Next Article in Journal
Review of Quasi-Solid Aqueous Zinc Batteries: A Bibliometric Analysis
Previous Article in Journal
Performance of Low-Cost Energy Dense Mixed Material MnO2-Cu2O Cathodes for Commercially Scalable Aqueous Zinc Batteries
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

All-Solid-State Anode-Free Sodium Batteries: Challenges and Prospects

by
Alexander M. Skundin
* and
Tatiana L. Kulova
A.N. Frumkin Institute of Physical Chemistry and Electrochemistry of the Russian Academy of Science, 119071 Moscow, Russia
*
Author to whom correspondence should be addressed.
Batteries 2025, 11(8), 292; https://doi.org/10.3390/batteries11080292 (registering DOI)
Submission received: 4 July 2025 / Revised: 25 July 2025 / Accepted: 28 July 2025 / Published: 2 August 2025

Abstract

All-solid-state anode-free sodium batteries present a special and especially important kind of energy storage device. Unfortunately, the industrial production of such batteries has been absent up to now, although the prospects of their development seem to be rather optimistic. The present mini review considers the fundamental advantages of all-solid-state anode-free sodium batteries as well as challenges in their creation. The advantages of all-solid-state anode-free sodium batteries reveal themselves when comparing them with ordinary sodium-ion batteries, sodium metal batteries, sodium batteries with liquid electrolyte, and their lithium counterparts.

Graphical Abstract

1. Introduction

So-called “anode-free” batteries are a specific but very promising kind of battery belonging to the post-lithium-ion era. Such batteries were first mentioned at the very end of the last century [1]. These were lithium metal batteries with solid electrolytes. Subsequently, various versions of anode-free lithium metal and sodium metal batteries with both solid and liquid electrolytes appeared. Surely, these batteries have two electrodes, and during discharge, one of them is the cathode and the other is the anode. Anode-free batteries are manufactured in a completely discharged state, and the negative electrode consists only of a current collector without active mass. During the first charge, lithium or sodium is deposited on this current collector at the expense of ions contained in the active material of the positive electrode. In the literature, one can find various synonyms for the term “anode-free battery”, such as “anode-less battery”, “zero-excess Li-metal battery”, “reservoir-free battery”, and so on.
The absence of active material on the negative electrode of a freshly assembled battery ensures a certain reduction in its weight and volume, i.e., an increase in energy density. An even more important advantage of anode-free sodium batteries over usual sodium metal ones is the significant simplification and reduction in cost of their production technology (at the assembly stage).
Unlike conventional batteries, the “electrolyte/negative electrode current collector” interface is of great importance for anode-free batteries, along with the “electrolyte/positive electrode active material” and “electrolyte/negative electrode active material” interfaces. There is currently extensive literature on anode-free lithium and sodium batteries with liquid electrolytes and on anode-free lithium batteries with solid electrolytes. Studies on the development of anode-free sodium batteries with solid electrolytes are still in their infancy. This very circumstance was the authors’ main motivation for writing this mini review. At the same time, many of the problems of such batteries are also inherent in anode-free sodium batteries with liquid electrolytes [2,3,4,5,6,7,8] and anode-free lithium batteries with solid electrolytes [7,9,10,11], as well as conventional (non anode-free) sodium batteries with solid electrolytes [12,13,14,15,16,17]. For the most part, these problems are related to sodium deposition during charging in the form of dendrites and their encapsulation, i.e., deposition of a solid electrolyte interphase (SEI) film on them.
The deposition of dendrites forming upon cathodic sodium is known to be connected with different factors. Firstly, in anode-free sodium metal batteries (no matter with liquid or solid electrolytes), sodium metal deposits directly onto foreign material of the current collector. The nucleation process in this case needs a rather high activation energy; therefore, it is non-uniform what is favorable for the growth of separate whiskers. Secondly, the above-mentioned SEIs, as a rule, contain a significant amount of flaws, which results in non-uniform electric fields and uncontrolled dendritic growth. The dendrite growth inevitably results in the breakdown of the separators and ultimately results in the short circuit of the battery. Sodium is a rather strong reducer, and in the course of its cathode deposition, a fresh, very active surface is formed. It is on this active surface that a passive film (SEI) grows, and since sodium is deposited in the form of dendrites, in many cases, the film completely envelops individual sodium particles, preventing their electronic contact with the current collector (encapsulation phenomenon).
The interface effect between electrode and solid electrolyte materials is the most critical issue for modern all-solid-state sodium batteries (both anode-free and ordinary). An ideal (or, at least, proper) interface between electrolyte and electrodes should have good compatibility and suitable mechanical strength as well as high ionic conductivity. In other words, this interface must reveal good sodiophility. Unfortunately, the contact of sodium metal and solid electrolytes faces interrelated challenges such as sluggish charge transfer kinetics, high interfacial resistance, uneven potential, and localized current density distribution, ultimately accelerating dendrite development.
The solid electrolytes for anode-free and ordinary sodium batteries must have high ionic conductivity along with a sufficiently high transference number of sodium ions, low electronic conductivity, a wide electrochemical window, chemical and mechanical stability, good mechanical properties, facile preparation, and environmental friendliness.
The solution to these problems usually comes down to modifying the current collector, electrolytes, and cycling mode (protocol).

2. The Interface Between a Solid Electrolyte and a Current Collector

Electrochemical processes at the interface between the solid electrolyte and the negative electrode current collector occur in anode-free batteries not only at the first charge, but also at further cycling during normal battery operation. In principle, such processes can also occur in a conventional (non-anode-free) sodium battery with a solid electrolyte during full discharge (or overdischarge) of the battery, but this is more likely to be related to improper operation (abuse) and should be avoided. The negative electrode current collector in a sodium battery can be made of different materials, and it is important to note that, unlike anode-free lithium batteries, in sodium batteries, this current collector can be made of both copper and aluminum, since sodium does not form alloys with aluminum and is not inserted into it.
The processes at the interface of the current collector with a solid electrolyte differ to a certain extent from the processes at the interface of the current collector with a liquid electrolyte. In the latter case, the entire surface of the current collector wetted with electrolytes is the contact surface. At the interface of the solid current collector with a solid electrolyte, contact is ensured only at individual points (spots), and the configuration of these spots may depend on the degree of compression of the electrolyte and the current collector (Figure 1).
A salient feature of such interfaces is the presence of voids (pores). As can be seen, the area of the electrochemically active surface is smaller than the total surface area, which leads to a certain decrease in the total current. This phenomenon, so-called current constriction, is well known in physics for contact between metals or semiconductors [19,20,21], but it has not been sufficiently studied in electrochemistry [18,22,23,24,25]. A natural consequence of current constriction is an increase in the total interfacial resistance [26,27], which also depends on the charge transfer resistance, the presence of SEI, and the presence of various contaminants. Of course, the picture shown in Figure 1 is not static, but it changes continuously when the current passes; the pore volume increases during the anodic dissolution of the metal and decreases during cathodic deposition. According to Faraday’s laws, the passage of 1 Ah is accompanied by a change in the volume of sodium by 0.88 cm3, which is 1.7 times greater than during the deposition/dissolution of lithium. In this regard, batteries with solid electrolytes compare favorably with batteries with liquid electrolytes, in which there is no free volume and the deposition of metallic sodium leads to the occurrence of internal stresses. Accordingly, the interfacial resistance increases in the anodic phase of cycling and decreases in the cathodic phase [24,25].
Some features of the interfaces of solid electrodes with a solid electrolyte associated with the occurrence of current constriction were studied using examples of different metals and different electrolytes. Thus, in [28,29,30,31], the interfaces of silver with AgI and AgBr and copper with CuBr, as well as platinum with CoO, were studied, and the occurrence of potential oscillations under galvanostatic conditions was discovered. The characteristic features of these oscillations depend on the current density, temperature, and compression force (pressure). Potential oscillations arise as a result of the periodic formation and disappearance of micropores and the corresponding changes in the contact surface area. Electrochemical dissolution occurs mainly at the points of dislocation exit to the surface, and it is in these places that micropores are formed. The walls of the micropores provide the rapid diffusion of metal adatoms.
Similar conclusions regarding the nature of current constriction were obtained in [32,33,34] using the example of lithium deposition from amorphous solid electrolyte LiPON (lithium phosphorus oxynitride) on substrates of different metals. In all cases, an increase in the current density of lithium deposition led to a decrease in the size of nuclei and to an increase in their number per unit surface area. The nucleation overpotential during lithium deposition on platinum (with which lithium forms an alloy) turned out to be significantly lower than the nucleation overpotential during lithium deposition on copper (when alloy formation is excluded). Potential oscillations were also observed during lithium deposition on platinum.
At the interface between copper and a crystalline lithium-conducting electrolyte with the garnet structure Li7La3Zr2O12 (LLZO), oscillations of even the open-circuit potential were observed, which, in turn, depended on the compression force [35] (Figure 2). An increase in the number of nuclei with increasing lithium deposition current density was also noted here, as well as preferential nucleation on surface defects [29,36].
The most interesting and important aspect from this article’s point of view is the work of [37], which studied the processes of sodium deposition and dissolution at the boundary of a copper current collector and a solid electrolyte NaSICON (Na SuperIonic CONductor). The electrolyte in this work had the composition Na3.4Zr2Si2.4P0.6O12 (NZSP) (it should be noted that this is the electrolyte often used in studies of all-solid-state sodium-metal batteries [14,38,39,40,41,42,43,44,45,46]; one can find in the literature both notations “NaSICON” and “NASICON”). The processes of nucleation and deposit growth were studied at different current densities and different stack pressures using time-dependent impedance spectroscopy, confocal light microscopy, and electron microscopy. The studies were carried out using the electrochemical system Cu|NZSP|Na, where sodium served simultaneously as an auxiliary electrode and a reference one (quasi-reference electrode). The reliability of using such a reference electrode was confirmed in [38]. The galvanostatic curves of sodium deposition had an usual appearance, with a sharp shift of the potential to the negative direction at the moment of switching on the current, corresponding to the nucleation overpotential, and a subsequent gradual shift of the potential to the positive direction until reaching a steady-state value, corresponding to the current density [32,47,48]. Intriguingly, in this stationary section of the galvanostatic curve, sporadic fluctuations were recorded, the amplitude of which decreased somewhat with an increase in current density and stack pressure. Potential fluctuations were caused by local delamination of the copper current collector from the solid electrolyte and the corresponding increase in interphase resistance.
It is interesting that the nucleation overvoltage at a constant stack pressure was practically independent on the current density, but at a constant current density, it increased linearly with the increase in stack pressure (Figure 3). The latter fact is most intriguing, and it could be associated with a change in the pattern of contact between the electrode and the electrolyte (Figure 1). It should be emphasized here that it is the overvoltage of the nucleation stage that is considered, and not the deposit growth in the steady state (in the latter case, the overvoltage increases with increasing current density [37]). It is important to note that the influence of the stack pressure on the processes at the boundary of the current collector with the solid electrolyte has a clear physical nature. Some effects described for the boundary of the current collector with the liquid electrolyte [49,50] more likely than not are some artifacts.
At the same time, an increase in the current density was found to result in the formation of more uniform sodium deposits. Surprisingly, in this work, a practical absence of the effect of stack pressure on the morphology of sodium deposits was found; although, for lithium deposition under similar conditions, this effect is very noticeable [35,51]. This difference in the nature of lithium and sodium deposition is explained by the higher plasticity of sodium (the yield strength for lithium is 16 MPa [52], while for sodium it is only 0.2 MPa [53]).
In [36], the processes of sodium deposition and dissolution at the interface of a current collector made of aluminum with a carbon black coating and NZSP as an electrolyte were studied using electron backscatter diffraction (EBSD). It was found that sodium under these conditions was deposited in the form of fairly large crystallites (up to 150 μm wide), with the intercrystallite boundaries being strictly perpendicular to the plane of the current collector. During anodic dissolution of these deposits, voids (pores) were formed, and the interfacial resistance increased sharply.
To reduce the interfacial resistance at the interface between the solid electrolyte Na3Zr2Si2PO12 (NZSP(it is worth noting that the abbreviation NZSP is used for all stoichiometry of the compounds NaxZr2SiyPzO12, and proper values of indexes x, y, z must be clear from context)) and the aluminum current collector, the work [54] proposes introducing an intermediate elastic ferroelectric layer with a thickness of about 3 μm, containing nanoparticles of barium titanate and carbon. Such a layer on the one hand dampens volumetric changes during the deposition and dissolution of sodium, and on the other hand, it makes the sodium ion flows more uniform. Accordingly, the sodium deposition is not in the form of dendrites but in the form of a smooth layer. In addition, the presence of this layer leads to a noticeable decrease in the nucleation overvoltage (from 53 to 4 mV at a current density of 0.1 mA/cm2, and from 73 to 14 mV at a current density of 1 mA/cm2). A laboratory model of an anode-free battery with the described intermediate layer, NZSP as an electrolyte, and a positive electrode based on Na3V2(PO4)3 (NVP) withstood 300 cycles with a current density of 0.1 mA/cm2 with an insignificant loss of capacity (4.6%).
In [55,56], the initial stages of sodium whisker deposition at the interface of a gold current collector and polycrystalline Na-β″-alumina as an electrolyte were studied using the elegant method of in situ transmission electron microscopy (TEM) in combination with crystal orientation analysis (Figure 4). The sodium whiskers were found to grow along grain boundaries, and nucleation occurs not only directly at the interface of the electrolyte with gold, but also in the bulk of the electrolyte at the grain boundaries.
Thus, reducing the intergranular volume (in other words, increasing the relative density of the solid electrolyte) is an effective means of combatting dendrite formation.

3. The Interface Between a Solid Electrolyte and Sodium

In anode-free sodium batteries, the interface “solid electrolyte/sodium” appears just after the start of charging a freshly assembled or completely discharged battery. Naturally, such a boundary is also present in conventional (not anode-free) sodium metal batteries, and in the literature, one can find a lot of works on the study of such a boundary (see, for example, the review [57]). Of course, the fundamental issue is the compatibility of the solid electrolyte and sodium [43,58]. The interface “solid electrolyte/sodium” has many features of the interface “solid electrolyte/current collector”, including the presence of voids and contact spots and, in general, the phenomenon of current constriction and, accordingly, increased interphase resistance. A significant problem is the risk of sodium deposition in the form of dendrites or whiskers, and the solution to this problem is associated with the modification of the solid electrolyte/sodium interface. It is the presence of voids and contact spots at such an interface that leads to uneven current density distribution and promotes the formation of dendrites [27,45,59]. It is worth noting that, since sodium has higher plasticity and a higher self-diffusion coefficient, the risk of dendrite formation at the solid electrolyte/sodium interface is less than at the solid electrolyte/lithium interface [60].
The penetration of sodium dendrites into a solid electrolyte depends on the presence of intercrystalline pores, i.e., on the deviation of the true density of the solid electrolyte from theoretical values, as well as on the ionic and electronic conductivity of the electrolyte (naturally, the former should be as large as possible, the latter as small as possible). In [61], it was shown that doping the solid electrolyte NASICON with ions of rare earth elements Pr3+, Eu3+, and Lu3+, which have a larger radius than the replaced ion Zr4+, leads to a noticeable increase in density, a decrease in electronic conductivity, and, accordingly, to a decrease in the risk of dendrite formation.
SEI on the sodium surface adjacent to a solid electrolyte compares favorably with SEI formed in a liquid electrolyte. In [39], the composition of SEI on a sodium electrode in a liquid electrolyte (1 M NaClO4 in the mixture 95% propylene carbonate and 5% fluoroethylene carbonate) and in contact with a solid electrolyte (NASICON, Na3Zr2Si2PO12) was studied using time-of-flight secondary ion mass spectrometry (TOF-SIMS). Figure 5 shows the material distribution across the SEI thickness in both cases. In the SEI formed in the liquid electrolyte, the products of electrolyte decomposition (ions CH2O, C2F, NaO, and ClO2) as well as the products transferred from the positive electrode (ions MnO, FeO, MnF3, and FeF3) were identified. In contact with the solid electrolyte, SEI was formed as a sodium oxide. It is important to emphasize that the SEI in contact with the solid electrolyte was much thinner than the SEI in the liquid electrolyte.
In [62] it is noted that natural SEIs at the sodium/NASICON interfaces have much higher protective properties than SEIs at the lithium/solid electrolyte interfaces.
High-quality SEI at the sodium/solid sulfide electrolyte (Na2.895W0.3Sb0.7S4) interface is obtained from a film of polyperfluoromethyl isopropyl ether [CF(CF3)CF2O]x(CF2O)y. When in contact with sodium metal, this film is enriched by NaF and reliably prevents dendrite growth and reduces interfacial resistance [63].
Various strategies of modification of the solid electrolyte/sodium interface contribute to the improvement of the quality of this boundary, including the reduction of the interfacial resistance and the suppression of dendrite formation. Thus, in [40], it was shown that one of the reasons for the increased interfacial resistance at the NZSP-sodium interface is the presence of functional groups (in particular, hydroxyl and carbonate) on the NZSP surface. Annealing of NZSP in air at a temperature of 800 °C allows for cleaning the surface, and additional annealing at a temperature of 1000 °C leads to partial surface restoration and an increase in sodiophilicity and completely eliminates dendrite formation. It was found, in [41], that annealing of NZSP leads to the formation of a protective nanometric layer of Na3PO4. Annealing of NASICON to clean its surface from oxygen-containing groups is also recommended in [64]. In [65,66], it was shown that the tendency to form oxygen-containing functional groups on the NZSP surface depends on the annealing mode. Synthesis of NZSP in a two-stage mode (first stage 1300 °C, 15 min, second stage 1200 °C, 12 h) allows for it to reduce the number of surface groups, increase the ceramics density, and also increase its ionic conductivity.
A similar result is described in [67], where it is shown that surface treatment of NASICON with liquid sodium at a temperature of 380 °C leads to the formation of a reduced sodiophilic surface layer, which provides a tenfold decrease in interfacial resistance. In [14], it is proposed to treat the surface of NZSP with a suspension of SbF3 in liquid sodium. With such treatment, the surface of the solid electrolyte is covered with a highly sodiophilic layer of Na-Sb alloy containing NaF particles, which provide ionic conductivity. An original method for creating a sodiophilic surface of a solid electrolyte is described in [68]. Here, a porous layer of the same electrolyte about 0.1 mm thick is applied to the surface of a monolithic electrolyte (NZSP doped with Ca2+). The porous layer is impregnated with a SnCl2 solution and then annealed in air, converting all the tin into SnO2. After this, the porous layer is impregnated with liquid sodium, whereby the pore walls are coated with metallic tin, and the entire porous layer acquires increased sodiophilicity.
A somewhat similar technique was used in [69]. Here, a thin SnO2 layer was applied to the surface of a solid NASICON-like electrolyte of the composition Na3.2Hf1.9Ca0.1Si2PO12. Upon contact with sodium metal, an intermediate Na2O/NaxSn layer was formed, which provided low interfacial resistance. To create a similar sodiophilic layer on the NASICON surface, the authors of [46] recommend applying a saturated aqueous solution of lead acetate to the surface and annealing in an argon atmosphere at a temperature of 550 °C. This creates a Pb/C@NZSP composite layer that is perfectly wetted by sodium. The interfacial resistance at the boundary of sodium with the solid electrolyte upon application of such a layer decreased from 391 to 1.5 Ohm·cm2.
The authors of [70] proposed applying a suspension of Super P carbon black in a liquid Na–K alloy to the surface of a solid electrolyte Na3Hf2Si2PO12 (NHSP) to form a quasi-liquid composite layer C@Na-K with a thickness of about 7 μm (such a suspension has the properties of a non-Newtonian liquid).
In [71], it is proposed to apply a thin layer of AlF3 to the NASICON surface (from an aqueous solution with subsequent drying and annealing at a temperature of 450 °C in a nitrogen atmosphere). This 350 nm thick AlF3 layer evens out the unevenness of the NASICON surface and therefore reduces the tendency to dendrite formation and interfacial resistance. When such a modified NASICON comes into contact with metallic sodium, Al, NaF, and also NamAlnF3 are formed. The resulting layer reliably prevents the deposition of sodium dendrites.
Another method for reducing the interfacial resistance at the solid electrolyte/sodium interface is by introducing a conducting polymer layer. In the already cited work [67], cross-linked poly (ethylene glycol) methyl ether acrylate (CPMEA) is proposed as such a polymer. In [72], a similar layer with a thickness of about 20 μm is made of a nanoporous product of copolymerization of 5,5′,6,6′-tetrahydroxy-3,3,3′,3′-tetramethyl-1,1′-spirobisindane and tetrafluoroterephthalo-nitrile, the molecular structure of which is shown in Figure 6.
A sodiophilic interlayer made of sulfurized polyacrylonitrile (sulfurized PAN) on the surface of NASICON is proposed in [73].
In addition to polymers, the sodiophilic interlayer can be made simply of metals, in particular tin [74], oxides, e.g., TiO2 [75], or carbon materials, specifically nanotubes [76] or graphene [77].
A relatively simple way to reduce the interfacial resistance at the solid electrolyte/sodium interface is proposed in [78]. It is shown here that doping Na3Zr2Si2PO12 with Zn2+ ions (replacing part of Zr with Zn) results in an increase in the electrolyte conductivity and, most importantly, in a decrease in the interfacial resistance. The electrolyte of the composition Na3.20Zr1.90Zn0.10Si2PO12 has a specific conductivity of 1.58 mS/cm, which is four times greater than that of the undoped electrolyte, and the interfacial resistance at the boundary with sodium for such an electrolyte was 25 Ohm·cm2, whereas, for the undoped electrolyte, this resistance was 785 Ohm·cm2. It was shown that zinc segregates on the surface of the solid electrolyte, where a layer of sodium-zinc alloy is formed, ensuring the sodiophilicity of the interface. The beneficial effect of doping the NASICON electrolyte with zinc was also reported in [79].
An even simpler way to reduce the interfacial resistance at the solid electrolyte/sodium interface is described in [80]. It is shown here that ultrasonic treatment of such an interface at room temperature for only 25 s (practically ultrasonic welding) results in the formation of a very close contact between sodium and the solid electrolyte NASICON (Na3Zr2Si2PO12) (Figure 7).
A symmetrical Na|Na3Zr2Si2PO12|Na cell assembled using such ultrasound solid welding successfully withstood more than 200 cycles at a current density of 0.1 mA/cm2 (1300 h) without any signs of dendrite formation or increase in overvoltage, which remained equal to 12 mV. A similar cell assembled without ultrasound solid welding did not withstand even 20 cycles.
The effect of stack pressure on the interfacial resistance at the solid electrolyte/current collector interface has already been noted above. Indicative data on the effect of uniaxial compression on the resistance at the solid electrolyte/sodium interface are presented in the study [81]. Here, the resistance of a NASICON pellet sandwiched between two sodium plates was measured at different compression pressures. The results are shown in Figure 8.
One can see that even a not too large compression pressure results in a decrease in interface resistance by an order of magnitude.
In the end of this section, it stands to right mentioning Refs. [82,83]. By using the density functional theory (DFT), it is shown here that monolayer coatings from some chalcogenides, namely ScTe2 and V2S2, demonstrate the excellent sodiophilic properties.

4. The Interface Between a Solid Electrolyte and a Cathode

The interface between solid electrolyte and positive electrode in anode-free sodium batteries differs both from the interface between a solid electrolyte and negative electrode in such batteries and from the interface between a solid electrolyte and positive electrode in batteries with liquid electrolytes (although, of course, the interface between a solid electrolyte and cathode in anode-less sodium batteries is no different from the same interface in conventional, non-anode-free batteries). First of all, it is worth noting that there is a wide variety of possible active materials for positive electrodes. In addition to the most popular sodium vanadate-phosphate Na3V2(PO4)3 [80,84,85,86,87,88,89], the following oxides can be named: NaxCoO2 [90], NaCrO2 [91], NaNi0.4Fe0.2Mn0.4O2 (NFM) [92], Na0.66Ni0.33Mn0.67O2 [93], P2–Na0.68Li0.10Ni0.25Mn0.63Ca0.01Ti0.01O2 (NM-L10CN) [94], sodium titanophosphate NaTi2(PO4)3 [67], double-cyano-perovskite Na2MnFe(CN)6 [39], perylene-3,4,9,10-tetracarboxylic dianhydride (PTCDA) [95], and pyrene-4,5,9,10-tetraone (PTO) [78].
Unlike the negative electrode, the positive one (cathode during discharge) contains a porous (essentially granular) active layer of a certain thickness. In order for such a layer to be able to operate across its entire thickness, it consists of a mixture of active material particles and a solid electrolyte, with the volume of solid electrolyte particles exceeding the percolation threshold to maintain the connectivity of all these particles. It is worth recalling that a certain amount of conductive additive, often carbon black, is also introduced into the active mass of the positive electrode to ensure electronic conductivity across the entire depth of the active layer. In addition, the actual particles of the active material are often coated with a nanometer layer of carbon. For example, in work [88], the active mass of the positive electrode included active material (Na3V2(PO4)3 with a carbon coating), a solid electrolyte (Na3.4Zr1.8Mg0.2Si2PO12), a conductive additive (super P), and binder (PVDF—polyvinylidene fluoride) in a weight ratio of 65:20:5:10.
If the liquid electrolyte permeates the entire porous space and enables the entire internal surface to participate in the current-producing process, then the mixture of particles of the active material and the solid electrolyte provides only point-to-point contacts. A more elegant method of organizing good contact between the active material and the electrolyte is to coat each particle of the active material with the electrolyte, for example, using pulse laser deposition.
As in the case of the negative electrode, cycling (sodium insertion/extraction) of the positive electrode is accompanied by volume changes, however, they are generally smaller than that at the negative electrode. Thus, when extracting two sodium ions from Na3V2(PO4)3 (during charging), the volume decrease is only 0.21 cm3/Ah, i.e., four times less than the change in the volume of sodium (with due account for the density of Na3V2(PO4)3 equal to 3.156 g/cm3 [87,96]). A significant difference between the interfaces of the solid electrolyte with the negative and positive electrodes also consists of the fact that, in the former case, the volume changes are one-dimensional and directed normally to the substrate surface, while, in the latter case, there is a three-dimensional change in the volume of each particle (granule) of the active material.
To improve the contact of the cathode material with the solid electrolyte (in essence, to reduce the interfacial resistance), in [85], it was proposed to wet the contact area of the NASICON solid electrolyte with the positive electrode based on sodium vanadate phosphate with a very small amount (0.5 μL/cm2) of liquid electrolyte (0.8 M NaPF6 in a mixture of ethylene carbonate with dimethyl carbonate) (Figure 9). This approach allowed them to significantly improve the battery characteristics, but due to partial evaporation and decomposition of the electrolyte, the cyclic resource was limited. A noticeable improvement was achieved by replacing the carbonate electrolyte with the same amount of ionic liquid (N-methyl-N-propylpiperidinium-bis(fluorosulfonyl) imide (PP13FSI)). In this version, the battery withstood 10,000 full charge–discharge cycles in the 10 C rate.
A similar approach was used in [93]. Here, the active mass of the positive electrode is a paste (in the authors’ terminology, “toothpaste”) of Na0.66Ni0.33Mn0.67O2 particles with a carbon coating and N-butyl-N-methylpyrrolidinium bis(fluorosulfonyl)imide (PY14FSI) ionic liquid, the structural formula of which is shown in Figure 10. This paste is simply applied to the surface of Na-β″-Al2O3 solid electrolyte using a doctor blade.
The authors of [84] proposed impregnating the active mass of the positive electrode with a so-called plastic–crystal Na+-electrolyte. This electrolyte is a complex obtained by heating a mixture of succinonitrile (N≡C(CH2)2C≡N) and NaClO4 with a molar ratio of 20:1 at 65 °C. When cooled to room temperature, it solidifies, remaining X-ray amorphous, and it has an ionic conductivity of 3.3 mS/cm, which exceeds the conductivity of the lithium analogue [97].
As already mentioned above, the authors of [67] proposed a CPMEA polymer interlayer to reduce the interfacial resistance at the solid electrolyte/sodium interface. The same work proposes the same interlayer to reduce the interfacial resistance at the solid electrolyte/cathode interface. In this case, the cathode was made of sodium fluorotitanate using CPMEA as a binder that also has the properties of an ionic electrolyte.
Another polymer, namely a copolymer of vinylidene fluoride with hexafluoropropylene (PVDF–HFP) for the same purposes, is proposed in [92]. Here, a mixture of PVDF–HFP with a suspension of the active material NFM in N-methylpyrrolidone was applied directly to the surface of the solid electrolyte (Na3Zr2Si2.2P0.8O12).
In [94], a very sophisticated active material of the positive electrode, specifically Na0.68Li0.10Ni0.25Mn0.63Ca0.01Ti0.01O2 (NM-L10CN), was reported. A battery with such an electrode and solid NZSP electrolyte, successfully withstood 500 cycles at 1 C rate. Intriguingly, this article says nothing about measures to reduce interphase resistance at the boundary of the solid electrolyte with the cathode.
It is known that materials such as Prussian blue, although attractive in terms of specific capacity, are unstable in contact with liquid organic electrolytes. In the above-mentioned work [39], the performance of such material, specifically double-cyano-perovskite with formula Na2MnFe(CN)6 (NMHCF), was assessed in contact with the Na3Zr2Si2PO12 solid electrolyte. The active mass of the positive electrode in this case was made from a mixture of NMHCF, carbon black as an electrically conductive additive, and a solid polymer electrolyte based on polyethylene oxide and NaClO4. Such a battery successfully withstood 200 cycles at the C/2 rate with a loss of only 10% of capacity.

5. Solid Electrolytes

In principle, the solid electrolyte for an anode-free sodium battery is no different from the solid electrolyte for a conventional (not-anode-free) battery. This explains the fact that there are a huge number of publications in the literature on solid electrolytes for conventional sodium batteries (it is high time to mention review papers such as [13,15,43,98,99,100,101,102,103,104,105,106,107]) but practically nothing on anode-free versions. Indeed, we are unaware of research devoted to this very topic. The main requirements for solid electrolytes for anode-free sodium batteries are obvious. Such electrolytes must possess sufficiently high conductivity for sodium ions and a sufficiently high transference number of such ions, a wide electrochemical stability window, chemical and mechanical stability, low electronic conductivity, ability to inhibit dendrites, good mechanical properties, facile preparation, and environmental friendliness. Of particular importance is the requirement for a high sodium ion transference number, since a deviation of this quantity from unity leads to the occurrence of concentration polarization with a corresponding decrease in discharge voltage and energy density.
Solid polymer electrolytes are sometimes included in the category of solid electrolytes, but polymers should be considered supercooled liquids rather than true solids. In addition, solid polymer electrolytes, as a rule, possess rather low sodium ion transference numbers. They are not considered in this article. Among the numerous inorganic solid electrolytes with sodium conductivity, groups such as NASICON-type materials, sulfides [108], and complex hydrides [100,108] can be distinguished. The most popular are NASICON-type materials, which were proposed as an electrolyte in 1976 as an alternative to β-alumina [109,110].
The composition of NASICON-type material can be described by the generalized formula NaMM′(PO4)3, where the M and M’ positions can contain various transition metal ions in the divalent, trivalent, tetravalent, or pentavalent states, and phosphorus can be partially replaced by silicon or arsenic [105]. Standard ceramic methods are used to synthesize NASICON-like electrolytes, i.e., sintering of precursor’s mixtures, which are obtained by simple dry mixing, sol-gel route, hydrothermal method, or ion exchange. Depending on the composition and synthesis method, NASICON can crystallize in the following structures: a rhombohedral (prevalent in NASICON system), monoclinic, triclinic, orthorhombic, corundum-like, etc. NASICON-like electrolytes have a specific ionic conductivity of about 0.1 mS/cm at room temperature, which depends on the specific composition and crystal structure [111]. For example, in [112], it was found that the introduction of a superstoichiometric amount of sodium into standard NASICON (transition from Na3Zr2Si2PO12 to Na3.3Zr2Si2PO12) leads to an increase in the room-temperature specific conductivity from 0.2 to 0.9 mS/cm. A similar result (2.4 mS/cm for Na3.3Zr2Si2PO12) was reported in [113]. In [114], materials of the compositions Na3.4Zr2Si2.5P0.5O12 and Na3.4Zr2Si2.4P0.6O12 with a specific conductivity of about 5 mS/cm were reported.
In [115], it was shown that doping NASICON with yttrium (by replacing the ZrO2 precursor with (ZrO2)0.97(Y2O3)0.03) leads to a fivefold increase in ionic conductivity. In general, doping with cations is an effective means of increasing the ionic conductivity of NASICON [106,116,117,118,119,120,121], with hafnium, yttrium, niobium, cobalt, and zinc being the most effective dopants. Among the NASICON-like materials with the highest ionic conductivity at room temperature, the following should be mentioned: Na3.2Hf2(SiO4)2.2(PO4)0.8 (specific conductivity 2.3 mS/cm) [122], Na3.4Zr1.8Ni0.2Si2PO12 (2.28 mS/cm) [116], Na3.3Hf2(SiO4)2(PO4) (1.1 mS/cm) [122], Na3.36Zn0.08Zr1.92(SiO4)2.2(PO4)0.8 (1.1 mS/cm) [123], Na3Zr1.8Zn0.2Si2PO11.8 (1.44 mS/cm) [120], Na3.232Co0.016Zr1.984(SiO4)2.2(PO4)0.8 (1.41 mS/cm) [124], Na2.1Y0.1Zr1.9(SiO4)2(PO4) (1.98 mS/cm) [117], Na3.5Zr1.75Mg0.25Si2O12 (2.5 mS/cm) [113], Na3.4Sc0.4Zr1.6(SiO4)2(PO4) (4.0 mS/cm) [125], Na2.96Nb0.04Zr1.96(SiO4)2(PO4) (4.95 mS/cm) [117], and Na3.4Zr1.9Zn0.1Si2.2P0.8O12 (5.27 mS/cm) [121].
As was indicated in Section 2, a decrease in the intergranular volume (an increase in the relative density of NASICON) is important from the point of view of combatting dendrite formation. Doping NASICON with cations in some cases leads to a noticeable increase in the relative density [68,85,121,126]. In [61], it was shown that large ions of rare earth elements such as Pr3+, Eu3+, and Lu3+ are particularly effective in this regard. Thus, if the relative density of undoped Na3Zr2Si2PO12 is 87.8%, then for Na3.3Zr1.7Pr0.3Si2PO12, Na3.3Zr1.7Eu0.3Si2PO12, and Na3.3Zr1.7Lu0.3Si2PO12, it increases to 94.2%, 94.5%, and 92.3%.
Besides doping, various technological methods are an important means of increasing ionic conductivity. For example, in [127], a NASICON-like electrolyte was synthesized by sintering a mixture of canonical Na3Zr2Si2PO12 with the addition of 5% Na2SiO3 at a temperature exceeding the melting point of Na2SiO3. The resulting material had an ionic conductivity at room temperature as high as 1.45 mS/cm, which was associated with the formation of a non-stoichiometrically silicon-enriched phase at the intercrystalline boundaries.
Sulfides are the second most important group of inorganic solid electrolytes with sodium ion conductivity [128]. The most important are electrolytes based on Na3PS4 [129], Na3SbS4 [130,131], and Na11Sn2PS12 [132]. Sulfide electrolytes are characterized by a wide electrochemical stability window (over 5 V) and stability to sodium [133,134,135,136]. There are amorphous (glassy), crystalline, and glass-ceramic samples. Plain amorphous examples of the Na2S–P2S5 system have rather modest ionic conductivity at room temperature (about 0.01 mS/cm). The transition to crystalline forms and to glass ceramics makes it possible to increase this conductivity to fractions of mS/cm [129,133,135,137,138,139]. An even greater increase in ionic conductivity can be achieved through doping, mainly anionic [140,141,142,143,144,145,146].
The positive effect of replacing part of the phosphorus in Na3PS4 with arsenic is described in [147]. The material with the composition Na3P0.62As0.38S4 has a room-temperature conductivity of about 1.46 mS/cm. The most effective cationic dopants are silicon, germanium, and tin. According to [145], the ionic conductivity at room temperature for materials with the compositions Na3.125Si0.125P0.875S4 and Na3.0625Sn0.0625P0.9375S4 amounts to 2.99 and 10.7 mS/cm. The ionic conductivity of Na3.1Ge0.1Sb0.9S4 at room temperature is 5.1 mS/cm [146].
In [142,144,148], it is shown that the substitution of sulfur in Na3PS4 with selenium leads to a noticeable increase in ionic conductivity. Thus, according to [144], Na3PSe4 has a specific conductivity of 1.16 mS/cm at room temperature. More importantly, the transference number for the Na+ cation in Na3PSe4 is practically equal to unity.
Probably, the compound Na2.88Sb0.88W0.12S4 has the highest ionic conductivity among doped sulfides; the literature gives room-temperature ionic conductivity values of 2.5 [149], 19 [150], and 32 mS/cm [151]. This discrepancy is due to different heat treatment conditions of the ceramics; the value of 32 mS/cm was obtained after 12 h annealing at 275 °C, which resulted in higher crystallinity and density. Figure 11 and Figure 12 compare the crystal structure of Na2.88Sb0.88W0.12S4 and undoped Na3SbS4. The inclusion of tungsten in crystalline Na3SbS4 results in the appearance of Na vacancies in the lattice, which promotes the formation of a cubic phase, which can facilitate isotropic three-dimensional conductivity of fast ions.
The transference number of the Na+ cation in this case was also close to unity. Interestingly, the Na2.88Sb0.88W0.12S4 composites with sodium iodide have an even higher ionic conductivity, which depends on the NaI content. The maximum conductivity of 36 mS/cm is exhibited by the Na2.88Sb0.88W0.12S4·0.50NaI composite [149] (Figure 13). At the same time, the introduction of NaI into the composite with Na3PS4 does not lead to a noticeable increase in conductivity [152].
The properties of the Na3–xP1–xWxS4 electrolyte depend on both the composition (degree of tungsten addition) and the synthesis method [153]. With the same mechanochemical synthesis, compositions with x = 0.09, 0.15, 0.18, and 0.21 had specific conductivities of 1.84, 8.8, 2.47, and 1.69 mS/cm, respectively. At the same time, according to data from [154], the Na2.9P0.9W0.1S4 electrolyte synthesized by solid-phase synthesis had a conductivity of 13 mS/cm, while Na2.9Sb0.9W0.1S4 had a conductivity of 41 mS/cm. It is worth noting that such properties of Na11Sn2PS12 are not too sensitive to small changes in composition. According to [155], electrolytes with compositions Na11Sn2PS12, Na10.875Sn2PS12, and Na11.125Sn2PS12 have conductivities of 2.4, 2.3, and 2.4 mS/cm, respectively. At the same time, the conductivity of Na10SnP2S12 is only 0.4 mS/cm [156].
Important dopants for sulfide electrolytes are halides. Thus, in [140], it was shown that Na2.9375PS3.9375Cl0.0625 has a conductivity of 1.38 mS/cm at room temperature. In a later work of the same team [157], for Na3PS3.8Cl0.2, Na2.875PS3.875Cl0.125, and Na2.9375PS3.875Cl0.125, the ionic conductivity values of 1.6, 6.38, and 9.03 mS/cm are reported. In [158], a series of halogen-substituted sulfides of the general formula Na3−xSbS4−xHalx (0 ≤ x ≤ 0.3, Hal = Cl, Br, and I) was studied, and it was found that Na2.85SbS3.85Br0.15 has the maximum room-temperature ionic conductivity of 2.87 mS/cm.
Of interest is the effect of doping sulfide electrolytes with oxygen on their conductivity and stability in contact with sodium. The authors of [159] investigated the introduction of small amounts of phosphorus and oxygen (in a ratio of 4 P:10 O) into the Na3SbS4 glass ceramic electrolyte and found that the highest ionic conductivity of 3.82 mS/cm was possessed by the composition Na3SbP0.16S4O0.4, pre-annealed in a vacuum at a temperature of 500 °C for 10 h (annealing led to an increase in crystallinity). At the same time, in [150], it was discovered that the introduction of oxygen into Na2.88Sb0.88W0.12S4 due to the substitution of sulfur led to a slight decrease in ionic conductivity. Moreover, based on first-principle calculations in [160], it was shown that the ionic conductivity of Na7P3O11 at room temperature is only 3 μS/cm, whereas for Na7P3S11 and Na7P3Se11, values of 10.97 and 12.56 mS/cm, respectively, were obtained.
Of other types of solid electrolytes, mention should be made of hydride (and borohydride) systems [161,162], thioborates [163], and the chloride electrolyte (Na1−xZrxLa1−xCl4) [164], although they are of much less importance than NASICON and sulfides.

6. Forecast of Possible Characteristics of Future All-Solid-State Anode-Free Sodium Batteries

As is clear from the Section 1, we are unaware of any literature data on the characteristics of specific all-solid-state anode-free sodium batteries. However, published data on the electrochemical properties of ordinary all-solid-state sodium battery systems provide an idea of the possible performance of future anode-free counterparts. Table 1 provides literature data on various electrochemical systems of all-solid-state sodium batteries.

7. Conclusions

All-solid-state anode-free sodium batteries present a special category of energy storage devices. Up to date, there are neither industrially produced batteries, nor laboratory ones like these such batteries, but the prospects for their creation are quite optimistic. This optimism is explained by the fundamental advantages of solid-state anode-free sodium batteries over other categories.
The advantages of anode-free sodium batteries over sodium-ion batteries are greater specific capacity and energy density. The advantages of anode-free sodium batteries over conventional sodium metal batteries are the absence of operations with metallic sodium during production and assembly, that is, in the radical simplification and reduction of the cost of technology. The advantages of batteries with a solid electrolyte over batteries with a liquid electrolyte are greater chemical stability, good mechanical properties, and increased safety. In addition, solid electrolytes usually have a sodium cation transference number close to one, which ensures an increase in energy density. The advantages of solid-state anode-free sodium batteries over their lithium counterparts include the much lower cost and greater availability of sodium, as well as the ability to use aluminum rather than copper as a current collector.
The above-listed advantages of all-solid-state anode-free sodium batteries have certain economic and environmental aspects. The already mentioned simplification and reduction in cost of the production technology of such batteries, associated with the exclusion of operations with sodium metal, also contribute to a significant scaling of production, which is especially important for energy storage systems. The exclusion of liquid electrolytes provides certain environmental advantages both in the production and in the recovering of batteries.

Author Contributions

A.M.S.—conceptualization, analysis of literature, writing and editing of the original draft; T.L.K.—conceptualization, analysis of literature, writing of original draft. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by Russian Science Foundation, grant number 25-13-00125. https://rscf.ru/en/project/25-13-00125/. Accessed on 30 May 2025.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Neudecker, B.J.; Dudney, N.J.; Bates, J.B. “Lithium-Free” Thin-Film Battery with In Situ Plated Li Anode. J. Electrochem. Soc. 2000, 147, 517–523. [Google Scholar] [CrossRef]
  2. Ni, Q.; Yang, Y.; Du, H.; Deng, H.; Lin, J.; Lin, L.; Yuan, M.; Sun, Z.; Sun, G. Anode-Free Rechargeable Sodium-Metal Batteries. Batteries 2022, 8, 272. [Google Scholar] [CrossRef]
  3. Hu, Z.; Liu, L.; Wang, X.; Zheng, Q.; Han, C.; Li, W. Current Progress of Anode-Free Rechargeable Sodium Metal Batteries: Origin, Challenges, Strategies, and Perspectives. Adv. Funct. Mater. 2024, 34, 2313823. [Google Scholar] [CrossRef]
  4. Chen, Y.; Ye, C.; Zhang, N.; Liu, J.; Li, H.; Davey, K.; Qiao, S. Prospects for practical anode-free sodium batteries. Mater. Today 2024, 73, 260–274. [Google Scholar] [CrossRef]
  5. Yang, T.; Luo, D.; Liu, Y.; Aiping Yu, A.; Chen, Z. Anode-free sodium metal batteries as rising stars for lithium-ion alternatives. iScience 2023, 26, 105982. [Google Scholar] [CrossRef]
  6. Xu, P.; Huang, F.; Sun, Y.; Lei, Y.; Cao, X.; Liang, S.; Fang, G. Anode-Free Alkali Metal Batteries: From Laboratory to Practicability. Adv. Funct. Mater. 2024, 34, 2406080. [Google Scholar] [CrossRef]
  7. Tian, Y.; An, Y.; Wei, C.; Jiang, H.; Xiong, S.; Feng, J.; Qian, Y. Recently advances and perspectives of anode-free rechargeable batteries. Nano Energy 2020, 78, 105344. [Google Scholar] [CrossRef]
  8. Li, Y.; Zhou, Q.; Weng, S.; Ding, F.; Qi, X.; Lu, J.; Li, Y.; Zhang, X.; Rong, X.; Lu, Y.; et al. Interfacial engineering to achieve an energy density of over 200 Wh kg−1 in sodium batteries. Nat. Energy 2022, 7, 511–519. [Google Scholar] [CrossRef]
  9. Heubner, C.; Maletti, S.; Auer, H.; Hüttl, J.; Voigt, K.; Lohrberg, O.; Nikolowski, K.; Partsch, M.; Michaelis, A. From Lithium-Metal toward Anode-Free Solid-State Batteries: Current Developments, Issues, and Challenges. Adv. Funct. Mater. 2021, 31, 2106608. [Google Scholar] [CrossRef]
  10. Xie, Z.; Wu, Z.; An, X.; Yue, X.; Wang, J.; Abudula, A.; Guan, G. Anode-free rechargeable lithium metal batteries: Progress and prospects. Energy Storage Mater. 2020, 32, 386–401. [Google Scholar] [CrossRef]
  11. Tong, Z.; Bazri, B.; Hu, S.; Liu, R.S. Interfacial chemistry in anode-free batteries: Challenges and strategies. J. Mater. Chem. A 2021, 9, 7396–7406. [Google Scholar] [CrossRef]
  12. Zhao, C.; Liu, L.; Qi, X.; Lu, Y.; Wu, F.; Zhao, J.; Yu, Y.; Hu, Y.; Chen, L. Solid-State Sodium Batteries. Adv. Energy Mater. 2018, 8, 1703012. [Google Scholar] [CrossRef]
  13. Fan, L.; Wei, S.; Li, S.; Li, Q.; Lu, Y. Recent Progress of the Solid-State Electrolytes for High-Energy Metal-Based Batteries. Adv. Energy Mater. 2018, 8, 1702657. [Google Scholar] [CrossRef]
  14. Lu, G.; Li, M.; Deng, R.; Hou, W.; Lu, L.; Song, S.; Xu, C.; Wang, R. Ultralong lifespan solid-state sodium battery with a supersodiophilic and fast ionic conductive composite sodium anode. Energy Storage Mater. 2024, 72, 103755. [Google Scholar] [CrossRef]
  15. Janek, J.; Zeier, W.G. Challenges in speeding up solid-state battery development. Nat. Energy 2023, 8, 230–240. [Google Scholar] [CrossRef]
  16. Zheng, X.; Bommier, C.; Luo, W.; Jiang, L.; Hao, Y.; Huang, Y. Sodium Metal Anodes for Room-Temperature Sodium-Ion Batteries: Applications, Challenges and Solutions. Energy Storage Mater. 2019, 16, 6–23. [Google Scholar] [CrossRef]
  17. Zhao, Y.; Adair, K.R.; Sun, X. Recent developments and insights into the understanding of Na metal anodes for Na-metal batteries. Energy Environ. Sci. 2018, 11, 2673–2695. [Google Scholar] [CrossRef]
  18. Eckhardt, J.K.; Klar, P.J.; Janek, J.; Heiliger, C. Interplay of Dynamic Constriction and Interface Morphology between Reversible Metal Anode and Solid Electrolyte in Solid State Batteries. ACS Appl. Mater. Interfaces 2022, 14, 35545–35554. [Google Scholar] [CrossRef]
  19. Mazur, R.G.; Dickey, D.H. A Spreading Resistance Technique for Resistivity Measurements on Silicon. J. Electrochem. Soc. 1966, 113, 255–259. [Google Scholar] [CrossRef]
  20. Denhoff, M.W. An accurate calculation of spreading resistance. J. Phys. D Appl. Phys. 2006, 39, 1761–1765. [Google Scholar] [CrossRef]
  21. Zhang, P.; Lau, Y.Y.; Timsit, R.S. On the Spreading Resistance of Thin-Film Contacts. IEEE Trans. Electron Devices 2012, 59, 1936–1940. [Google Scholar] [CrossRef]
  22. Eckhardt, J.K.; Fuchs, T.; Burkhardt, S.; Klar, P.J.; Janek, J.; Heiliger, C. 3D Impedance Modeling of Metal Anodes in Solid-State Batteries–Incompatibility of Pore Formation and Constriction Effect in Physical-Based 1D Circuit Models. ACS Appl. Mater. Interfaces 2022, 14, 42757–42769. [Google Scholar] [CrossRef] [PubMed]
  23. Fleig, J.; Maier, J. Rough electrodes in solid and liquid electrochemistry: Impact of morphology on the impedance. Solid State Ionics 1997, 94, 199–207. [Google Scholar] [CrossRef]
  24. Jolly, D.S.; Ning, Z.; Darnbrough, J.E.; Kasemchainan, J.; Hartley, G.O.; Adamson, P.; Armstrong, D.E.J.; Marrow, J.; Bruce, P.G. Sodium/Na β″ Alumina Interface: Effect of Pressure on Voids. ACS Appl. Mater. Interfaces 2020, 12, 678–685. [Google Scholar] [CrossRef]
  25. Koshikawa, H.; Matsuda, S.; Kamiya, K.; Miyayama, M.; Kubo, Y.; Uosaki, K.; Hashimoto, K.; Nakanishi, S. Dynamic changes in charge-transfer resistance at Li metal/Li7La3Zr2O12 interfaces during electrochemical Li dissolution/deposition cycles. J. Power Sources 2018, 376, 147–151. [Google Scholar] [CrossRef]
  26. Bay, M.; Wang, M.; Grissa, R.; Heinz, M.V.F.; Sakamoto, J.; Battaglia, C. Sodium Plating from Na-β″-Alumina Ceramics at Room Temperature, Paving the Way for Fast-Charging All-Solid-State Batteries. Adv. Energy Mater. 2020, 10, 1902899. [Google Scholar] [CrossRef]
  27. Krauskopf, T.; Hartmann, H.; Zeier, W.G.; Janek, J. Toward a Fundamental Understanding of the Lithium Metal Anode in Solid-State Batteries—An Electrochemo-Mechanical Study on the Garnet-Type Solid Electrolyte Li6.25Al0.25La3Zr2O12. ACS Appl. Mater. Interfaces 2019, 11, 14463–14477. [Google Scholar] [CrossRef]
  28. Janek, J. Oscillating reactions at metal electrodes in solid electrolytes. Solid State Ion. 1997, 101–103, 721–727. [Google Scholar] [CrossRef]
  29. Rohnke, M.; Rosenkranz, C.; Janek, J. The influence of non-equilibrium defects on the anodic dissolution of a metal into a solid electrolyte. Solid State Ion. 2006, 177, 447–456. [Google Scholar] [CrossRef]
  30. Janek, J.; Majoni, S. Investigation of Charge Transport Across the Ag I AgI-interface: (I) Occurrence of Periodic Phenomena During Anodic Dissolution of Silver. Ber. Bunsenges. Phys. Chem. 1995, 99, 14–20. [Google Scholar] [CrossRef]
  31. Majoni, S.; Janek, J. Investigation of Charge Transport Across the Agl AgI-Interface: (II) Dilatometric Study of the Anodic Dissolution of Silver. Ber. Bunsenges. Phys. Chem. 1998, 102, 756–762. [Google Scholar] [CrossRef]
  32. Motoyama, M.; Ejiri, M.; Iriyama, Y. Modeling the Nucleation and Growth of Li at Metal Current Collector/LiPON Interfaces. J. Electrochem. Soc. 2015, 162, A7067–A7071. [Google Scholar] [CrossRef]
  33. Motoyama, M.; Ejiri, M.; Yamamoto, T.; Iriyama, Y. In Situ Scanning Electron Microscope Observations of Li Plating/Stripping Reactions with Pt Current Collectors on LiPON Electrolyte. J. Electrochem. Soc. 2018, 165, A1338–A1347. [Google Scholar] [CrossRef]
  34. Motoyama, M.; Hirota, M.; Yamamoto, T.; Iriyama, Y. Temperature Effects on Li Nucleation at Cu/LiPON Interfaces. ACS Appl. Mater. Interfaces 2020, 12, 38045–38053. [Google Scholar] [CrossRef]
  35. Kazyak, E.; Wang, M.J.; Lee, K.; Yadavalli, S.; Sanchez, A.J.; Thouless, M.D.; Sakamoto, J.; Dasgupta, N.P. Understanding the electro-chemo-mechanics of Li plating in anode-free solid-state batteries with operando 3D microscopy. Matter 2022, 5, 3912–3934. [Google Scholar] [CrossRef]
  36. Fuchs, T.; Ortmann, T.; Becker, J.; Haslam, C.G.; Ziegler, M.; Singh, V.K.; Rohnke, M.; Mogwitz, B.; Peppler, K.; Nazar, L.F.; et al. Imaging the microstructure of lithium and sodium metal in anode-free solid-state batteries using electron backscatter diffraction. Nat. Mater. 2024, 23, 1678–1685. [Google Scholar] [CrossRef]
  37. Ortmann, T.; Fuchs, T.; Eckhardt, J.K.; Ding, Z.; Ma, Q.; Tietz, F.; Kübel, C.; Rohnke, M.; Janek, J. Deposition of Sodium Metal at the Copper-NaSICON Interface for Reservoir-Free Solid-State Sodium Batteries. Adv. Energy Mater. 2024, 14, 2302729. [Google Scholar] [CrossRef]
  38. Ortmann, T.; Burkhardt, S.; Eckhardt, J.K.; Fuchs, T.; Ding, Z.; Sann, J.; Rohnke, M.; Ma, Q.; Tietz, F.; Fattakhova-Rohlfing, D.; et al. Kinetics and Pore Formation of the Sodium Metal Anode on NASICON-Type Na3.4Zr2Si2.4P0.6O12 for Sodium Solid-State Batteries. Adv. Energy Mater. 2023, 13, 2202712. [Google Scholar] [CrossRef]
  39. Gao, H.; Xin, S.; Xue, L.; Goodenough, J.B. Stabilizing a High-Energy-Density Rechargeable Sodium Battery with a Solid Electrolyte. Chem 2018, 4, 833–844. [Google Scholar] [CrossRef]
  40. Oh, J.A.S.; Wang, Y.; Zeng, Q.; Sun, J.; Sun, Q.; Goh, M.; Chua, B.; Zeng, K.; Lu, L. Intrinsic low sodium/NASICON interfacial resistance paving the way for room temperature sodium-metal battery. J. Colloid Interface Sci. 2021, 601, 418–426. [Google Scholar] [CrossRef]
  41. Quérel, E.; Seymour, I.D.; Cavallaro, A.; Ma, Q.; Tietz, F.; Aguadero, A. The role of NaSICON surface chemistry in stabilizing fast-charging Na metal solid-state batteries. J. Phys. Energy 2021, 3, 044007. [Google Scholar] [CrossRef]
  42. Quérel, E.; Williams, N.J.; Seymour, I.D.; Skinner, S.J.; Aguadero, A. Operando Characterization and Theoretical Modeling of Metal|Electrolyte Interphase Growth Kinetics in Solid-State Batteries. Part I: Experiments. Chem. Mater. 2023, 35, 853–862. [Google Scholar] [CrossRef] [PubMed]
  43. Lu, Y.; Li, L.; Zhang, Q.; Niu, Z.; Chen, J. Electrolyte and Interface Engineering for Solid-State Sodium Batteries. Joule 2018, 2, 1747–1770. [Google Scholar] [CrossRef]
  44. Tang, X.; Han, W.; Zhang, Y.; Liu, S. Interface engineering of sodium metal anode for all-solid-state sodium batteries. J. Power Sources 2024, 623, 235397. [Google Scholar] [CrossRef]
  45. Tsai, C.; Lan, T.; Dellen, C.; Ling, Y.; Ma, Q.; Fattakhova-Rohlfing, D.; Guillon, O.; Tietz, F. Dendrite-tolerant all-solid-state sodium batteries and an important mechanism of metal self-diffusion. J. Power Sources 2020, 476, 228666. [Google Scholar] [CrossRef]
  46. Li, R.; Jiang, D.; Du, P.; Yuan, C.; Cui, X.; Tang, Q.; Zheng, J.; Li, Y.; Lu, K.; Ren, X.; et al. Negating Na‖Na3Zr2Si2PO12 interfacial resistance for dendrite-free and “Na-less” solid-state batteries. Chem. Sci. 2022, 13, 14132–14140. [Google Scholar] [CrossRef]
  47. Pei, A.; Zheng, G.; Shi, F.; Li, Y.; Cui, Y. Nanoscale Nucleation and Growth of Electrodeposited Lithium Metal. Nano Lett. 2017, 17, 1132–1139. [Google Scholar] [CrossRef]
  48. Xu, P.; Li, X.; Yan, M.; Ni, H.; Huang, H.; Lin, X.; Liu, X.; Fan, J.; Zheng, M.; Yuan, R.; et al. A highly reversible sodium metal anode by mitigating electrodeposition overpotential. J. Mater. Chem. A 2021, 9, 22892–22900. [Google Scholar] [CrossRef]
  49. Willow, A.; Hussein, H.E.M.; Vajirakaphan, S.; Chasri, A.; Margadonna, S. Improving In-Situ Sodium Metal Plating on Copper Foil Through Optimization of Mechanical Pressure: Towards High-Performance Anode-Free Sodium Ion Batteries. Front. Energy Res. 2022, 10, 888321. [Google Scholar] [CrossRef]
  50. Sayahpour, B.; Li, W.; Bai, S.; Lu, B.; Han, B.; Chen, Y.; Deysher, G.; Parab, S.; Ridley, P.; Raghavendran, G.; et al. Quantitative analysis of sodium metal deposition and interphase in Na metal batteries. Energy Environ. Sci. 2024, 17, 1216–1228. [Google Scholar] [CrossRef]
  51. Louli, A.J.; Genovese, M.; Weber, R.; Hames, S.G.; Logan, E.R.; Dahn, J.R. Exploring the Impact of Mechanical Pressure on the Performance of Anode-Free Lithium Metal Cells. J. Electrochem. Soc. 2019, 166, A1291–A1299. [Google Scholar] [CrossRef]
  52. Zhang, X.; Wang, Q.J.; Harrison, K.L.; Roberts, S.A.; Harris, S.J. Pressure-Driven Interface Evolution in Solid-State Lithium Metal Batteries. Cell Rep. Phys. Sci. 2020, 1, 100012. [Google Scholar] [CrossRef]
  53. Wang, M.J.; Chang, J.; Wolfenstine, J.B.; Sakamoto, J. Analysis of elastic, plastic, and creep properties of sodium metal and implications for solid-state batteries. Materialia 2020, 12, 100792. [Google Scholar] [CrossRef]
  54. Sun, C.; Li, Y.; Sun, Z.; Yuan, X.; Jin, H.; Zhao, Y. Ferroelectric interface for efficient sodium metal cycling in anode-free solid-state batteries. Mater. Today 2024, 80, 395–405. [Google Scholar] [CrossRef]
  55. Ding, Z.; Tang, Y.; Ortmann, T.; Eckhardt, J.K.; Dai, Y.; Rohnke, M.; Melinte, G.; Heiliger, C.; Janek, J.; Kübel, C. The Impact of Microstructure on Filament Growth at the Sodium Metal Anode in All-Solid-State Sodium Batteries. Adv. Energy Mater. 2023, 13, 2302322. [Google Scholar] [CrossRef]
  56. Ding, Z.; Tang, Y.; Ortmann, T.; Eckhardt, J.K.; Rohnke, M.; Melinte, G.; Heiliger, C.; Janek, J.; Kübel, C. Microstructural Influence on Sodium Filament Growth in All Solid-state Batteries. BIO Web Conf. 2024, 129, 25020. [Google Scholar] [CrossRef]
  57. Lou, S.; Zhang, F.; Fu, C.; Chen, M.; Ma, Y.; Yin, G.; Wang, J. Interface Issues and Challenges in All-Solid-State Batteries: Lithium, Sodium, and Beyond. Adv. Mater. 2020, 32, 2000721. [Google Scholar] [CrossRef]
  58. Tian, Y.; Shi, T.; Richards, W.D.; Li, J.; Kim, J.C.; Bo, S.; Ceder, G. Compatibility issues between electrodes and electrolytes in solid-state batteries. Energy Environ. Sci. 2017, 10, 1150–1166. [Google Scholar] [CrossRef]
  59. Wang, S.; Xu, H.; Li, W.; Dolocan, A.; Manthiram, A. Interfacial Chemistry in Solid-state Batteries: Formation of Inter-phase and Its Consequences. J. Am. Chem. Soc. 2018, 140, 250–257. [Google Scholar] [CrossRef]
  60. Ma, Q.; Ortmann, T.; Yang, A.; Sebold, D.; Burkhardt, S.; Rohnke, M.; Tietz, F.; Fattakhova-Rohlfing, D.; Janek, J.; Guillon, O. Enhancing the Dendrite Tolerance of NaSICON Electrolytes by Suppressing Edge Growth of Na Electrode along Ceramic Surface. Adv. Energy Mater. 2022, 12, 2201680. [Google Scholar] [CrossRef]
  61. Wang, X.; Chen, J.; Mao, Z.; Wang, D. Effective resistance to dendrite growth of NASICON solid electrolyte with lower electronic conductivity. Chem. Eng. J. 2022, 427, 130899. [Google Scholar] [CrossRef]
  62. Zhang, Z.; Wenzel, S.; Zhu, Y.; Sann, J.; Shen, L.; Yang, J.; Yao, X.; Hu, Y.; Wolverton, C.; Li, H.; et al. Na3Zr2Si2PO12: A Stable Na+-Ion Solid Electrolyte for Solid-State Batteries. ACS Appl. Energy Mater. 2020, 3, 7427–7437. [Google Scholar] [CrossRef]
  63. Hu, X.; Wang, M.; Liu, Y.; Meng, X.; Zhong, Y.; Wang, X.; Tu, J. Fluorinated solid electrolyte interphase enables interfacial stability for sulfide-based solid-state sodium metal batteries. Chem. Eng. J. 2024, 499, 156290. [Google Scholar] [CrossRef]
  64. Gao, Z.; Yang, J.; Yuan, H.; Fu, H.; Li, Y.; Li, Y.; Ferber, T.; Guhl, C.; Sun, H.; Jaegermann, W.; et al. Stabilizing Na3Zr2Si2PO12/Na Interfacial Performance by Introducing a Clean and Na-Deficient Surface. Chem. Mater. 2020, 32, 3970–3979. [Google Scholar] [CrossRef]
  65. Li, Y.; Sun, C.; Sun, Z.; Li, M.; Jin, H.; Zhao, Y. Boosting Na-O Affinity in Na3Zr2Si2PO12 Electrolyte Promises Highly Rechargeable Solid-State Sodium Batteries. Adv. Funct. Mater. 2024, 34, 2403937. [Google Scholar] [CrossRef]
  66. Wang, C.; Jin, H.; Zhao, Y. Surface Potential Regulation Realizing Stable Sodium/Na3Zr2Si2PO12 Interface for Room-Temperature Sodium Metal Batteries. Small 2021, 17, 2100974. [Google Scholar] [CrossRef]
  67. Zhou, W.; Li, Y.; Xin, S.; Goodenough, J.B. Rechargeable Sodium All-Solid-State Battery. ACS Cent. Sci. 2017, 3, 52–57. [Google Scholar] [CrossRef]
  68. Lu, Y.; Alonso, J.A.; Yi, Q.; Lu, L.; Wang, Z.L.; Sun, C. A High-Performance Monolithic Solid-State Sodium Battery with Ca2+ Doped Na3Zr2Si2PO12 Electrolyte. Adv. Energy Mater. 2019, 9, 1901205. [Google Scholar] [CrossRef]
  69. Tian, H.; Liu, S.; Deng, L.; Wang, L.; Dai, L. New-type Hf-based NASICON electrolyte for solid-state Na-ion batteries with superior long-cycling stability and rate capability. Energy Storage Mater. 2021, 39, 232–238. [Google Scholar] [CrossRef]
  70. Suo, J.; Zhao, Q.; Tian, H.; Wang, L.; Dai, L.; Luo, J.; Liu, S. Designing a quasi-liquid alloy interface for solid Na-ion battery. ACS Nano 2023, 17, 10229–10235. [Google Scholar] [CrossRef]
  71. Miao, X.; Di, H.; Ge, X.; Zhao, D.; Wang, P.; Wang, R.; Wang, C.; Yin, L. AlF3-modified anode-electrolyte interface for effective Na dendrites restriction in NASICON-based solid-state electrolyte. Energy Storage Mater. 2020, 30, 170–178. [Google Scholar] [CrossRef]
  72. Yu, X.; Manthiram, A. Sodium-Sulfur Batteries with a Polymer-Coated NASICON-type Sodium-Ion Solid Electrolyte. Matter 2019, 1, 439–451. [Google Scholar] [CrossRef]
  73. Miao, X.; Wang, H.; Sun, R.; Ge, X.; Zhao, D.; Peng Wang, P.; Wang, R.; Yin, L. Isotropous Sulfurized Polyacrylonitrile Interlayer with Homogeneous Na+ Flux Dynamics for Solid-State Na Metal Batteries. Adv. Energy Mater. 2021, 11, 2003469. [Google Scholar] [CrossRef]
  74. Chi, X.; Hao, F.; Zhang, J.; Wu, X.; Zhang, Y.; Gheytani, S.; Wen, Z.; Yao, Y. A high-energy quinone-based all-solid-state sodium metal battery. Nano Energy 2019, 62, 718–724. [Google Scholar] [CrossRef]
  75. Yang, J.; Gao, Z.; Ferber, T.; Zhang, H.; Guhl, C.; Yang, L.; Li, Y.; Deng, Z.; Liu, P.; Cheng, C.; et al. Guided-formation of favorable interface for stabilizing Na metal solid-state batteries. J. Mater. Chem. A 2020, 8, 7828–7835. [Google Scholar] [CrossRef]
  76. Wu, T.; Wen, Z.; Sun, C.; Wu, X.; Zhang, S.; Yang, J. Disordered Carbon Tubes Based on Cotton Cloth for Modulating the Interface Impedance in β”-Al2O3-based Solid-state Sodium Metal Batteries. J. Mater. Chem. A 2018, 6, 12623–12629. [Google Scholar] [CrossRef]
  77. Matios, E.; Wang, H.; Wang, C.; Hu, X.; Lu, X.; Luo, J.; Li, W. Graphene Regulated Ceramic Electrolyte for Solid-State Sodium Metal Battery with Superior Electrochemical Stability. ACS Appl. Mater. Interfaces 2019, 11, 5064–5072. [Google Scholar] [CrossRef] [PubMed]
  78. Fang, D.; Li, Y.; Wang, C.; Miao, R.; Yang, S.; Zhao, Y.; Ding, Y.; He, J.; Chen, L.; Li, N.; et al. Wide-temperature solid-state sodium metal batteries using Na+ superionic conductor-type solid electrolytes. Energy Storage Mater. 2025, 74, 103973. [Google Scholar] [CrossRef]
  79. Jaschin, P.W.; Tang, C.R.; Eric, D.; Wachsman, E.D. High-rate cycling in 3D dual-doped NASICON architectures toward room-temperature sodium-metal-anode solid-state batteries. Energy Environ. Sci. 2024, 17, 727–737. [Google Scholar] [CrossRef]
  80. Wang, X.; Chen, J.; Wang, D.; Mao, Z. Improving the alkali metal electrode/inorganic solid electrolyte contact via room-temperature ultrasound solid welding. Nat. Commun. 2021, 12, 7109. [Google Scholar] [CrossRef]
  81. Uchida, Y.; Hasegawa, G.; Shima, K.; Inada, M.; Enomoto, N.; Akamatsu, H.; Hayashi, K. Insights into Sodium Ion Transfer at the Na/NASICON Interface Improved by Uniaxial Compression. ACS Appl. Energy Mater. 2019, 2, 2913–2920. [Google Scholar] [CrossRef]
  82. Chowdhury, S.; Mahato, J.C.; Chung, J.S.; Kang, S.G.; Gupta, B.C. Theoretical prediction of the potential of using two-dimensional V2S2 as an effective anode for alkali metal ion batteries. Surf. Interfaces 2024, 51, 104750. [Google Scholar] [CrossRef]
  83. Dey, T.; Chowdhury, S.; Kang, S.G.; Sen, P.; Gupta, B.C.; Mahato, J.C. Two-dimensional ScTe2 monolayer: An efficient anode material for sodium-ion battery and cathode material for lithium-ion and potassium-ion battery. Comput. Mater. Sci. 2025, 253, 113824. [Google Scholar] [CrossRef]
  84. Gao, H.; Xue, L.; Xin, S.; Park, K.; Goodenough, J.B. A Plastic–Crystal Electrolyte Interphase for All-Solid-State Sodium Batteries. Angew. Chem. Int. Ed. 2017, 56, 5541–5545. [Google Scholar] [CrossRef]
  85. Zhang, Z.; Zhang, Q.; Shi, J.; Chu, Y.S.; Yu, X.; Xu, K.; Ge, M.; Yan, H.; Li, W.; Gu, L.; et al. A Self-Forming Composite Electrolyte for Solid-State Sodium Battery with Ultralong Cycle Life. Adv. Energy Mater. 2016, 6, 1601196. [Google Scholar] [CrossRef]
  86. Wang, Q.; Yu, C.; Li, L.; Liu, X.; Zhang, X.; Gao, G.; Wang, Y.; Li, G. Sc-doping in Na3Zr2Si2PO12 electrolytes enables preeminent performance of solid-state sodium batteries in a wide temperature range. Energy Storage Mater. 2023, 54, 135–145. [Google Scholar] [CrossRef]
  87. Zatovsky, I.V. NASICON-type Na3V2(PO4)3. Acta Cryst. 2010, 66, i12. [Google Scholar] [CrossRef]
  88. Zhang, Z.; Zhang, Q.; Ren, C.; Luo, F.; Ma, Q.; Hu, Y.; Zhou, Z.; Li, H.; Huang, X.; Chen, L. A ceramic/polymer composite solid electrolyte for sodium batteries. J. Mater. Chem. A 2016, 4, 15823–15828. [Google Scholar] [CrossRef]
  89. Pandit, B.; Johansen, M.; Martínez-Cisneros, C.S.; Naranjo-Balseca, J.M.; Levenfeld, B.; Ravnsbæk, D.B.; Varez, A. Na3V2(PO4)3 Cathode for Room-Temperature Solid-State Sodium-Ion Batteries: Advanced In Situ Synchrotron X-ray Studies to Understand Intermediate Phase Evolution. Chem. Mater. 2024, 36, 2314–2324. [Google Scholar] [CrossRef]
  90. Kehne, P.; Guhl, C.; Mac, Q.; Tietz, F.; Alff, L.; Hausbrand, R.; Komissinskiy, P. Sc-substituted Nasicon solid electrolyte for an all-solid-state NaxCoO2/Nasicon/Na sodium model battery with stable electrochemical performance. J. Power Sources 2019, 409, 86–93. [Google Scholar] [CrossRef]
  91. Li, L.; Yao, J.; Xu, R.; Lin, Q.; Yan, X.; Yu, C.; Zhang, L. Highly stable and encapsulation-microstructural cathode derived by self-pressurization behavior in Na-halides-based all-solid-state batteries. Energy Storage Mater. 2023, 63, 103016. [Google Scholar] [CrossRef]
  92. Sun, C.; Xie, L.; Wang, B.; Tan, H.; Zhang, Q.; Han, L.; Liu, Y.; Sydorov, D. Improve the cycling performance of solid-state sodium batteries through the design of a cathode interface buffering layer. Appl. Surf. Sci. 2025, 680, 161265. [Google Scholar] [CrossRef]
  93. Liu, L.; Qi, X.; Ma, Q.; Rong, X.; Hu, Y.; Zhou, Z.; Li, H.; Huang, X.; Chen, L. Toothpaste-like Electrode: A Novel Approach to Optimize the Interface for Solid-State Sodium-Ion Batteries with Ultralong Cycle Life. ACS Appl. Mater. Interfaces 2016, 8, 32631–32636. [Google Scholar] [CrossRef] [PubMed]
  94. Liang, B.; Lv, Y.; Wang, H.; Li, B.; Zhao, S.; Zheng, L.; Huang, Y.; Hong, Z. Durable high voltage solid-state sodium batteries with Pseudocapacitive P2 layered oxide cathode. Energy Storage Mater. 2024, 73, 103804. [Google Scholar] [CrossRef]
  95. Yang, S.; Shao, C.; Xiao, X.; Fang, D.; Li, N.; Zhao, E.; Wang, C.; Chen, L.; Li, N.; Li, J.; et al. Revealing stable organic cathode/solid electrolyte interface to promote all-solid-state sodium batteries using organic cathodes. Energy Storage Mater. 2024, 73, 103857. [Google Scholar] [CrossRef]
  96. Du, K.; Guo, H.; Hu, G.; Peng, Z.; Cao, Y. Na3V2(PO4)3 as cathode material for hybrid lithium ion batteries. J. Power Sources 2013, 223, 284–288. [Google Scholar] [CrossRef]
  97. Zhu, X.; Zhao, R.; Deng, W.; Ai, X.; Yang, H.; Cao, Y. An All-solid-state and All-organic Sodium-ion Battery based on Redox-active Polymers and Plastic Crystal Electrolyte. Electrochim. Acta 2015, 178, 55–59. [Google Scholar] [CrossRef]
  98. Ma, Q.; Tietz, F. Solid-State Electrolyte Materials for Sodium Batteries: Towards Practical Applications. ChemElectroChem 2020, 7, 2693–2713. [Google Scholar] [CrossRef]
  99. Åvall, G.; Mindemark, J.; Brandell, D.; Johansson, P. Sodium-Ion Battery Electrolytes: Modeling and Simulations. Adv. Energy Mater. 2018, 8, 1703036. [Google Scholar] [CrossRef]
  100. Kim, J.; Yoon, K.; Park, I.; Kang, K. Progress in the Development of Sodium-Ion Solid Electrolytes. Small Methods 2017, 1, 1700219. [Google Scholar] [CrossRef]
  101. Zhang, Z.; Shao, Y.; Lotsch, B.; Hu, Y.; Li, H.; Janek, J.; Nazar, L.F.; Nan, C.; Maier, J.; Armand, M.; et al. New Horizons for Inorganic Solid State Ion Conductors. Energy Environ. Sci. 2018, 11, 1945–1976. [Google Scholar] [CrossRef]
  102. Zhao, Q.; Stalin, S.; Zhao, C.; Archer, L.A. Designing solid-state electrolytes for safe, energy-dense batteries. Nat. Rev. Mater. 2020, 5, 229–252. [Google Scholar] [CrossRef]
  103. Ahmad, H.; Kubra, K.T.; Butt, A.; Nisar, U.; Iftikhar, F.J.; Ali, G. Recent progress, challenges, and perspectives in the development of solid-state electrolytes for sodium batteries. J. Power Sources 2023, 581, 233518. [Google Scholar] [CrossRef]
  104. Hou, W.; Guo, X.; Shen, X.; Amine, K.; Yu, H.; Lu, J. Solid electrolytes and interfaces in all-solid-state sodium batteries: Progress and perspective. Nano Energy 2018, 52, 279–291. [Google Scholar] [CrossRef]
  105. Kumar, P.P.; Yashonath, S. Ionic conduction in the solid state. J. Chem. Sci. 2006, 118, 135–154. [Google Scholar] [CrossRef]
  106. Guin, M.; Tietz, F. Survey of the transport properties of sodium superionic conductor materials for use in sodium batteries. J. Power Sources 2015, 273, 1056–1064. [Google Scholar] [CrossRef]
  107. Wu, J.; Zhang, R.; Fu, Q.; Zhang, J.; Zhou, X.; Gao, P.; Xu, C.; Liu, J.; Guo, X. Inorganic Solid Electrolytes for All-Solid-State Sodium Batteries: Fundamentals and Strategies for Battery Optimization. Adv. Funct. Mater. 2021, 31, 2008165. [Google Scholar] [CrossRef]
  108. Yang, Z.; Tang, B.; Ren, D.; Yu, X.; Gao, Y.; Wu, Y.; Yang, Y.; Chen, Z.; Zhou, Z. Advancing solid-state sodium batteries: Status quo of sulfide-based solid electrolytes. Mater. Today 2024, 80, 429–449. [Google Scholar] [CrossRef]
  109. Goodenough, J.B.; Hong, H.Y.; Kafalas, J.A. Fast Na+-ion transport in skeleton structures. Mater. Res. Bull. 1976, 11, 203–220. [Google Scholar] [CrossRef]
  110. Hong, H.Y. Crystal structures and crystal chemistry in the system Na1+xZr2SixP3−xO12. Mater. Res. Bull. 1976, 11, 173–182. [Google Scholar] [CrossRef]
  111. Boilot, J.P.; Collin, G.; Colomban, P. Relation structure-fast ion conduction in the NASICON solid solution. J. Solid State Chem. 1998, 73, 160–171. [Google Scholar] [CrossRef]
  112. Park, H.; Jung, K.; Nezafati, M.; Kim, C.; Kang, B. Sodium Ion Diffusion in Nasicon (Na3Zr2Si2PO12) Solid Electrolytes: Effects of Excess Sodium. ACS Appl. Mater. Interfaces 2016, 8, 27814–27824. [Google Scholar] [CrossRef]
  113. Zhang, Z.; Zou, Z.; Kaup, K.; Xiao, R.; Shi, S.; Avdeev, M.; Hu, Y.; Wang, D.; He, B.; Li, H.; et al. Correlated Migration Invokes Higher Na+-Ion Conductivity in NaSICON-Type Solid Electrolytes. Adv. Energy Mater. 2019, 9, 1902373. [Google Scholar] [CrossRef]
  114. Ma, Q.; Tsai, C.; Wei, X.; Heggen, M.; Tietz, F.; Irvine, J.T.S. Room temperature demonstration of a sodium superionic conductor with grain conductivity in excess of 0.01 S cm-1 and its primary applications in symmetric battery cells. J. Mater. Chem. A 2019, 7, 7766–7776. [Google Scholar] [CrossRef]
  115. Fuentes, R.O.; Figueiredo, F.; Marques, F.M.B.; Franco, J.I. Optimised NASICON Ceramics for Na+ Sensing. Ionics 2002, 8, 383–390. [Google Scholar] [CrossRef]
  116. He, J.; Yang, S.; Xiao, X.; Fang, D.; Miao, R.; Wang, C.; Chen, L.; Li, N.; Li, J.; Su, Y.; et al. Aliovalent cation substitution in Na3Zr2Si2PO12 for practical solid-state sodium metal batteries. Energy Storage Mater. 2025, 75, 104037. [Google Scholar] [CrossRef]
  117. Takahashi, T.; Kuwabara, K.; Shibata, M. Solid-state Ionics—Conductivities of Na+ ion conductors based on NASICON. Solid State Ion. 1980, 1, 163–175. [Google Scholar] [CrossRef]
  118. Bohnke, O.; Ronchetti, S.; Mazza, D. Conductivity measurements on nasicon and nasicon-modified materials. Solid State Ion. 1999, 122, 127–136. [Google Scholar] [CrossRef]
  119. Anantharamulu, N.; Rao, K.K.; Rambabu, G.; Kumar, B.V.; Radha, V.; Vithal, M. A wide-ranging review on Nasicon type materials. J. Mater. Sci. 2011, 46, 2821–2837. [Google Scholar] [CrossRef]
  120. Chen, D.; Luo, F.; Zhou, W.; Zhu, D. Influence of Nb5+, Ti4+, Y3+ and Zn2+ doped Na3Zr2Si2PO12 solid electrolyte on its conductivity. J. Alloys Compd. 2018, 757, 348–355. [Google Scholar] [CrossRef]
  121. Yang, J.; Liu, G.; Avdeev, M.; Wan, H.; Han, F.; Shen, L.; Zou, Z.; Shi, S.; Hu, Y.; Wang, C.; et al. Ultrastable All-Solid-State Sodium Rechargeable Batteries. ACS Energy Lett. 2020, 5, 2835–2841. [Google Scholar] [CrossRef]
  122. Vogel, E.M.; Cava, R.J.; Rietman, E. Na+ ion conductivity and crystallographic cell characterization in the Hf-nasicon system Na1+xHf2SixP3−xO12. Solid State Ion. 1984, 14, 1–6. [Google Scholar] [CrossRef]
  123. Bogusz, W.; Krok, F.; Jakubowski, W. Influence of doping on some physical properties of NASICON. Solid State Ion. 1983, 9–10, 803–807. [Google Scholar] [CrossRef]
  124. Krok, F.; Kony, D.; Dygas, J.R.; Jakubowski, W.; Bogusz, W. On some properties of NASICON doped with MgO and CoO. Solid State Ion. 1989, 36, 251–254. [Google Scholar] [CrossRef]
  125. Ma, Q.; Guin, M.; Naqash, S.; Tsai, C.; Tietz, F.; Guillon, O. Scandium-Substituted Na3Zr2(SiO4)2(PO4) Prepared by a Solution-Assisted Solid-State Reaction Method as Sodium-Ion Conductors. Chem. Mater. 2016, 28, 4821–4828. [Google Scholar] [CrossRef]
  126. Yang, J.; Wan, H.; Zhang, Z.; Liu, G.; Xu, X.; Hu, Y.; Yao, X. NASICON-structured Na3.1Zr1.95Mg0.05Si2PO12 solid electrolyte for solid-state sodium batteries. Rare Met. 2018, 37, 480–487. [Google Scholar] [CrossRef]
  127. Oh, J.A.S.; He, L.; Plewa, A.; Morita, M.; Zhao, Y.; Sakamoto, T.; Song, X.; Zhai, W.; Zeng, K.; Lu, L. Composite NASICON (Na3Zr2Si2PO12) Solid-State Electrolyte with Enhanced Na+ Ionic Conductivity: Effect of Liquid Phase Sintering. ACS Appl. Mater. Interfaces 2019, 11, 40125–40133. [Google Scholar] [CrossRef]
  128. Nafis, M.S.; Liang, Z.; Lee, S.; Ban, C. Structural engineering developments in sulfide solid-state electrolytes for lithium and sodium solid-state batteries. Nano Energy 2025, 133, 110447. [Google Scholar] [CrossRef]
  129. Hayashi, A.; Noi, K.; Tanibata, N.; Nagao, M.; Tatsumisago, M. High sodium ion conductivity of glasseceramic electrolytes with cubic Na3PS4. J. Power Sources 2014, 258, 420–423. [Google Scholar] [CrossRef]
  130. Zhang, L.; Zhang, D.; Yang, K.; Yan, X.; Wang, L.; Mi, J.; Xu, B.; Li, Y. Vacancy-Contained Tetragonal Na3SbS4 Superionic Conductor. Adv. Sci. 2016, 3, 1600089. [Google Scholar] [CrossRef]
  131. Zhang, D.; Cao, X.; Xu, D.; Wang, N.; Yu, C.; Hu, W.; Yan, X.; Mi, J.; Wen, B.; Wang, L.; et al. Synthesis of cubic Na3SbS4 solid electrolyte with enhanced ion transport for all-solid-state sodium-ion batteries. Electrochim. Acta 2018, 259, 100–109. [Google Scholar] [CrossRef]
  132. Zhang, Z.; Ramos, E.; Lalère, F.; Assoud, J.; Kaup, K.; Hartman, P.; Nazar, L.F. Na11Sn2PS12: A New Solid State Sodium Superionic Conductor. Energy Environ. Sci. 2018, 11, 87–93. [Google Scholar] [CrossRef]
  133. Hayashi, A.; Noi, K.; Sakuda, A.; Tatsumisago, M. Superionic glass-ceramic electrolytes for room-temperature rechargeable sodium batteries. Nat. Commun. 2012, 3, 856. [Google Scholar] [CrossRef]
  134. Banerjee, A.; Park, K.H.; Heo, J.W.; Nam, Y.J.; Moon, C.K.; Oh, S.M.; Seung-Tae Hong, S.; Jung, Y.S. Na3SbS4: A Solution Processable Sodium Superionic Conductor for All-Solid-State Sodium-Ion Batteries. Angew. Chem. Int. Ed. 2016, 55, 9634–9638. [Google Scholar] [CrossRef]
  135. Noi, K.; Hayashi, A.; Tatsumisago, M. Structure and properties of the Na2S–P2S5 glasses and glass—ceramics prepared by mechanical milling. J. Power Sources 2014, 269, 260–265. [Google Scholar] [CrossRef]
  136. Wang, H.; Chen, Y.; Hood, Z.D.; Sahu, G.; Pandian, A.S.; Keum, J.K.; An, K.; Liang, C. An Air-Stable Na3SbS4 Superionic Conductor Prepared by a Rapid and Economic Synthetic Procedure. Angew. Chem. Int. Ed. 2016, 55, 8551–8555. [Google Scholar] [CrossRef] [PubMed]
  137. Kim, S.K.; Mao, A.; Sen, S.; Kim, S. Fast Na-Ion Conduction in a Chalcogenide Glass–Ceramic in the Ternary System Na2Se–Ga2Se3–GeSe2. Chem. Mater. 2014, 26, 5695–5699. [Google Scholar] [CrossRef]
  138. Jansen, M.; Henseler, U. Synthesis, structure determination, and ionic conductivity of sodium tetrathiophosphate. J. Solid State Chem. 1992, 99, 110–119. [Google Scholar] [CrossRef]
  139. Tatsumisago, M.; Hayashi, A. Sulfide Glass-Ceramic Electrolytes for All-Solid-State Lithium and Sodium Batteries. Int. J. Appl. Glass Sci. 2014, 5, 226–235. [Google Scholar] [CrossRef]
  140. Chu, I.; Kompella, C.S.; Nguyen, H.; Zhu, Z.; Hy, S.; Deng, Z.; Meng, Y.S.; Ong, S.P. Room-Temperature All-solid-state Rechargeable Sodium-ion Batteries with a Cl-doped Na3PS4 Superionic Conductor. Sci. Rep. 2016, 6, 33733. [Google Scholar] [CrossRef]
  141. de Klerk, N.J.J.; Wagemaker, M. Diffusion Mechanism of the Sodium-Ion Solid Electrolyte Na3PS4 and Potential Improvements of Halogen Doping. Chem. Mater. 2016, 28, 3122–3130. [Google Scholar] [CrossRef]
  142. Bo, S.; Wang, Y.; Kim, J.C.; Richards, W.D.; Ceder, G. Computational and Experimental Investigations of Na-Ion Conduction in Cubic Na3PSe4. Chem. Mater. 2016, 28, 252–258. [Google Scholar] [CrossRef]
  143. Tanibata, N.; Noi, K.; Hayashi, A.; Tatsumisago, M. Preparation and characterization of highly sodium ion conducting Na3PS4–Na4SiS4 solid electrolytes. RSC Adv. 2014, 4, 17120–17123. [Google Scholar] [CrossRef]
  144. Zhang, L.; Yang, K.; Mi, J.; Lu, L.; Zhao, L.; Wang, L.; Li, Y.; Zeng, H. Na3PSe4; A Novel Chalcogenide Solid Electrolyte with High Ionic Conductivity. Adv. Energy Mater. 2015, 5, 1501294. [Google Scholar] [CrossRef]
  145. Zhu, Z.; Chu, I.; Deng, Z.; Ong, S.P. Role of Na+ Interstitials and Dopants in Enhancing the Na+ Conductivity of the Cubic Na3PS4 Superionic Conductor. Chem. Mater. 2015, 27, 8318–8325. [Google Scholar] [CrossRef]
  146. Yu, L.; Jiao, Q.; Liang, B.; Shan, H.; Lin, C.; Gao, C.; Shen, X.; Dai, S. Exceptionally high sodium ion conductivity and enhanced air stability in Na3SbS4 via germanium doping. J. Alloys and Compd. 2022, 913, 165229. [Google Scholar] [CrossRef]
  147. Yu, Z.; Shang, S.; Seo, J.; Wang, D.; Luo, X.; Huang, Q.; Chen, S.; Lu, J.; Li, X.; Liu, Z.; et al. Exceptionally High Ionic Conductivity in Na3P0.62As0.38S4 with Improved Moisture Stability for Solid-State Sodium-Ion Batteries. Adv. Mater. 2017, 29, 1605. [Google Scholar] [CrossRef]
  148. Bo, S.; Wang, Y.; Ceder, G. Structural and Na-ion conduction characteristics of Na3PSxSe4−x. J. Mater. Chem. A 2016, 4, 9044–9053. [Google Scholar] [CrossRef]
  149. Takayanagi, T.; Nasu, A.; Tsuji, F.; Sakuda, A.; Tatsumisago, M.; Hayashi, A. Mechanochemically Prepared Highly Conductive Na2.88Sb0.88W0.12S4–NaI Composite Electrolytes for All-Solid-State Sodium Battery. Electrochemistry 2022, 90, 047005. [Google Scholar] [CrossRef]
  150. Takayanagi, T.; Nasu, A.; Tsuji, F.; Motohashi, K.; Sakuda, A.; Tatsumisago, M.; Hayashi, A. Preparation and characterization of Na2.88Sb0.88W0.12S4–xOx solid electrolyte. J. Ceram. Soc. Jpn. 2022, 130, 498–503. [Google Scholar] [CrossRef]
  151. Hayashi, A.; Masuzawa, N.; Yubuchi, S.; Tsuji, F.; Hotehama, C.; Sakuda, A.; Tatsumisago, M. A sodium-ion sulfide solid electrolyte with unprecedented conductivity at room temperature. Nat. Commun. 2019, 10, 5266. [Google Scholar] [CrossRef]
  152. Hibi, Y.; Tanibata, N.; Hayashi, A.; Tatsumisago, M. Preparation of sodium ion conducting Na3PS4–NaI glasses by a mechanochemical technique. Solid State Ion. 2015, 270, 6–9. [Google Scholar] [CrossRef]
  153. Tsuji, F.; Nasu, A.; Sakuda, A.; Tatsumisago, M.; Hayashi, A. Mechanochemical synthesis and characterization of Na3–xP1–xWxS4 solid electrolytes. J. Power Sources 2021, 506, 230100. [Google Scholar] [CrossRef]
  154. Fuchs, T.; Culver, S.P.; Till, P.; Zeier, W.G. Defect-Mediated Conductivity Enhancements in Na3−xPn1−xWxS4 (Pn = P, Sb) Using Aliovalent Substitutions. ACS Energy Lett. 2020, 5, 146–151. [Google Scholar] [CrossRef]
  155. Oh, K.; Chang, D.; Park, I.; Yoon, K.; Kang, K. First-Principles Investigations on Sodium Superionic Conductor Na11Sn2PS12. Chem. Mater. 2019, 31, 6066–6075. [Google Scholar] [CrossRef]
  156. Richards, W.R.; Tsujimura, T.; Miara, L.J.; Wang, Y.; Kim, J.C.; Ong, S.P.; Uechi, I.; Suzuki, N.; Ceder, G. Design and synthesis of the superionic conductor Na10SnP2S12. Nat. Commun. 2016, 7, 11009. [Google Scholar] [CrossRef] [PubMed]
  157. Feng, X.; Chien, P.; Zhu, Z.; Chu, I.; Wang, P.; Immediato-Scuotto, M.; Arabzadeh, H.; Ong, S.P.; Hu, Y. Studies of Functional Defects for Fast Na-Ion Conduction in Na3−yPS4−xClx with a Combined Experimental and Computational Approach. Adv. Funct. Mater. 2019, 29, 1807951. [Google Scholar] [CrossRef]
  158. Yu, L.; Yin, J.; Gao, C.; Lin, C.; Shen, X.; Dai, S.; Jiao, Q. Halogen Doping Mechanism and Interface Strengthening in the Na3SbS4 Electrolyte via Solid-State Synthesis. ACS Appl. Mater. Interfaces 2023, 15, 31635–31642. [Google Scholar] [CrossRef]
  159. Shu, L.; Gao, C.; Liu, Y.; Zhou, X.; Ma, H.; Zhang, X.; Shen, X.; Dai, S.; Lin, C.; Jiao, Q. Enhancing interface stability and ionic conductivity in the designed Na3SbP0.4xS4−xOx sulfide solid electrolyte through bridging oxygen. J. Colloid Interface Sci. 2023, 652, 2042–2053. [Google Scholar] [CrossRef]
  160. Wang, Y.; Richards, W.D.; Bo, S.; Miara, L.J.; Ceder, G. Computational Prediction and Evaluation of Solid-State Sodium Superionic Conductors Na7P3X11 (X = O, S, Se). Chem. Mater. 2017, 29, 7475–7482. [Google Scholar] [CrossRef]
  161. Matsuo, M.; Kuromoto, S.; Sato, T.; Oguchi, H.; Takamura, H.; Orimo, S. Sodium ionic conduction in complex hydrides with [BH4] and [NH2] anions. Appl. Phys. Lett. 2012, 100, 203904. [Google Scholar] [CrossRef]
  162. Udovic, T.J.; Motoaki Matsuo, M.; Unemoto, A.; Verdal, N.; Stavila, V.; Skripov, A.V.; Rush, J.J.; Takamura, H.; Orimo, S. Sodium superionic conduction in Na2B12H12. Chem. Commun. 2014, 50, 3750–3752. [Google Scholar] [CrossRef]
  163. Zhou, L.; Bazak, J.D.; Singh, B.; Li, C.; Assoud, A.; Washton, N.M.; Murugesan, V.; Nazar, L.F. A New Sodium Thioborate Fast Ion Conductor: Na3B5S9. Angew. Chem. Int. Ed. 2023, 62, e202300404. [Google Scholar] [CrossRef]
  164. Fu, C.; Li, Y.; Xu, W.; Feng, X.; Gu, W.; Liu, J.; Deng, W.; Wang, W.; Abeykoon, A.M.M.; Su, L.; et al. LaCl3-based sodium halide solid electrolytes with high ionic conductivity for all-solid-state batteries. Nat. Commun. 2024, 15, 4315. [Google Scholar] [CrossRef]
Figure 1. Schematic of the contact of the current collector (electrode) with a solid electrolyte. The contact surface on which charge transfer is possible is marked in red (from [18] open access).
Figure 1. Schematic of the contact of the current collector (electrode) with a solid electrolyte. The contact surface on which charge transfer is possible is marked in red (from [18] open access).
Batteries 11 00292 g001
Figure 2. Oscillations of the open-circuit potential at the copper/LLZO interface under different compression forces (from [35] open access).
Figure 2. Oscillations of the open-circuit potential at the copper/LLZO interface under different compression forces (from [35] open access).
Batteries 11 00292 g002
Figure 3. Dependence of nucleation overpotential on current density at a constant stack pressure of 2 MPa (a) and on stack pressure at a constant current density of 0.2 A/cm2 (b) (constructed according to data from [37]).
Figure 3. Dependence of nucleation overpotential on current density at a constant stack pressure of 2 MPa (a) and on stack pressure at a constant current density of 0.2 A/cm2 (b) (constructed according to data from [37]).
Batteries 11 00292 g003
Figure 4. Schematic of the in situ transmission electron microscopy experiment used to study the deposition of sodium whiskers at grain boundaries (from [55] open access).
Figure 4. Schematic of the in situ transmission electron microscopy experiment used to study the deposition of sodium whiskers at grain boundaries (from [55] open access).
Batteries 11 00292 g004
Figure 5. Distribution of SEI composition by depth for a liquid electrolyte (a,b) and solid electrolyte (c,d) (from [39] open access).
Figure 5. Distribution of SEI composition by depth for a liquid electrolyte (a,b) and solid electrolyte (c,d) (from [39] open access).
Batteries 11 00292 g005
Figure 6. Molecular structure of the copolymerization product of 5,5′,6,6′-tetrahydroxy-3,3,3′,3′-tetramethyl-1,1′-spirobisindane and tetrafluoroterephthalonitrile (from [72] open access).
Figure 6. Molecular structure of the copolymerization product of 5,5′,6,6′-tetrahydroxy-3,3,3′,3′-tetramethyl-1,1′-spirobisindane and tetrafluoroterephthalonitrile (from [72] open access).
Batteries 11 00292 g006
Figure 7. Schematic illustration of the ultrasound solid welding method (from [80] open access).
Figure 7. Schematic illustration of the ultrasound solid welding method (from [80] open access).
Batteries 11 00292 g007
Figure 8. Dependence of interfacial resistance (R) at the NASICON/Na interface on the compression pressure (P) (constructed according to data from [81]).
Figure 8. Dependence of interfacial resistance (R) at the NASICON/Na interface on the compression pressure (P) (constructed according to data from [81]).
Batteries 11 00292 g008
Figure 9. Contact of the cathode material with a solid electrolyte without (a) and with (b) wetting with a liquid electrolyte (from [85] with permission of Wiley).
Figure 9. Contact of the cathode material with a solid electrolyte without (a) and with (b) wetting with a liquid electrolyte (from [85] with permission of Wiley).
Batteries 11 00292 g009
Figure 10. Structural formula of ionic liquid PY14FSI.
Figure 10. Structural formula of ionic liquid PY14FSI.
Batteries 11 00292 g010
Figure 11. The crystal structure of Na3SbS4 viewed along the c (left panel) and a axes (middle panel), and Na1 and Na2 coordination environments (right panel). The inset of the middle panel illustrates diffusion channels along the a-direction composed by S-anion distorted cubic sublattices (from [130] open access).
Figure 11. The crystal structure of Na3SbS4 viewed along the c (left panel) and a axes (middle panel), and Na1 and Na2 coordination environments (right panel). The inset of the middle panel illustrates diffusion channels along the a-direction composed by S-anion distorted cubic sublattices (from [130] open access).
Batteries 11 00292 g011
Figure 12. Crystal structure of cubic Na2.88Sb0.88W0.12S4 with the unit cell outlined. The Na, Sb, W, and S sites are represented by blue, orange, gray, and yellow balls, respectively. Na is linearly arranged, and distortion of the SbS4/WS4 tetrahedra is very small compared to the tetragonal structure of Na3SbS4 (from [151] open access).
Figure 12. Crystal structure of cubic Na2.88Sb0.88W0.12S4 with the unit cell outlined. The Na, Sb, W, and S sites are represented by blue, orange, gray, and yellow balls, respectively. Na is linearly arranged, and distortion of the SbS4/WS4 tetrahedra is very small compared to the tetragonal structure of Na3SbS4 (from [151] open access).
Batteries 11 00292 g012
Figure 13. The effect of NaI upon conductivity of Na2.88Sb0.88W0.12S4·xNaI (from [149] open access).
Figure 13. The effect of NaI upon conductivity of Na2.88Sb0.88W0.12S4·xNaI (from [149] open access).
Batteries 11 00292 g013
Table 1. Comparison of the characteristics of all-solid-state sodium batteries based on different electrochemical systems.
Table 1. Comparison of the characteristics of all-solid-state sodium batteries based on different electrochemical systems.
ElectrolyteCathodeAnodeVoltage Range, VIonic
Conductivity of Electrolyte, RT, S/cm
Cathode
Specific
Capacity,
mAh/g
CyclesRef.
Na3Zr2Si2PO12Na2MnFe(CN)6Na4.0–2.51.2·10−3
at 60 °C
120 at 0.5 C200[39]
Na3.4Zr2(SiO4)2.4(PO4)0.6Na3V2(PO4)3Na3.8–2.05·10−30.6 mAh/cm2 at 0.5 mA/cm270[45]
Na3Zr2Si2PO12Na3V2(PO4)3Na3.8–2.06·10−488 at 0.2 C100[46]
Na3Zr2Si2PO12NaCu1/9Ni2/9Fe1/3Mn1/3O2BaTiO3/C@Al3.8–2.06·10−493.1 at 0.1 mA/cm2300[54]
Na3Zr2Si2PO12Na3.5V0.5Mn0.5Fe0.5
Ti0.5(PO4)3
Na4.4–2.06·10−3152 at 1 C1000[65]
Na3Zr2Si2PO12Na3V1.5Cr0.5(PO4)3Na4.3–2.50.85·10−3103 at 1 C400[66]
Na3.2Zr1.9Ca0.1Si2PO12Na3V2(PO4)3Na3.8–2.71.67·10−394.9 at 1 C450[68]
Na3.2Hf1.9Ca0.1Si2PO12Na3V2(PO4)3Na/SnO23.8–2.71.07·10−3103.1 at 0.5 C300[69]
β″-Al2O3 Na3V2(PO4)3Na3.8–2.51.17·10−4100 at 0.2 C100[76]
Na3Zr2Si2PO12Na3V2(PO4)3Na4.0–2.64.3·10−4110 at 0.1 mA/cm2900[80]
Na3Zr2Si2PO12Na3V2(PO4)3Na4.0–2.53.3·10−490 at 0.1 C100[84]
Na3.4Zr1.6Sc0.4Si2PO12Na3V2(PO4)3Na3.8–2.31.77·10−4100 at 1 C300[86]
Na-β″-Al2O3Na0.66Ni0.33Mn0.67O2Na3.8–2.51.0·10−379 at 6 C10,000[93]
Na3.4Zr1.9Zn0.1Si2.2P0.8O12FeS2Na2.5–0.75.27·10−3236.5 at 0.1 C
133.1 at 0.5 C
100
300
[121]
Na3.1Zr1.95Mg0.05Si2PO12Na0.9Cu0.22Fe0.3Mn0.48O2Na3.8–2.51.33·10−357.9 at 0.5 C100[126]
Na3PS4NaCrO2Na15Sn44.0–1.04.6·10−490 at 0.013 mA/cm220[129]
Na3PS4NaCrO2Na15Sn43.0–1.04.0·10−425010[131]
Na3SbS4NaCrO2Na4.0–1.51.1·10−3108 at 0.05 mA/cm21[134]
t-Na2.9375PS3.9375Cl0.0625TiS2Na2.4–1.21.14·10−380 at 0.1 C10[140]
Na3P0.62As0.38S4TiS2Na-Sn2.4–1.21.46·10−3150 at 0.02 C10[147]
Na2.88Sb0.88W0.12S4·0.5NaITiS2Na-Sn2.4–1.210−2100 at 0.057 C5[149]
Na2.88Sb0.88W0.12S3.7O0.3TiS2Na-Sn2.4–1.22.2·10−3130 at 0.057 C5[150]
Na2.85P0.85W0.15S4TiS2Na-Sn2.4–1.28.8·10−3140 at 0.057 C5[153]
Na3.0PS3.8Cl0.2Na3V2(PO4)3Na3.8–2.51.96·10−3100 at 0.1 C10[157]
Na0.7Zr0.3La0.7Cl4NaCrO2Na2Sn3.4–2.02.9·10−4114 at 0.3 C70[164]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Skundin, A.M.; Kulova, T.L. All-Solid-State Anode-Free Sodium Batteries: Challenges and Prospects. Batteries 2025, 11, 292. https://doi.org/10.3390/batteries11080292

AMA Style

Skundin AM, Kulova TL. All-Solid-State Anode-Free Sodium Batteries: Challenges and Prospects. Batteries. 2025; 11(8):292. https://doi.org/10.3390/batteries11080292

Chicago/Turabian Style

Skundin, Alexander M., and Tatiana L. Kulova. 2025. "All-Solid-State Anode-Free Sodium Batteries: Challenges and Prospects" Batteries 11, no. 8: 292. https://doi.org/10.3390/batteries11080292

APA Style

Skundin, A. M., & Kulova, T. L. (2025). All-Solid-State Anode-Free Sodium Batteries: Challenges and Prospects. Batteries, 11(8), 292. https://doi.org/10.3390/batteries11080292

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

Article Metrics

Back to TopTop