12) at intermediate 206. https://​doi.​org/​10.​1016/j.​jelec​hem.​2015.​11.​022
temperatures by modifying phase transition behavior. J Power 170. Ruan Y, Guo F, Liu J, Song S, Jiang N, Cheng B (2019) Opti-
Sources 399:329–336. https://​doi.​org/​10.​1016/j.​jpows​our.​2018.​ mization of ­Na3Zr2Si2PO12 ceramic electrolyte and interface
07.​113 for high performance solid-state sodium battery. Ceram Int
152. Pal SK, Saha R, Kumar GV, Omar S (2020) Designing High Ionic 45(2):1770–1776. https://​doi.​org/​10.​1016/j.​ceram​int.​2018.​10.​
Conducting NASICON-type ­Na3Zr2Si2PO12 Solid-Electrolytes 062
for Na-Ion Batteries. J Phys Chem C 124(17):9161–9169. https://​ 171. Uchida Y, Hasegawa G, Shima K, Inada M, Enomoto N, Aka-
doi.​org/​10.​1021/​acs.​jpcc.​0c005​43 matsu H, Hayashi K (2019) Insights into sodium ion transfer at
153. Srikanth V, Subbarao EC, Agrawal DK, Huang CY, Roy R, Rao the Na/NASICON interface improved by uniaxial compression.
GV (1991) Thermal expansion anisotropy and acoustic emission ACS Appl Energy Mater 2(4):2913–2920. https://​doi.​org/​10.​
of ­NaZr2P3O12 family ceramics. J Am Ceram Soc 74(2):365–368. 1021/​acsaem.​9b002​50
https://​doi.​org/​10.​1111/j.​1151-​2916.​1991.​tb068​88.x 172. Zhu Y, He X, Mo Y (2015) Origin of Outstanding Stability in
154. Naqash S, Gerhards M-T, Tietz F, Guillon O (2018) Coefficients the Lithium Solid Electrolyte Materials: Insights from Thermo-
of Thermal Expansion of Al-and Y-Substituted NaSICON Solid dynamic Analyses Based on First-Principles Calculations. ACS
Solution ­Na3+2xAlxYxZr2−2xSi2PO12. Batteries 4(3):33. https://​ Appl Mater Interfaces 7(42):23685–23693. https://​doi.​org/​10.​
doi.​org/​10.​3390/​batte​ries4​030033 1021/​acsami.​5b075​17
155. Oota T, Yamai I (1986) Thermal Expansion Behavior of 173. Zhu Y, He X, Mo Y (2016) First principles study on electro-
­NaZr2(PO4)3  Type Compounds. J Am Ceram Soc 69(1):1–6. chemical and chemical stability of solid electrolyte–electrode
https://​doi.​org/​10.​1111/j.​1151-​2916.​1986.​tb046​82.x interfaces in all-solid-state Li-ion batteries. J Mater Chem A
156. Kaus M, Guin M, Yavuz M, Knapp M, Tietz F, Guillon O, Ehren- 4(9):3253–3266. https://​doi.​org/​10.​1039/​C5TA0​8574H
berg H, Indris S (2017) Fast ­Na+ ion conduction in NASICON- 174. Lacivita V, Wang Y, Bo S-H, Ceder G (2019) Ab initio investiga-
type ­Na3.4Sc2(SiO4)0.4(PO4)2.6 observed by 23Na NMR relaxom- tion of the stability of electrolyte/electrode interfaces in all-solid-
etry. J Phys Chem C 121(3):1449–1454. https://​doi.​org/​10.​1021/​ state Na batteries. J Mater Chem A 7(14):8144–8155. https://d​ oi.​
acs.​jpcc.​6b105​23 org/​10.​1039/​C8TA1​0498K
157. Alamo J, Roy R (1984) Ultralow-expansion ceramics in the sys- 175. Fu H, Yin Q, Huang Y, Sun H, Chen Y, Zhang R, Yu Q, Gu
tem ­Na2O-ZrO2-P2O5-SiO2. J Am Ceram Soc 67(5):c78–c80 L, Duan J, Luo W (2019) Reducing interfacial resistance by

13
Ionics

Na-SiO 2  composite anode for NASICON-based solid-state all-solid-state sodium batteries with robust ceramic interface
sodium battery. ACS Mater Lett 2(2):127–132. https://​doi.​org/​ between rigid electrolyte and electrode materials. Nano Energy
10.​1021/​acsma​teria​lslett.​9b004​42 65:104040. https://​doi.​org/​10.​1016/j.​nanoen.​2019.​104040
176. Shi C, Wang L, Chen X, Li J, Wang S, Wang J, Jin H (2022) Chal- 184. Liu L, Qi X, Ma Q, Rong X, Hu Y-S, Zhou Z, Li H, Huang X,
lenges of layer-structured cathodes for sodium-ion batteries. Nanoscale Chen L (2016) Toothpaste-like electrode: A novel approach to
Horiz 7(4):338–351. https://​doi.​org/​10.​1039/​D1NH0​0585E optimize the interface for solid-state sodium-ion batteries with
177. Wei F, Zhang Q, Zhang P, Tian W, Dai K, Zhang L, Mao J, Shao ultralong cycle life. ACS Appl Mater Interfaces 8(48):32631–
G (2021) Review—research progress on layered transition metal 32636. https://​doi.​org/​10.​1021/​acsami.​6b117​73
oxide cathode materials for sodium ion batteries. J Electrochem 185. Zhang Z, Xu K, Rong X, Hu Y-S, Li H, Huang X, Chen L
Soc 168(5):050524. https://​doi.​org/​10.​1149/​1945-​7111/​abf9bf (2017) ­Na3.4Zr1.8Mg0.2Si2PO12  filled poly (ethylene oxide)/
178. Bhange DS, Anang DA, Ali G, Park J-H, Kim J-Y, Bae J-H, Yoon Na(CF3SO2)2N as flexible composite polymer electrolyte for
WY, Chung KY, Nam K-W (2020) N ­ aFeSnO4: Tunnel structured solid-state sodium batteries. J Power Sources 372:270–275.
anode material for rechargeable sodium-ion batteries. Electro- https://​doi.​org/​10.​1016/j.​jpows​our.​2017.​10.​083
chem Commun 121:106873. https://​doi.​org/​10.​1016/j.​elecom.​ 186. Płcharski J, Weiczorek W (1988) PEO based composite solid
2020.​106873 electrolyte containing nasicon. Solid State Ionics 28:979–982.
179. Nuti M, Spada D, Quinzeni I, Capelli S, Albini B, Galinetto P, https://​doi.​org/​10.​1016/​0167-​2738(88)​90315-3
Bini M (2020) From tunnel NMO to layered polymorphs oxides 187. Wang X, Chen J, Mao Z, Wang D (2022) Effective resistance
for sodium ion batteries. SN Appl Sci 2(11):1893. https://d​ oi.o​ rg/​ to dendrite growth of NASICON solid electrolyte with lower
10.​1007/​s42452-​020-​03607-z electronic conductivity. Chem Eng J 427:130899. https://d​ oi.o​ rg/​
180. Barpanda P, Lander L, Nishimura S-I, Yamada A (2018) Poly- 10.​1016/j.​cej.​2021.​130899
anionic Insertion Materials for Sodium-Ion Batteries. Adv Energy 188. Goodenough JB, Singh P (2015) Solid electrolytes in recharge-
Mater 8(17):1703055. https://​doi.​org/​10.​1002/​aenm.​20170​3055 able electrochemical cells. J Electrochem Soc 162(14):A2387–
181. Niu Y, Zhang Y, Xu M (2019) A review on pyrophosphate frame- A2392. https://​doi.​org/​10.​1149/2.​00215​14jes
work cathode materials for sodium-ion batteries. J Mater Chem
A 7(25):15006–15025. https://​doi.​org/​10.​1039/​C9TA0​4274A Publisher's note Springer Nature remains neutral with regard to
182. Wang D, Bie X, Fu Q, Dixon D, Bramnik N, Hu YS, Fauth F, jurisdictional claims in published maps and institutional affiliations.
Wei Y, Ehrenberg H, Chen G, Du F (2017) Sodium vanadium
titanium phosphate electrode for symmetric sodium-ion batter- Springer Nature or its licensor holds exclusive rights to this article under
ies with high power and long lifespan. Nat Commun 8:15888. a publishing agreement with the author(s) or other rightsholder(s);
https://​doi.​org/​10.​1038/​ncomm​s15888 author self-archiving of the accepted manuscript version of this article
183. Lan T, Tsai C-L, Tietz F, Wei X-K, Heggen M, Dunin-Borkowski is solely governed by the terms of such publishing agreement and
RE, Wang R, Xiao Y, Ma Q, Guillon O (2019) Room-temperature applicable law.

13

You might also like