Next Article in Journal
Aerodynamic Analysis of a Low-Speed Tandem-Channel Wing for eVTOL Aircraft Considering Propeller–Wing Interaction
Next Article in Special Issue
Thin-Film Lithium Cobalt Oxide for Lithium-Ion Batteries
Previous Article in Journal
Potential Application of Canola Hull Fuel Pellets for the Production of Synthesis Gas and Hydrogen
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Electrode/Electrolyte Interphases of Sodium-Ion Batteries

by
Tatiana L. Kulova
* and
Alexander M. Skundin
Frumkin Institute of Physical Chemistry and Electrochemistry, Russian Academy of Sciences, Leninsky Prospect, 31-4, 119071 Moscow, Russia
*
Author to whom correspondence should be addressed.
Energies 2022, 15(22), 8615; https://doi.org/10.3390/en15228615
Submission received: 16 October 2022 / Revised: 9 November 2022 / Accepted: 14 November 2022 / Published: 17 November 2022
(This article belongs to the Special Issue Advanced and Multifunctional Materials for Energy Storage Systems)

Abstract

:
The performance of sodium-ion batteries largely depends on the presence and properties of passive films formed on the electrode/electrolyte interfaces. Passive films on negative electrodes inevitably result from the reduction in electrolyte components (solvent and salt anion). They have the properties of a solid electrolyte with sodium ion conductivity and are insulators in terms of electronic conductivity. Usually, they are called SEI—solid electrolyte interphase. The formation of SEI is associated with the consumption of a certain charge, which is an irreversible capacity. Passive films on the surface of positive electrodes (CEI—cathode electrolyte interphase) arise as a result of electrolyte oxidation. The present review summarizes the literature of the recent 15 years concerning the effects of electrode nature (hard carbon, other carbon materials, various metals, oxides, chalcogenides, etc.), electrolyte composition, and other factors on composition and properties of SEIs in sodium-ion batteries. Literary data on CEIs are reviewed as well, although their volume is inferior to that of data on SEIs.

1. Introduction

One of the most important notions inherent in lithium-ion batteries is the concept of a passive film on the negative electrode, commonly called a solid electrolyte interface (SEI) [1]. Actually, the concept of SEI was formulated long before the advent of lithium-ion batteries in connection with the question of the kinetic stability of metallic lithium in aprotic electrolytes [2,3]. Because it is the most negative metal, lithium reduces the electrolyte (primarily an aprotic solvent) and simultaneously corrodes. Thus, lithium is thermodynamically unstable in aprotic electrolytes. Under certain conditions, insoluble electrolyte reduction products form a passive film on the lithium surface, which has the properties of a solid electrolyte with lithium-ion conductivity and negligible electronic conductivity. The presence of this film (SEI) does not prevent the flow of the current-generating process, namely the anodic dissolution of lithium (for rechargeable batteries, both anodic dissolution and cathodic reduction in lithium) but prevents direct contact of lithium with the electrolyte and the reduction in the latter. The electrolyte reduction is an undesirable (parasitic) process, and the electricity expended in this process represents an irreversible capacity loss. A good SEI contributes to minimizing irreversible capacity, i.e., improving battery performance in general.
Later it was found that SEI is formed not only on the metal lithium electrode but also on the negative electrodes of lithium-ion batteries, the operating potential of which is quite negative. The formation of SEI on electrodes made of carbon materials has been investigated in length.
On the positive electrodes of lithium-ion batteries, which are characterized by rather high values of operating potentials, electrolyte oxidation is possible with the formation of corresponding passive films. Such films were named cathode electrolyte interphase (CEI) [4]. (The term solid permeable interphase (SPI) [5] is less commonly used).
The very existence and role of SEI and CEI in sodium-ion batteries are taken for granted. There are much fewer studies of SEI and CEI in the sodium system than similar studies in the lithium one, but there are still several reviews on sodium SEI and CEI [6,7,8,9,10,11,12,13,14,15].
The principle of SEI functioning is depicted in Figure 1. Coarse-solvated Li+ ions approach an electrode surface coated with SEI. They cannot penetrate SEI, but after desolvation, small Li+ ions are transferred through SEI and fit in the crystal lattice. SEI consists of inorganic particles embedded in an organic (polymer, oligomer) matrix.
The composition and properties of SEI and CEI depend on many factors, including the nature of the electrode material, the nature of the electrolyte and the presence of additives in the electrolyte, the pre-history of the electrode, etc.

2. SEI on Sodium Metal

Sodium metal is not used in sodium-ion batteries (except for all-solid-state ones). Sodium metal electrodes are intended for sodium-oxygen [16], sodium-sulfur [17], sodium-carbon dioxide [18], and so on systems. However, it is reasonable to consider the formation of SEIs on sodium metal, especially considering the difference between such SEIs and SEIs on lithium metal. Although lithium and sodium are very similar in physical and chemical properties, and lithium and sodium power sources have a similar principle of operation, their electrochemical behavior is very different. The nucleation and growth of sodium dendrites are more difficult than for lithium ones due to the higher energy barrier and the slowness of processes with sodium compared to lithium. Accordingly, sodium deposition overpotential is typically higher than for lithium, resulting in a lower discharge voltage for power sources with sodium electrodes. At the same time, sodium is more reactive with respect to electrolyte than lithium.
At the first contact of sodium metal with an electrolyte under open circuit conditions, a continuous primary passive film is formed on its surface. With further cathodic polarization, the growth of sodium dendrites begins, mainly in the places of cracks and other defects in the passive film. The interaction of the surface of the dendrites with the electrolyte leads to the well-known phenomenon of encapsulation and the removal of part of the sodium from the further electrochemical process. In contrast to lithium needle-like dendrites, sodium dendrites have a branched, bushy, or mossy shape with a much more developed surface [19]. Thus, natural SEIs on sodium has a much lower protective ability than on lithium [10,11,12,19,20,21,22,23]. For example, it was shown in [12] that the potential of a lithium electrode at both anode and cathode current density of 1 mA/cm2 in a 1 M LiPF6 in a mixture of ethylene carbonate (EC), propylene carbonate (PC), and diethyl carbonate (DEC) retains a value that does not change with time, while the potential of a sodium electrode at the same current densities in a 1 M NaClO4 solution in the same solvent oscillates with a large amplitude. These oscillations point to the continuous destruction and healing of SEI on sodium under these conditions. Moreover, it turned out that SEIs on sodium is much more soluble in the electrolyte than SEIs in lithium systems [24]. It was shown in [25] that, under open circuit conditions, the impedance of the interface between sodium metal and an electrolyte (1 M NaPF6 in a mixture of EC with dimethyl carbonate (DMC) or in a mixture of EC:PC:DMC) is way more than the impedance of the interface between lithium metal and a similar electrolyte. Unlike the lithium system, the impedance in the sodium system increases continuously with time. That is why repeated attempts have been made to create “artificial” SEIs on sodium metal by optimizing the composition of the electrolyte, including using various additives to the electrolyte.
The properties of SEI formed in carbonate electrolytes depend markedly on the electrolyte composition, in particular, on the solvent nature. For example, it was shown in [26] that the polarization of a sodium electrode in 1 M NaClO4 in pure PC significantly exceeds that in 1 M NaClO4 in a PC:EC mixture.
As a rule, SEI formed in carbonate electrolytes has a much lower protective ability than SEI in ether electrolytes [10,12,22]. For example, it was shown in [22] that the replacement of the traditional carbonate electrolyte with NaPF6 solutions in glymes (mono-, di-, and tetraglyme) makes it possible to carry out the cathodic deposition of sodium without SEI destruction and without the formation of dendrites. SEIs, in this case, contain Na2O and NaF as mineral constituents and sodium alkoxides (RCH2ONa) as organic constituents. It was emphasized that the formation of high-quality SEI, in this case, is provided by both the solvent and the salt. Replacing sodium hexafluorophosphate with other salts, such as sodium imide, triflate, or perchlorate, while retaining glyme as a solvent, or replacing glyme with carbonates, resulted in SEI that did not prevent dendrite formation. However, in [27,28,29], the formation of reliable SEIs in a carbonate electrolyte, specifically 1 M sodium difluorooxalatoborate in an EC-DEC mixture, was reported.
Another example of a successful electrolyte for the functioning of a rechargeable sodium electrode is highly concentrated (about 5 M) solutions of sodium bisfluorosulfonylimide (NaFSI) in 1,2-dimethoxyethane (DME) [30], with a high concentration (close to saturation) of NaFSI being an indispensable condition for the formation of high-quality SEI. The authors of [31] came to the same conclusion somewhat earlier. It was shown in [32] that the same effect could be achieved in much less concentrated (1.5 M) NaFSI solutions in mixtures of 1,2-dimethoxyethane (DME) with bis(2,2,2-trifluoroethyl) ether (BTFE). It is high time to stress the fact that the elemental composition of SEI formed in a concentrated solution of NaFSI in DME and in a dilute solution of NaFSI in a DME-BTFE mixture turned out to be almost the same and differed from the composition of SEI formed in a dilute solution of NaFSI in DME (Figure 2). Figure 2 also clearly shows the change in SEI composition with depth. The content of carbon and nitrogen (that is, the organic component) on the surface of the SEI is much greater than at a depth of 10 nm. The ratio of solvents DME:BTFE in the mixture should be from 1:2 to 1:3. A solution of NaFSI in an ionic liquid, tri(isobutyl)methylphosphonium bis(fluorosulfonyl)imide, also proved to be a successful electrolyte [33]. At a NaFSI concentration of 45% in such an electrolyte, dense, high-quality SEIs were formed.
Somewhat exotic electrolytes that ensure the formation of high-quality SEIs on the surface of sodium metal were proposed in [34]. Here, ammonia-based quasi-ionic liquids were used, having the composition NaY·xNH3, where Y is any anion, in particular, NaI·3.3NH3, NaBH4·1.5NH3, and NaBF4·2.5NH3. These compounds have boiling points of 40, 18, and 10 °C, respectively, and are highly concentrated (7.6, 12.3, and 9.7 M, respectively) solutions of Na+ in liquid ammonia. (It is not out of place to mention that NaI·3.3NH3, as an electrolyte for primary cells with a sodium anode, was proposed back in 1988 [35]). In [36], a similar electrolyte based on the quasi-ionic liquid NaAlCl4·2SO2 is described. In such an electrolyte, SEI is formed on the sodium surface, consisting mainly of NaCl (along with small amounts of Na2O and Na2S), and completely prevents the formation of dendrites during cathodic sodium deposition.
As for various additives to the electrolyte, perhaps the most discussed additive is fluoroethylene carbonate (FEC) [37], which has good film-forming properties [38], suppresses the destruction of solvents, and improves the conditions for the electrode process [19,39,40,41,42]. The addition of FEC has proven itself well in the lithium system. When used in power sources with a sodium electrode, the concentration of this additive should be increased compared to lithium systems [12]. However, in the literature, one can also find an indication of the negative effect of the addition of FEC on the properties of SEI on sodium metal [30,43]. (But, for example, in [43], it is shown that the harmful effect of FEC can be neutralized by the addition of 1,1,1,3,3,3-hexafluoroisopropyl methyl ether). The addition of FEC leads to a significant enrichment of SEI with sodium fluoride, which was shown both by theoretical calculations [44,45,46,47] and by direct experiments [19,41,42]. Another important additive to the electrolyte that provides high-quality SEI on sodium metal is sodium sulfide, more precisely, sodium polysulfides [48].
In general, the nomenclature of additives that improve SEI on sodium metal is not at all systematic. For example, potassium bis(trifluoromethylsulfonyl)imide was mentioned as a useful additive [49]. (It was assumed that in the presence of TFSI–anions, the composition of SEI would include sodium nitride (Na3N) and sodium oxynitrides (NaNxOy); it is surprising that this idea is not mentioned in work [30], which is especially devoted to the imide electrolyte). In [50], SbF3 is announced as a useful additive to a concentrated imide electrolyte. Figure 3 vividly shows the effect of SbF3 additive to the electrolyte on dendrite formation. In the presence of such an additive, a bilayer SEI is formed on the sodium surface. The inner layer consists of a Na-Sb alloy. The outer layer is a regular SEI enriched with NaF (Figure 3c). This bilayer SEI is thinner than the SEI formed in the same electrolyte without the addition of SbF3 (Figure 3b) and ensures the deposition of a uniform dendrite-free sodium layer. A comparison of Figure 3a,b shows the advantage of the ether electrolyte over the carbonate one.
The authors of [51] propose the addition of NaAsF6 to carbonate electrolytes. It is indicated that in this case, SEI contains significant amounts of sodium fluoride and O–As–O polymer, and it is this combination that provides high protective properties of SEI. At the same time, authors of [52] postulate the detrimental effect of NaF on the properties of SEI on sodium metal and show that excellent SEI can be obtained in completely fluoride-free electrolytes, for example, in a solution of sodium tetraphenylborate (NaBPh4) in DME. In [53], tin chloride is proposed as an additive to carbonate electrolytes. It is assumed that in such electrolytes, an Sn–Na alloy is formed on the sodium surface (due to the contact reduction in tin ions), and SEI will be enriched with NaCl. It has also been proposed to apply a thin (sub-nanometer) layer of protective coating to the surface of sodium metal as some kind of “artificial SEI.” These can be layered 2D materials (hexagonal boron nitride, graphene, silicene, germanene, stannene, phosphorene, etc.) [54,55] or composite layers consisting of a polymer (for example, a copolymer of polyvinylidene fluoride with hexafluoropropylene (PVdF-HFP) saturated with an ordinary liquid electrolyte (plasticizer) with a filler of Al2O3 nanopowder) [56,57]. Another option for “artificial SEI” is a thin layer of sodium disulfide [58]. To form such an “artificial SEI,” a layer of molybdenum disulfide is applied to the surface of metallic sodium, which interacts with sodium according to the equation
4Na + MoS2 = 2Na2S + Mo
In [59], it is proposed to use not pure sodium metal but a sodium composite with reduced graphene oxide. In this case, it is not sodium that contacts the electrolyte, but graphene, on the surface of which a stable SEI is formed.

3. SEI on Hard Carbon

Of the wide variety of anode materials proposed for sodium-ion batteries, the most popular is hard carbon, the use of which in sodium-ion batteries was first described in [60,61]. In the non-graphitized disordered turbostratic structure of hard carbon, there are various places for sodium insertion (randomly oriented graphene layers with a “house of cards” structure, various defects, and micropores in stacks of graphene layers) [62,63,64]. As a result, reversible insertion and extraction of sodium occur over a relatively wide range of potentials. The galvanostatic charge curves show two almost straight segments: a decreasing segment in the potential range from about 1.0 to 0.1 V (hereinafter, all potential values are given relative to the sodium electrode) and a flat segment (plateau) from 0.1 to 0.01 V [62,65,66]. It is on the first falling section that the main charge is spent on the formation of SEI. Figure 4 explains how a comparison of the charge and discharge galvanostatic curves gives a possibility to find the reversible and irreversible capacity of the electrode.
It has been shown in many works that the composition of SEI on hard carbon electrodes substantially depends on the composition of the electrolyte [66,67,68,69,70,71,72,73,74,75,76,77,78,79,80]. Thus, according to [66], SEI formed in NaClO4 solution in butylene carbonate (BC) possesses worst protective properties than SEI formed in solutions in EC and PC (Figure 5a). Even more revealing is the comparison of NaClO4 solutions in mixtures EC:DMC, EC:ethyl methyl carbonate (EMC), and EC:DEC (Figure 5b).
The composition of the electrolyte affects not only the brutto-composition of the SEI but also the distribution of the SEI components over its thickness. Figure 6 shows the fundamental difference between conventional carbonate electrolytes (Figure 6a) and electrolytes based on ionic liquid (Figure 6b).
The composition of SEI formed in 1 M solutions of NaPF6, NaClO4, NaFSI, NaTFSI (sodium bis(trifluoromethanesulfonyl)imide), and NaFTFSI (sodium fluorosulfonyl-(trifluoromethanesulfonyl)imide) in a mixture of EC-DEC was studied in length in [69]. It was shown by X-ray photoelectron spectroscopy (XPS) that the organic part of SEI contains sodium carbonate (Na2CO3), sodium alkyl carbonates (ROCO3Na), sodium alkoxides (RONa), ethylene oxide oligomers, semicarbonates and double sodium alkyl carbonates formed as a result of the reductive destruction of solvents, and the relative amounts of these components depend on the nature of the salt anion. The nature of the salt anion also affects the density and thickness of SEIs: thicker SEIs are formed in the electrolyte with NaClO4, while thinner SEIs are formed in the electrolytes with NaPF6 and NaTFSI. The distribution of organic and inorganic (NaCl, NaF, Na2CO3) components over the SEI depth also depends on the nature of the anion. The authors of [70] came to similar conclusions. It is also stated in [76] that double alkyl carbonates constitute the main organic component of SEI on hard carbon.
It was shown in [71] that, in a solution of sodium tetraphenylborate in DME, quite stable SEIs are formed on hard carbon electrodes (as well as on sodium metal [52]), which do not contain fluorine compounds and ensure electrode cycling at current densities up to 6 A/g. The behavior of hard carbon in carbonate-based (1 M NaClO4 in an EC-PC mixture) and ether-based (1 M NaClO4 in a mixture of tetraglyme with DME) electrolytes were compared in more detail in [80]. It was found that in the former case, loose, non-continuous SEIs, whereas in the latter case, thin, dense, continuous SEIs of uniform thickness is formed. As a result, in the carbonate electrolyte, the electrode processes proceed with greater polarization. Just as in the case of sodium metal, the composition and properties of SEI on hard carbon depend markedly on the presence of additives in the electrolyte. The most popular additive is again FEC [37,67,70,78,81,82,83], which promotes the formation of thinner dense SEIs. (Intriguingly, such an additive as vinylene carbonate (VC) turned out to be much less effective in sodium-ion batteries than in lithium-ion batteries [37]). The addition of FEC leads to a significant improvement in the properties of SEIs on hard carbon in those electrolytes in which low-quality SEIs generally form, for example, in solutions of NaClO4 in plain PC [84]. At the same time, in electrolytes based on an EC-PC mixture, the addition of FEC led to a deterioration in the quality of SEI [63].
The reductive decomposition of FEC was found to proceed with significantly lower activation energy than the propylene carbonate reduction [46,47,85]. This conclusion was made by using the density functional theory (DFT) and the method of molecular dynamics. This very feature determines the easier formation of SEI enriched with sodium fluoride. The efficiency of the addition of FEC was shown to depend on its concentration [81,83]. For 1 M solutions of NaPF6 in PC and EC:PC mixture, the optimal concentration of FEC is 0.5%. It is also stated in [81] that it is the combination of NaPF6 with FEC that provides a synergistically successful SEI on hard carbon electrodes.
It is commonly accepted that SEI is formed mainly in the first cycle. However, a number of studies have shown that the evolution of the composition and structure of SEI on hard carbon continues over the initial 10–20 cycles. Such a conclusion, in particular, was made in [76], where it was found that in a solution of 1 M NaTFSI in PC with the addition of 3% FEC, SEI on hard carbon thickens and enriches with sodium carbonate and sodium fluoride over the first 20 cycles.
Other additives that have a beneficial effect on the formation of high-quality SEI on hard carbon include tris(trimethylsilyl)phosphite (TMSP, [(CH3)3Si]3PO3) [86], 1,3-propanesultone (C3H6O3S) and ethylene sulfate (C2H4SO4) [87], ionic liquids [88,89,90]. The role of TMSP in improving SEI on hard carbon is that the TMSP additive traps traces of oxygen, moisture, as well as HF and PF5 in the electrolyte [91,92]. Additives of sulfur-containing 1,3-propanesultone and ethylene sulfate contribute to the formation of sulfates and sulfites (ROSO2Na and RSO3Na), which stabilize SEI on hard carbon during long-term cycling. The examples of ionic liquids are 1-butyl-3-methyl-imidazolium bis(trifluoromethylsulfonyl)imide, 1-ethyl-3-methyl-imidazolium bis(trifluoromethylsulfonyl)imide, and N-propyl-N-methylpyrrolidinium bis(trifluoromethylsulfonyl)imide [88], N-propyl-N-methylpyrrolidinium bisfluorosulfonylamide [89] and 1-ethyl-3-methylimidazolium bisfluorosulfonylimide [90]. It is worth noting that in all cases, the presence of ionic liquids in the electrolyte contributes to an increase in the mechanical and electrochemical stability of SEI due to a change in its composition, although no quantitative data on the composition of SEI are given.
Somewhat unexpected was the effect of nano-zeolite additions to the active mass of hard-carbon-based electrodes on the SEI characteristics [93]. The replacement of part of the carbon black (as an electrically conductive additive) in the active mass of the electrode with alumina-silicate zeolite Nano-ZSM5 P91 (Zeolite Socony Mobil-5) led to a certain decrease in the thickness and resistance of the SEI. It puzzles why the authors of [93] do not comment on the fact that the replacement of 80% of the electronic conductor (carbon black) with an ionic conductor (zeolite) was not accompanied by a deterioration in the characteristics of the electrode even at elevated currents (up to 5 C).
Attempts have been made to create artificial SEIs on solid carbon. For example, paper [94] describes hard carbon fibers on the surface of which arrays of V2O5 nanosheets are grown. The total amount of V2O5 was 14–20%. When V2O5 nanosheets were deposited, the surface of hard carbon was “healed” (decrease in the density of defects and nanovoids), which ensured the formation of a thinner dense SEI. In [95], it is proposed to deposit an Al2O3 layer about 2 nm thick on the surface of hard carbon. Such a layer is applied by atomic layer deposition (ALD) and serves as a sublayer for SEI. The authors showed that the SEI formed on such a sublayer has a lower thickness and a higher ionic conductivity. The alumina sublayer is assumed to slow down the process of electrolyte reduction on the surface of hard carbon. Amorphous (soft) carbon coatings obtained from coal tar pitch exert approximately the same effect [96,97].
In general, the composition and properties of SEIs formed on hard carbon are highly dependent on the state of its surface. As a rule, the less defective the surface and the lower the porosity of the surface layer, the perfect SEIs are formed on such surfaces [98,99,100]. In the synthesis of hard carbon by pyrolysis of various precursors, the state of the surface significantly depends on the temperature and mode of pyrolysis. Thus, it was shown in [98] that when hard carbon is obtained by pyrolysis of peat, samples synthesized at a temperature of 1400 °C are characterized by the maximum reversible and minimum irreversible capacity and, accordingly, the thinnest and densest SEI. In [100], it was found that at a constant temperature of sucrose pyrolysis, the surface of the resulting hard carbon is more perfect (defect-free), and the slower the heating during pyrolysis occurs. A somewhat different conclusion was formulated in [101], where it was found that thinner and more perfect SEIs are formed on microporous hard carbon than on samples with low porosity.
An original method for stabilizing SEI on hard carbon is described in [102]. Here it is proposed to carry out preliminary short-term treatment of the surface of hard carbon in oxygen plasma [103,104]. In this case, oxygen-containing functional groups are formed on the carbon surface, and the surface itself becomes hydrophilic. On such a surface, very thin SEIs are formed with negligible irreversible capacitance.
Some interesting observations were reported in [105]. It is shown here that dense thin (8 nm) SEIs are formed on electrodes made of hard carbon obtained by pyrolysis of longan peel at room temperature, providing stable cycling with a sufficiently high reversible capacity. Thicker (21 nm) SEIs are formed on the same electrodes at a temperature of −20 °C, which leads to strong degradation during cycling (capacity decreased from 280 to 80 mAh/g in 5 cycles). However, upon subsequent heating to room temperature, the SEI was transformed, i.e., it returned to its previous thickness with a simultaneous increase in capacity up to 250 mAh/g. The expected results on the effect of current density during the charging of electrodes based on hard carbon are reported in [106]. It is shown here that increasing the cathode current density leads to the formation of thinner perfect SEIs. Thus, according to the data of electrochemical impedance spectroscopy, the total electrolyte, and SEI resistance, corresponding to the active component of the impedance at an infinitely high frequency, was 1.3 Ohm·cm2 at a charge with a current of 100 A/g and 15.1 Ohm·cm2 at a charge with a current of 1 A/g.
When SEI is formed, a significant charge is consumed on the first cycle, and accordingly, a certain amount of sodium ions is lost from the positive electrode. In order tto compensate for this loss, it is recommended that the electrode, in particular, a hard-carbon-based one [107], be pre-sodiated. A simple but effective method of pre-sodiation consists in spraying a sodium naphthaline (Naph-Na) solution onto a carbon electrode. Figure 7a shows how the first charge curve changes upon pre-sodiation: open circuit voltage shifts from 2.35 V to 0.75 V, and irreversible capacity diminishes by 60 mAh/g. Some interfacial layer with a thickness of ≈20 nm is formed upon pre-sodiation (Figure 7b). This layer consists of sodium-containing-organic and inorganic compounds, such as sodium alkyl carbonates, sodium carboxylate, sodium carbonate, sodium fluoride, etc.
Works [108,109,110] are devoted to a detailed analysis of the morphology and composition of SEI on hard carbon. Various characteristics of sodium alkyl carbonates are given in [111]. It was found in [112] that SEI on hard carbon grows inward rather than outward, i.e., in the direction from the interface with the electrolyte to the carbon surface.

4. SEI on Other Anode Materials

The list of anode materials in addition to hard carbon is extensive [113] and includes various other forms of carbon [114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145], metals, and their composites with carbon (Sb [146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161], Sn [162,163,164,165,166,167,168], Bi [169,170,171,172,173,174,175], Cu [176]), silicon [177,178], germanium [179,180] oxides [181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198,199,200,201,202,203,204,205,206,207,208,209,210,211,212,213,214,215,216,217,218,219,220,221,222,223,224,225,226,227,228], chalcogenides [229,230,231,232,233,234,235,236,237,238,239,240,241,242,243], phosphides and phosphorus composites [244,245,246,247,248,249,250,251], as well as more complex compounds from sodium titanates [252,253,254,255], to very exotic compounds [256,257,258,259,260]. As a result, information on each specific material is rather limited.
Some regularities outlined in Section 1 and Section 2 happen to be valid also for many other anode materials. For example, the advantages of ether-based electrolytes over carbonate electrolytes have been confirmed for the formation of SEI on graphite [130,142], tin [162,165], bismuth [261], bismuth sulfide [229], bismuth oxychloride [182], Na3(TiOPO4)2F [256], molybdenum and iron sulfides [231,237], TiO2 [187], etc. A favorable effect of FEC is noted in the formation of SEI on antimony [147,148,149,153], tin [167], bismuth-carbon composites [169], SnO [222], SnO2 composites with reduced oxide graphene [224,262], a composite of antimony sulfide with graphene [263], etc. There are examples in the literature of the positive effect of pre-sodiation on the reduction in the irreversible capacity of the first cycle [138,253,264].
At the same time, the composition and structure of SEIs and the features of the processes of their formation and destruction on different materials are, in general, individual. So, on different forms of carbon, even in the same electrolyte, SEIs are formed in several different potential ranges. For example, on carbon nanofibers doped with fluorine and nitrogen, stable high-quality SEIs in an EC-PC-based electrolyte are formed at potentials of about 0.92 and 0.46 V (these values correspond to extrema on voltammograms) [115]. On a carbon material obtained by heat treatment of a mixture of resorcinol and formaldehyde and having the shape of smooth spherical particles (i.e., having a small specific surface area), the SEI formation potential is close to 1.1 V [114]. It is of deep interest that the cited work reports the formation of very thin dense SEIs in an electrolyte based on a mixture of EC-DEC. This fact comes as rather a surprise and contradicts most other works in which SEIs in such an electrolyte have a large thickness and defectiveness (see, e.g., [80]). Thin SEIs on hollow fibers obtained by carbonization of polyaniline is formed at potentials of 0.43 and 1.13 V [117]. At the same time, in the experiments with carbon fibers made by cellulose carbonization, a certain irreversible capacity was registered in the potential range of 0.2 to 0.3 V in EC-PC-based electrolytes [118].
Although graphite is inferior to hard carbon as an anode material for sodium-ion batteries, a sufficient number of studies have been published on the reversible sodium insertion into graphite and the formation of SEI [120,121,127,128,129,130,131,141]. It has been established that, such as lithium, sodium is intercalated into graphite in stages, with the formation of distinct steps on galvanostatic curves [127]. Usually, solvated (or incompletely desolvated) sodium ions are intercalated into graphite. For example, sodium ions bound to one solvent molecule are intercalated from an electrolyte based on diethylene glycol dimethyl ether (DEGDME) [130]. It was shown in [128] that sodium ions solvated by two solvent molecules are reversibly intercalated from a diglyme-based electrolyte. In [141], the formation of SEI was studied in detail upon the sodium intercalation into graphite from a solution of NaFSI in tetraethylene glycol dimethyl ether (TEGDME). Under these conditions, SEI has a thickness of 3 to 8 nm and consists of the reduction products of both the salt anion and the solvent. In [131], a paradoxical conclusion was made that when a graphite electrode is cycled in an electrolyte based on diglyme, SEI is not formed at all. An explanation of this paradox can be found in [140]. Here, using soft X-ray absorption spectroscopy (XAS) it was shown that in an ether-based electrolyte (1 M NaPF6 in DEGDME), SEI on graphite is formed during cathodic polarization and is destroyed during subsequent anodic polarization, i.e., shows signs of reversibility. Under the same conditions on disordered (hard or soft) carbon, such reversibility is only partially revealed, and an appreciable part of SEI is formed irreversibly during cathodic polarization and does not undergo transformations during subsequent anodic polarization. In carbonate electrolytes, sodium is not reversibly intercalated into graphite at all [142].
In the example of various carbon electrodes, a certain correlation was found between the specific surface area (or dispersion), and the volume of SEIs formed, i.e., irreversible capacity. For example, it was shown in [124] that when sodium is inserted into carbon xerogels obtained from a resorcinol–formaldehyde mixture and having pores ranging in size from 10 to 200 nm, the irreversible capacity in the first cycle is 85% for the finest and 67% for the largest pores samples. A similar conclusion was made in [125] for porous carbon obtained by carbonization of polyethylene oxide and polystyrene, in [121] for a material obtained by carbonization of peanut hulls, and in [126] for a nanoporous material obtained by carbonization of ZIF-8 zeolite.
The authors of [119,144] describe a carbon material that is, in fact, highly porous (with a specific surface area of more than 1000 m2/g and sizes of micropores from 0.5 to 2 nm and mesopores from 2 to 100 nm) carbon impregnated with sodium. Sodium occupied only part of the pore space. The rest was filled with an electrolyte so that the electrode worked to the entire depth of the active layer. In this case, the layer of oxygen-doped carbon foam was obtained from starch and deposited on a current copper collector. A dense thin SEI containing NaCO3R, NaF, and Na2O was formed on such an electrode in a diglyme-based electrolyte, which ensured stable cycling. The outer layer of SEI was enriched in organic components, while the inner layer was enriched in inorganic ones, which is generally typical of SEIs formed in electrolytes based on ethers [120].
The authors of [137] studied electrodes obtained from graphene, on which Al2O3 particles about 1 nm in size are fixed. The introduction of these nanoparticles led to a sharp decrease in the defectiveness of graphene layers and the formation of thin homogeneous SEI. Moreover, the presence of such nanoparticles protected the SEI from the harmful effects of HF traces. This effect is similar to that described in [95] for hard carbon.
Of particular interest is the behavior of carbon electrodes in electrolytes based on ionic liquids, i.e., electrolytes that do not contain traditional aprotic solvents [265]. It has been found in many studies that even in electrolytes based on ionic liquids, SEIs are formed on the electrode surface during the first cathodic polarization, the characteristics of which differ from those of “ordinary” SEIs. In particular, the insertion of sodium into electrodes, which are a composite of graphene nanosheets and carbon microspheres in 1 M NaFSI in N-propyl-N-methylpyrrolidinium bis(fluorosulfonyl)imide (PMP–FSI) and in 1 M NaTFSI in N-propyl-N-methylpyrrolidinium bis(trifluoromethanesulfonyl)imide (PMP-TFSI) was studied in [145]. On the voltammograms recorded in 1 M NaFSI in PMP-FSI, a large cathode peak was noted at potentials of about 0.6 V during the first cycle, which the authors identified with the SEI formation due to anion reduction. This peak did not appear in subsequent cycles. According to the XPS analysis, the composition of the SEIs formed in the ionic liquid electrolyte differed from the composition of the SEIs formed in the traditional EC-DEC electrolyte. In the former case, SEIs were enriched in polyolefins ((CH2)n) and compounds with S=C=O groups; in the latter case, they had a high content of sodium carbonates and alkyl carbonates and compounds with C–O and C=O groups. It is worth noting that SEIs formed in 1 M NaTFSI in PMP-FSI had a much higher Ohmic resistance and did not provide reliable cycling. A similar conclusion about the composition of SEIs formed in an electrolyte based on PMP–FSI was made in [266].
Antimony has a specific capacity of about 600 mAh/g, is an attractive anode material for sodium-ion batteries. True enough, the working discharge potential of antimony electrodes is noticeably more positive than that of carbon-based electrodes and amounts to 0.7–1.2 V. Even with such positive potentials, SEI is formed on the Sb electrodes, the composition, and properties of which being depended on the nature of the electrolyte. Noteworthy is that it was shown in [157] that thinner and more reliable SEIs are formed on antimony powder electrodes made using carboxymethyl cellulose as a binder in 1 V NaClO4 in PC with the addition of FEC than on electrodes made with polyvinylidene fluoride binder (in the first case, the irreversible capacity of the first cycle was half that of the second case).
The properties of SEI on antimony powder electrodes with carboxymethyl cellulose binder were studied in [146] using three electrolytes: 1 M NaClO4 in PC, 1 M NaPF6 in PC, and 1 M NaPF6 in an EC:DMC mixture. All electrolytes were supplemented with 5% FEC. The irreversible capacity of the first cycle was minimal in an electrolyte based on an EC:DMC mixture and maximal in a sodium perchlorate solution.
In the example of antimony electrodes, the conclusion made in the study of carbon electrodes was confirmed that an increase in dispersion is accompanied by an increase in the irreversible capacity of the first cycle [148].
An interesting observation is reported in [152]. Here, cathodic deposition was performed from a solution containing nickel chloride (NiCl2), L-antimony potassium tartrate (C8H4K2O12Sb2), boric acid (H3BO3), and sodium hypophosphite (NaH2PO2). The resulting precipitate was cauliflower-like antimony particles and some (up to 12%) SnNi alloy particles. According to the authors, it was the presence of this alloy that provided dense, high-quality SEIs and high cyclability. In [154], it was found that a change in temperature in the range from 5 to 50 °C leads not so much to a change in the SEI Ohmic resistance on nanocrystalline antimony but to a change in the charge transfer resistance at the interface between the SEI and the electrode. The paper [155] describes an original material that has a garnet structure and consists of secondary spherical particles ranging in size from 2 to 10 microns, each of which contains many yolk-shell particles (antimony core, carbon shell). On such a structure, relatively thick but sufficiently conductive SEIs are formed, which provide damping of volumetric changes during sodium insertion and extraction. A composite consisting of antimony nanoparticles in a porous carbon matrix turned out to be an effective material for the negative electrode of sodium-ion batteries [159]. SEI on such a composite in a carbonate electrolyte with the addition of FEC is formed in a wide range of potentials and ensures stable cycling. Other antimony composites with different carbon materials [160,161] form similar SEIs. Another interesting example of an antimony-containing material is given in [156]. Here, thin plates consisted of antimony nanoparticles with shells of amorphous TiO2−x. We deem it possible to assume that TiO2−x stabilizes the SEI on such electrodes.
The properties of SEIs on tin formed in electrolytes based on alkyl carbonates and based on diglyme were investigated in [162] using cryogenic transmission electron microscopy and XPS. It was shown that in the first case, thick loose SEIs are formed, while in the second case, ultrathin dense SEIs are formed, consisting of a polymer matrix in which amorphous inorganic nanoparticles are dispersed. It is confirmed in [166] that SEIs on tin formed in 1 M NaClO4 in PC are enriched mainly in sodium carbonate and alkyl carbonates and are partially destroyed by anodic polarization.
The additives of FEC have a significant effect on the properties of SEI. Figure 8a shows the discharge curves on an SnSb-carbon nanotube composite electrode in solutions of NaPF6 in EC:DEC containing and not containing FEC additives. This Figure also shows the effect of FEC on Coulombic efficiency. With regard to tin electrodes, glyme-based electrolytes also have advantages over carbonate-based ones (Figure 8b).
The features of SEI on bismuth have been studied mainly on samples of various composites of bismuth with carbon. In [169], a composite obtained by carbonization of bismuth citrate and containing 83% bismuth was studied. It is also shown here that loose SEIs with a thickness of about 15 nm containing ester polymerization products are formed in the carbonate electrolyte. Dense SEIs about 3 nm thick was deposited in an electrolyte based on ethylene glycol dimethyl ether. Similar results were obtained in [172,173]. The authors of [170] describe a composite in which bismuth nanospheres are encapsulated in porous carbon doped with nitrogen. On such a composite in 1 M NaPF6 in DME, stable SEIs are realized, which ensure the reversible sodium insertion at currents up to 100 A/g. However, without some special measures, such electrodes demonstrate rather high irreversible capacity at the first cycle (Figure 9a). Almost the same material is described in [175]. A peculiar composite of bismuth with carbon is described in [171,174]. Here, bismuth nanorods are enclosed in carbon nanotubes doped with nitrogen. On such a composite, SEIs at cathodic polarization appeared at potentials of about 0.64 V (which was noted by a rather narrow peak in the voltammogram) and ensured stable cycling (2600 cycles at a current of 1 A/g). In [176], copper was used as a model material for studying the mechanical properties of SEIs formed in 1 M NaPF6 in an EC–DEC mixture. It is unclear, however, to what extent the conclusions of this work can be extended to other substrates and other electrolytes.
It is known that sodium (unlike lithium) practically does not insert into crystalline silicon [267,268]. However, there are indications of the possibility of reversible insertion of sodium into amorphous silicon [178] from an electrolyte based on an EC-PC mixture with the consumption of a noticeable irreversible capacity for the formation of SEI at potentials about 1.1 V (Figure 9b). In [177], the process of sodium insertion into clathrate (openwork) structures of type II silicon (type II clathrates have the formula NaxSi136, where 0 < x < 24), and it was found that in this case, the formation potential and some properties of SEI depend on temperature, since side processes are accelerated with increasing temperature, leading to irreversible capacity.
The closest analog of silicon, germanium, on the contrary, is capable of reversible sodium incorporation [179]. In this case, side processes of electrolyte reduction also occur, accompanied by irreversible capacity, but we are unaware of research on the composition and structure of the resulting SEI. It is worth mentioning that the irreversible capacity of germanium (in contrast to carbon materials) is noticeably reduced by the addition of VC [180]. The effect of VC is vividly shown in Figure 10.
Many oxides, including Fe3O4, are used as anode materials in the form of composites with carbon or in the form of carbon-coated nanoparticles. One can venture a guess that it is the presence of carbon components that determines the nature of SEI on such materials. In [181], on an electrode of Fe3O4 nanoparticles with a thin carbon coating in an electrolyte consisting of 1 M NaClO4 in ethyl methanesulfonate with the addition of FEC, the formation of SEI was recorded at potentials of about 1.3 V. In [188], Fe3O4 nanorods with a thin carbon coating in an electrolyte, consisting of 1 M NaClO4 in the usual EC:PC mixture were studied, and the formation of SEI was found only at potentials of about 0.4 V. This difference is undoubtedly due to the difference in the nature of the electrolyte. In [190], as well as in [189,191], nanostructured Fe2O3 electrodes without carbon coating on individual particles are described (however, using 20% carbon black as an electrically conductive additive in the active mass of electrodes). On such electrodes in 1 M NaClO4 in an EC:DEC mixture, the formation of SEI occurred in the potentials about 0.4 V, i.e., in the same way as on Fe3O4 nanorods with a carbon coating.
Titania is one of the popular candidates for the negative electrode material of sodium-ion batteries. As a rule, it is used in the modification of anatase, and both carbon-coated and uncoated materials are described. Already in the first publications [196,197,198], the existence of an appreciable irreversible capacitance on such electrodes associated with the formation of SEI was noted. In [196], the irreversible capacity appeared only in the first cycle, while in [197,198], it gradually decreased from the first to the tenth cycle. As a rule, in solutions based on an EC:PC mixture, SEIs are formed at potentials of 1.1–1.2 V [199,200], although in [196], a potential of about 0.8 V is indicated, and in [201], about 0.4 V for NaClO4 solution in pure PC. In [200], the influence of the nature of the electrolyte on the formation of SEI was studied in more detail, and the effect of the solvent was shown to be stronger than that of the salt anion. Thus, in a solution of NaTFSI in PC, the irreversible capacity in the first 10–15 cycles were significantly greater than in solutions of NaPF6 and NaClO4 in the same solvent. For NaClO4 solutions, the irreversible capacity decreased in the series of solvents (EC:DMC)-PC-(EC:PC). In the same work, the practical absence of the effect of the addition of 2% FEC on the electrochemical behavior of TiO2 in 1 M NaClO4 solutions in an EC:PC mixture was noted (which contradicts the results of many other studies, for example, [205]). The works [192,202] show the effect of pre-sodiation of electrodes based on TiO2 on the reduction in the irreversible capacity.
An original approach for reducing the irreversible capacitance on titania-based electrodes is described in [185]. Here, an array of anatase nanotubes is proposed to be preliminarily lithiated in a LiClO4 solution in PC, and then Li-containing SEIs are converted into Na-containing SEIs in a NaClO4 solution.
Tin oxides and their composites with carbon have also been considered promising materials for the negative electrodes of sodium-ion batteries. The processes of SEI formation on such materials were studied in [219,220,221,222,223,224,225,226,227,228]. It was found in [219] that the composition of SEI on the SnO2–reduced graphene oxide composite formed in an electrolyte based on EC:DEC with the addition of FEC includes NaF, Na2CO3, and sodium alkyl carbonates. On a SnO2 composite with multiwalled carbon nanotubes in a NaClO4 solution in EC:PC, SEI was formed at potentials in the range from 0.5 to 0.2 V [220]. A somewhat unusual composition of SEI on the SnO2 nanocomposite with carbon Super P is presented in [228]. Here, in a 1 M solution of NaClO4 in an EC:DEC mixture with the addition of FEC, SEIs were formed containing, in addition to the usual sodium fluoride, carbonate, and alkyl carbonates, also sodium chloride, and chlorate. In [262], the behavior of electrodes made of a SnO2 composite with graphene in an electrolyte based on an ionic liquid (1 M NaFSI in PMP:FSI) was studied, and a cathode peak was recorded on voltammograms at potentials of about 0.6 V, associated with the formation of SEI, which turned out to be thinner and more homogeneous than the SEI formed in the PC-EC electrolyte.
When analyzing the processes of sodium insertion into materials based on oxides, it should be taken into account that, in these cases, there are at least two mechanisms for the appearance of the irreversible capacity of the initial cycles. First, these are the processes of reduction in electrolyte components, leading, in particular, to the formation of SEI. Secondly, these are the processes of reduction in oxides themselves. For example, for electrodes based on SnO2, during the first charge at potentials of 3–0.8 V, sodium ions are introduced into the structure of tin dioxide. At potentials more negative than 0.8 V, the sodium ion interacts with tin dioxide to form tin and Na2O nanoparticles. Finally, at potentials more negative than 0.1 V, the formation of the NaxSn alloy is observed. The authors of [269] have shown that during the electrochemical reduction in thin layers of highly porous SnO2 on a copper substrate, firstly, Na2O and tin are formed according to the equation, and then a series of intermetallic compounds NaSn3, α-NaSn, Na9Sn4 and Na15Sn4 appear. During the discharge process, the latter again turns into the tin. The same remark applies to the insertion of sodium into materials based on chalcogenides.
SnO2 + 4Na+ + 4e → Sn + 2Na2O
Among other oxide materials on which SEI formation processes were studied, mention should be made of oxides of cobalt [186,207,208,209,210,211,212], antimony [214,215,216,217], niobium [218], and nickel [184,213]. According to the data of [214], SEI on Sb2O4 in an electrolyte based on EC-DMC are formed at potentials of about 0.43 V. The authors of [215] report that on a SbOx composite with reduced graphene oxide in an EC-PC-based electrolyte, SEI at the first cycle is formed at potentials of about 0.46 V.
Of the chalcogenides proposed as active anode materials in sodium-ion batteries, sulfides (FeS [230], Fe3S4 [237], Bi2S3 [229], MoS2 [231,241], WS2 [234], SnS2 [235,243], Sb2S3 [242], Ni3S2 [236]) and selenides (SnSe2 [232], CoSe [238], ZnSe [239], MoSe0.85S0.15 [240], FeSe2 [270]) arouse keen interest.
The work [229] confirmed the sodium insertion mechanism described above, including the stage of bismuth sulfide reduction with the formation of bismuth metal and Na2S. This work and [231] also confirmed the trivial conclusion that on electrodes based on Bi2S3 nanorods in an electrolyte based on EC:DEC loose unstable SEIs are formed, which are the main reason for capacity fading, while thin dense SEIs are obtained in a DME-based electrolyte (Figure 11a). In addition, the cycling of Bi2S3 nanorods in a DME-based electrolyte was found to result in surface coarsening, which increases the irreversible capacity. Doping Bi2S3 with iron and depositing a thin carbon coating on the nanorods prevent etching and radically change the character of SEIs, enriching them with sodium fluoride and reducing their impedance.
In [230], an unusual binder, specifically sodium polyacrylate, was used for FeS-based electrodes. In this case, in a solution of 1 M NaCF3SO3 in diethylene glycol dimethyl ether as an electrolyte, ductile SEIs were formed at potentials of about 0.8 V. Some amounts of sodium polyacrylate were found in the organic components of SEI, and here the organic components were localized in surface layers of the SEI, and mineral ones in the inner layers, too.
The example of chalcogenide-based electrodes shows the influence of the structure of electrodes on the properties of SEI. For instance, composite electrodes with a sandwich-like structure consisting of SnSe2 and reduced graphene oxide layers were studied in [232], with the content of reduced graphene oxide being as low as 7.3%. According to the authors, it was this electrode structure that ensured the formation of thin reliable SEIs as a result of the abundance of Sn–O–C bonds. It was shown in [233] that for Sb2Te3 composites with carbon nanotubes, the thickness and continuity of the SEI depend on the ratio of these components, and, in the optimal case (10% carbon nanotubes), the thickness of the SEI formed in 1 M EC:PC with the addition of FEC is 19 nm. For plain Sb2Te3 (without carbon nanotubes), the SEI thickness was 67 nm. [234] states that it is the thin carbon coating on the tungsten disulfide particles that ensures the formation of thin, high-quality SEIs. The original globule-like FeSe2/graphene structure (B-FeSe2/G) is described in [270]. Here graphene layers wrap the surfaces of FeSe2 particles and stretch into the interior of these particles. Such structures display higher electrochemical performance than particulate P-FeSe2 (Figure 11b).
An interesting example of the use of tin sulfides is the work [139]. Here, the active material of the negative electrode is activated carbon, on the surface of which particles the thinnest layer of SnS and SnS2 nanosheets, which play the role of an artificial SEI, are deposited.
In [263], data are presented on the formation of reliable SEIs on electrodes made of Sb2S composite with graphene in an electrolyte based on an ionic liquid (1 M NaFSI in PMP:FSI).
Phosphorus and some phosphides are promising materials for the negative electrode of sodium-ion batteries [271]. As a rule, phosphorus is used as composites with various forms of carbon. SEIs are also formed on such electrodes during cathodic polarization, and the mechanism of this process does not have any specific features; however, the properties of SEIs depend on the composition of the composite, i.e., from the ratio P:C [247]. In [250], a “core-shell” structure was proposed, where the core was composed of a composite of red phosphorus with ball-milled carbon nanotubes, and the shell was a thin (less than 10 nm) polydopamine film playing the role of SEI. In [249], it is proposed to deposit a thin coating of amorphous TiO2 on particles of a red phosphorus composite with milled carbon nanotubes. The presence of such a coating, according to the authors, provides the formation of thin continuous ductile SEI enriched with NaF and has a reduced resistance. In [251], it is proposed to apply a thin coating of nickel to red phosphorus particles.
Of the phosphides proposed as negative electrodes, the most popular are Sn4P3 [244,245,272], as well as VP2 [248], CoSi3P3, and FeSi4P4 [246]. It was shown in [244] that the combined addition of FEC + TMSP provides the same reliable SEI on the surface of the phosphide as on the surface of hard carbon. In [248], high-quality SEIs were obtained on a VP2 electrode in an ionic liquid-based electrolyte.
The properties of SEIs formed on such negative electrodes as sodium titanate have been studied in [252,253,254,255,273]. In addition, information is provided on SEI on Na3(TiOPO4)2F [256], NaTiOPO4 [182], ZnSnO3 composite with carbon [257], Bi2MoO6 [259], and CuCrP2S6 [260].

5. CEI on Cathode Materials

Passive films on the materials of positive electrodes of sodium-ion batteries have been studied much less than SEI on negative electrodes. Moreover, since a wide variety of materials have been proposed as functional cathode materials, the characteristics of CEI on each specific material have been studied at best in one or two papers. The very existence of passive films as products of electrolyte oxidation was confirmed by direct electron microscopic studies in [274] for Na0.70Mn0.80Co0.15Zr0.05O2 (it is noted that it was zirconium doping that ensured the formation of dense thin CEI), in [275] for (Ni0.4Co0.4Mn0.2)3O4 and in [276] for Na0.67Mn0.625Fe0.25Ni0.125O2.
In [277], reliable conclusions about the properties of CEIs formed on Na2FexFe1−x(SO4)2(OH)x during the first anodic polarization in 1 M NaClO4 in an EC:DMC:EMC mixture were made based on the results of electrochemical impedance spectroscopy. In an equivalent circuit, CEI is described by a parallel combination of a constant phase element QCEI and resistance RCEI. The last element depends on the potential and, at potentials from 1.5 to 3.9 V, is approximately 1.5 Ohm (for a particular cell), and as the potential increases to 4.5 V, it increases almost linearly to 7 Ohm as a result of an increase in the CEI thickness.
It was shown in [278] that stable, reliable CEIs are formed on NaNi0.68Mn0.22Co0.1O2 in an electrolyte consisting of 1.5 M NaFSI in a mixture of DEC with tris (2,2,2-trifluoroethyl) phosphate (TFP). We deem it possible to assume that the stability of such CEIs is determined by the optimum solvating capacity of said solvent.
Like SEI, the properties of CEI depend on the presence of an FEC additive in the electrolyte. According to [279], the addition of FEC to a solution of NaClO4 in PC inhibits the oxidation of PC on the surface of NaCo0.7Mn0.3O2 and promotes the formation of CEI enriched in NaF.
The authors of [280] found that the behavior of Prussian blue cathodes depends on the content of uncoordinated water molecules (so-called zeolite water). The content of zeolite water, in turn, depends on the Prussian blue synthesis temperature, the optimal temperature being 80 °C. CEI enriched with Na2CO3 are deposited on the material synthesized in this mode, which ensures stable cycling of the positive electrodes.
It was shown in [281,282] that the addition of ionic liquids to PC-based electrolytes leads to a noticeable improvement in CEI on Na3V2(PO4)3–carbon composites.

6. Conclusions

Characteristics of sodium-ion batteries largely depend on the presence and properties of passive films formed on the surface of active materials of negative and positive electrodes. Passive films on negative electrodes result from the reduction in electrolyte components (solvent and salt anion). They have the properties of a solid electrolyte with sodium ion conductivity and are insulators in terms of electronic conductivity. Their thickness is a few (or dozen) nanometers. Usually, they are called SEI—solid electrolyte interphase. The formation of SEI is associated with the consumption of a certain charge, which is an irreversible capacity. Passive films on the surface of positive electrodes (CEI—cathode electrolyte interphase) arise as a result of electrolyte oxidation.
The formation process and properties of SEI depend both on the nature of the electrode material and on the nature of the electrolyte and the conditions for its reduction (including temperature, current density, etc.). In some cases, thin dense SEIs of uniform thickness with high ionic conductivity are formed, which provide long-term efficient cycling of negative electrodes with low-capacity fading. In other cases, SEIs are thick, unevenly distributed over the surface, and have high Ohmic resistance, resulting in poor cyclability. In many cases, electrolyte additives have a beneficial effect on SEI properties. The most effective and popular additive is fluoroethylene carbonate. The most studied are SEI on electrodes based on hard carbon, other carbon materials, some metals, oxides, chalcogenides, and phosphides. In Table 1, some information on SEI composition on various electrodes in various electrolytes, as well as on irreversible capacity at initial cycles, is summarized.
The analysis of Table 1 shows that SEI principal composition on various electrodes is, by and large, almost the same. The main variations in SEI composition, e.g., the presence of Na2SO4, Na2S, and SiF, are associated with features of special electrolytes. It is clear that the composition of inorganic components of SEIs is determined by the nature of the salt’s anion, whereas the composition of organic components is associated mainly with the nature of solvents.
CEI studies are few and far between. However, the very existence of CEI has been proven by electron microscopy studies, and the beneficial effect of fluoroethylene carbonate additives on the characteristics of CEI has also been confirmed.

Author Contributions

T.L.K.—conceptualization, analysis of literature, writing of original draft; A.M.S.—conceptualization, analysis of literature, writing and editing of the 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 21-13-00160.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

SEIsolid electrolyte interphase
CEIcathode electrolyte interphase
ECethylene carbonate
PCpropylene carbonate
DECdiethyl carbonate
DMCdimethyl carbonate
RCH2ONa, RONasodium alkoxides
NaFSIsodium bisfluorosulfonylimide
DME1,2-dimethoxyethane
BTFEbis(2,2,2-trifluoroethyl) ether
FECfluoroethylene carbonate
PVdF–HFPpolyvinylidene fluoride with hexafluoropropylene
Qrevreversible capacity
Qirrirreversible capacity
NaTFSIsodium bis(trifluoromethanesulfonyl)imide
NaFTFSIsodium fluorosulfonyl-(trifluoromethanesulfonyl)imide
XPSphotoelectron spectroscopy
ROCO3Nasodium alkyl carbonates
VCvinylene carbonate
DFTdensity functional theory
TMSPtris(trimethylsilyl)phosphite
NaODFBsodium-difluoro(oxalate)borate
ROSO2Nasodium alkyl sulfates
RSO3Nasodium alkyl and sulfites
DEGDMEdiethylene glycol dimethyl ether, diglyme
TEGDMEtetraethylene glycol dimethyl ether
PMP–FSIN-propyl-N-methylpyrrolidinium bis(fluorosulfonyl)imide
PMP-TFSIN-propyl-N-methylpyrrolidinium bis(trifluoromethanesulfonyl)imide
TFPtris (2,2,2-trifluoroethyl) phosphate

References

  1. Balbuena, P.B.; Wang, Y. (Eds.) Lithium-Ion Batteries: Solid-Electrolyte Interface; Imperial College Press: London, UK, 2004; 407p. [Google Scholar]
  2. Peled, E. The Electrochemical Behavior of Alkali and Alkaline Earth Metals in Nonaqueous Battery Systems—The Solid Electrolyte Interphase Model. J. Electrochem. Soc. 1979, 126, 2047–2051. [Google Scholar] [CrossRef]
  3. Peled, E.; Golodnitsky, D.; Ardel, G. Advanced Model for Solid Electrolyte Interphase Electrodes in Liquid and Polymer Electrolytes. J. Electrochem. Soc. 1997, 144, L208–L210. [Google Scholar] [CrossRef]
  4. Qian, Y.; Hu, S.; Zou, X.; Deng, Z.; Xu, Y.; Cao, Z.; Kang, Y.; Deng, Y.; Shi, Q.; Xu, K.; et al. How electrolyte additives work in Li-ion batteries. Energy Storage Mater. 2019, 20, 208–215. [Google Scholar] [CrossRef]
  5. Haregewoin, A.M.; Wotango, A.S.; Hwang, B.-J. Electrolyte additives for lithium ion battery electrodes: Progress and perspectives. Energy Environ. Sci. 2016, 9, 1955–1988. [Google Scholar] [CrossRef]
  6. Huang, Y.; Zhao, L.; Li, L.; Xie, M.; Wu, F.; Chen, R. Electrolytes and Electrolyte/Electrode Interfaces in Sodium-Ion Batteries: From Scientific Research to Practical Application. Adv. Mater. 2019, 31, 1808393. [Google Scholar] [CrossRef]
  7. Eshetu, G.G.; Elia, G.A.; Armand, M.; Forsyth, M.; Komaba, S.; Rojo, T.; Passerini, S. Electrolytes and Interphases in Sodium-Based Rechargeable Batteries: Recent Advances and Perspectives. Adv. Energy Mater. 2020, 10, 2000093. [Google Scholar] [CrossRef] [Green Version]
  8. Wang, E.; Niu, Y.; Yin, Y.-X.; Guo, Y.-G. Manipulating Electrode/Electrolyte Interphases of Sodium-Ion Batteries: Strategies and Perspectives. ACS Mater. Lett. 2021, 3, 18–41. [Google Scholar] [CrossRef]
  9. Li, Y.; Liu, M.; Feng, X.; Li, Y.; Wu, F.; Bai, Y.; Wu, C. How Can the Electrode Influence the Formation of the Solid Electrolyte Interface? ACS Energy Lett. 2021, 6, 3307–3320. [Google Scholar] [CrossRef]
  10. Bao, C.; Wang, B.; Liu, P.; Wu, H.; Zhou, Y.; Wang, D.; Liu, H.; Dou, S. Solid Electrolyte Interphases on Sodium Metal Anodes. Adv. Funct. Mater. 2020, 30, 2004891. [Google Scholar] [CrossRef]
  11. Zhang, J.; Gai, J.; Song, K.; Chen, W. Advances in electrode/electrolyte interphase for sodium-ion batteries from half cells to full cells. Cell Rep. Phys. Sci. 2022, 3, 100868. [Google Scholar] [CrossRef]
  12. Lee, B.; Paek, E.; Mitlin, D.; Lee, S.W. Sodium metal anodes: Emerging solutions to dendrite growth. Chem. Rev. 2019, 119, 5416–5460. [Google Scholar] [CrossRef]
  13. Ma, L.; Cui, J.; Yao, S.; Liu, X.; Luo, Y.; Shen, X.; Kim, J.-K. Dendrite-Free Lithium Metal and Sodium Metal Batteries. Energy Storage Mater. 2020, 27, 522–554. [Google Scholar] [CrossRef]
  14. Fang, L.; Bahlawane, N.; Sun, W.; Pan, H.; Xu, B.B.; Yan, M.; Jiang, Y. Conversion-Alloying Anode Materials for Sodium Ion Batteries. Small 2021, 17, 2101137. [Google Scholar] [CrossRef] [PubMed]
  15. Liu, Q.; Xu, R.; Mu, D.; Tan, G.; Gao, H.; Li, N.; Chen, R.; Wu, F. Progress in electrolyte and interface of hard carbon and graphite anode for sodium-ion battery. Carbon Energy 2022, 4, 458–479. [Google Scholar] [CrossRef]
  16. Yadegari, H.; Sun, Q.; Sun, S. Sodium-Oxygen Batteries: A Comparative Review from Chemical and Electrochemical Fundamentals to Future Perspective. Adv. Mater. 2016, 28, 7065–7093. [Google Scholar] [CrossRef]
  17. Manthiram, A.; Yu, X. Ambient Temperature Sodium–Sulfur Batteries. Small 2015, 11, 2108–2114. [Google Scholar] [CrossRef]
  18. Hu, X.; Sun, J.; Li, Z.; Zhao, Q.; Chen, C.; Chen, J. Rechargeable Room-Temperature Na–CO2 Batteries. Angew. Chem. Int. Ed. 2016, 55, 6482–6486. [Google Scholar] [CrossRef]
  19. Rodriguez, R.; Loeffler, K.E.; Nathan, S.S.; Sheavly, J.K.; Dolocan, A.; Heller, A.; Mullins, C.B. In Situ Optical Imaging of Sodium Electrodeposition: Effects of Fluoroethylene Carbonate. ACS Energy Lett. 2017, 2, 2051–2057. [Google Scholar] [CrossRef]
  20. Zhang, J.; Wang, D.-W.; Lv, W.; Qin, L.; Niu, S.; Zhang, S.; Cao, T.; Kang, F.; Yang, Q.-H. Ethers Illume Sodium-Based Battery Chemistry: Uniqueness; Surprise; and Challenges. Adv. Energy Mater. 2018, 8, 1801361. [Google Scholar] [CrossRef]
  21. Mandl, M.; Becherer, J.; Kramer, D.; Mönig, R.; Diemant, T.; Behm, R.J.; Hahn, M.; Böse, O.; Danzer, M.A. Sodium metal anodes: Deposition and dissolution behaviour and SEI formation. Electrochim. Acta 2020, 354, 136698. [Google Scholar] [CrossRef]
  22. She, Z.W.; Sun, J.; Sun, Y.; Cui, Y. A Highly Reversible Room-Temperature Sodium Metal Anode. ACS Cent. Sci. 2015, 1, 449–455. [Google Scholar] [CrossRef]
  23. Dugas, R.; Ponrouch, A.; Gachot, G.; David, R.; Palacin, M.R.; Tarascon, J.M. Na Reactivity toward Carbonate-Based Electrolytes: The Effect of FEC as Additive. J. Electrochem. Soc. 2016, 163, A2333–A2339. [Google Scholar] [CrossRef] [Green Version]
  24. Mogensen, R.; Brandell, D.; Younesi, R. Solubility of the Solid Electrolyte Interphase (SEI) in Sodium Ion Batteries. ACS Energy Lett. 2016, 1, 1173–1178. [Google Scholar] [CrossRef]
  25. Iermakova, D.I.; Dugas, R.; Palacín, M.R.; Ponrouch, A. On the Comparative Stability of Li and Na Metal Anode Interfaces in Conventional Alkyl Carbonate Electrolytes. J. Electrochem. Soc. 2015, 162, A7060–A7066. [Google Scholar] [CrossRef]
  26. Rudola, A.; Aurbach, D.; Balaya, P. A new phenomenon in sodium batteries: Voltage step due to solvent interaction. Electrochem. Commun. 2014, 46, 56–59. [Google Scholar] [CrossRef]
  27. Cao, Y.; Xiao, L.; Wang, W.; Choi, D.; Nie, Z.; Yu, J.; Saraf, L.V.; Yang, Z.; Liu, J. Reversible Sodium Ion Insertion in Single Crystalline Manganese Oxide Nanowires with Long Cycle Life. Adv. Mater. 2011, 23, 3155–3160. [Google Scholar] [CrossRef]
  28. Chen, J.; Huang, Z.; Wang, C.; Porter, S.; Wang, B.; Lie, W.; Liu, H.K. Sodium-difluoro(oxalato)borate (NaDFOB): A new electrolyte salt for Na-ion batteries. Chem. Commun. 2015, 51, 9809–9812. [Google Scholar] [CrossRef] [Green Version]
  29. Gao, L.; Chen, J.; Liu, Y.; Yamauchi, Y.; Huang, Z.; Kong, X. Revealing the chemistry of an anode-passivating electrolyte salt for high rate and stable sodium metal batteries. J. Mater. Chem. A 2018, 6, 12012–12017. [Google Scholar] [CrossRef]
  30. Lee, J.; Lee, Y.; Lee, L.; Lee, S.-M.; Choi, J.-H.; Kim, H.; Kwon, M.-S.; Kang, K.; Lee, K.T.; Choi, N.-S. Ultraconcentrated Sodium Bis(fluorosulfonyl)imide-Based Electrolytes for High-Performance Sodium Metal Batteries. ACS Appl. Mater. Interfaces 2017, 9, 3723–3732. [Google Scholar] [CrossRef]
  31. Cao, R.; Mishra, K.; Li, X.; Qian, J.; Engelhard, M.H.; Bowden, M.E.; Han, K.S.; Mueller, K.T.; Henderson, W.A.; Zhang, J.-G. Enabling room temperature sodium metal batteries. Nano Energy 2016, 30, 825–830. [Google Scholar] [CrossRef]
  32. Zheng, J.; Chen, S.; Zhao, W.; Song, J.; Engelhard, M.; Zhang, J.-G. Extremely Stable Sodium Metal Batteries Enabled by Localized High Concentration Electrolytes. ACS Energy Lett. 2018, 3, 315–321. [Google Scholar] [CrossRef]
  33. Basile, A.; Makhlooghiazad, F.; Yunis, R.; MacFarlane, D.R.; Forsytha, M.; Howlett, P.C. Extensive sodium metal plating and stripping in a highly concentrated inorganic-organic ionic liquid electrolyte via surface pretreatment. ChemElectroChem 2017, 4, 986–991. [Google Scholar] [CrossRef]
  34. Ruiz-Martínez, D.; Kovacs, A.; Gómez, R. Development of novel inorganic electrolytes for room temperature rechargeable sodium metal batteries. Energy Environ. Sci. 2017, 10, 1936–1941. [Google Scholar] [CrossRef]
  35. Badoz-Lambling, J.; Bardin, M.; Bernard, C.; Fahys, B.; Herlem, M.; Thiebault, A.; Robert, G. New Battery Electrolytes for Low and High Temperatures: Liquid and Solid Ammoniates for High Energy Batteries I. Sodium Iodide and Lithium Perchlorate Ammoniates. J. Electrochem. Soc. 1988, 135, 578–591. [Google Scholar] [CrossRef]
  36. Song, J.; Jeong, G.; Lee, A.; Park, J.H.; Kim, H.; Kim, Y.-J. Dendrite-Free Polygonal Sodium Deposition with Excellent Interfacial Stability in a NaAlCl4–2SO2 Inorganic Electrolyte. ACS Appl. Mater. Interfaces. 2015, 7, 27206–27214. [Google Scholar] [CrossRef]
  37. Komaba, S.; Ishikawa, T.; Yabuuchi, N.; Murata, W.; Ito, A.; Ohsawa, Y. Fluorinated Ethylene Carbonate as Electrolyte Additive for Rechargeable Na Batteries. ACS Appl. Mater. Interfaces 2011, 3, 4165–4168. [Google Scholar] [CrossRef]
  38. Xu, C.; Lindgren, F.; Philippe, B.; Gorgoi, M.; Björefors, F.; Edström, K.; Gustafsson, T. Improved Performance of the Silicon Anode for Li-Ion Batteries: Understanding the Surface Modification Mechanism of Fluoroethylene Carbonate as an Effective Electrolyte Additive. Chem. Mater. 2015, 27, 2591–2599. [Google Scholar] [CrossRef]
  39. Han, M.; Zhu, C.; Ma, T.; Pan, Z.; Tao, Z.; Chen, J. In situ atomic force microscopy study of nano–micro sodium deposition in ester-based electrolytes. Chem. Commun. 2018, 54, 2381–2384. [Google Scholar] [CrossRef]
  40. Shiraz, M.H.A.; Zhao, P.; Liu, J. High-performance sodium–selenium batteries enabled by microporous carbon/selenium cathode and fluoroethylene carbonate electrolyte additive. J. Power Sources 2020, 453, 227855. [Google Scholar] [CrossRef]
  41. Wang, Q.; Zhao, C.; Lv, X.; Lu, Y.; Lin, K.; Zhang, S.; Kang, F.; Hu, Y.-S.; Li, B. Stabilizing a sodium-metal battery with the synergy effects of a sodiophilic matrix and fluorine-rich interface. J. Mater. Chem. A 2019, 7, 24857–24867. [Google Scholar] [CrossRef]
  42. Lee, Y.; Lee, J.; Lee, J.; Kim, K.; Cha, A.; Kang, S.; Wi, T.; Kang, S.J.; Lee, H.-W.; Choi, N.-S. Fluoroethylene Carbonate-Based Electrolyte with 1 M Sodium Bis(fluorosulfonyl)imide Enables High-Performance Sodium Metal Electrodes. ACS Appl. Mater. Interfaces 2018, 10, 15270–15280. [Google Scholar] [CrossRef] [PubMed]
  43. Yi, Q.; Lu, Y.; Sun, X.; Zhang, H.; Yu, H.; Sun, C. Fluorinated Ether Based Electrolyte Enabling Sodium-Metal Batteries with Exceptional Cycling Stability. ACS Appl. Mater. Interfaces 2019, 11, 46965–46972. [Google Scholar] [CrossRef] [PubMed]
  44. Liu, Q.; Mu, D.; Wu, B.; Wang, L.; Gai, L.; Wu, F. Density Functional Theory Research into the Reduction Mechanism for the Solvent/Additive in a Sodium-Ion Battery. ChemSusChem 2016, 10, 786–796. [Google Scholar] [CrossRef]
  45. Åvall, G.; Mindemark, J.; Brandell, D.; Johansson, P. Sodium-Ion Battery Electrolytes: Modeling and Simulations. Adv. Energy Mater. 2018, 8, 1703036. [Google Scholar] [CrossRef]
  46. Takenaka, N.; Sakai, H.; Suzuki, Y.; Uppula, P.; Nagaoka, M. A Computational Chemical Insight into Microscopic Additive Effect on Solid Electrolyte Interphase Film Formation in Sodium-Ion Batteries: Suppression of Unstable Film Growth by Intact Fluoroethylene Carbonate. J. Phys. Chem. C 2015, 119, 18046–18055. [Google Scholar] [CrossRef]
  47. Kumar, H.; Detsi, E.; Abraham, D.P.; Sheno, V.B. Fundamental Mechanisms of Solvent Decomposition Involved in Solid-Electrolyte Interphase Formation in Sodium Ion Batteries. Chem. Mater. 2016, 28, 8930–8941. [Google Scholar] [CrossRef]
  48. Wang, H.; Wang, C.; Matios, E.; Li, W. Facile Stabilization of Sodium Metal Anode with Additives: Unexpected Key Role of Sodium Polysulfide and Adverse Effect of Sodium Nitrate. Angew. Chem. Int. Ed. 2018, 57, 7734–7737. [Google Scholar] [CrossRef]
  49. Shi, Q.; Zhong, Y.; Wu, M.; Wang, H.; Wang, H. High performance sodium metal anodes enabled by a bifunctional potassium salt. Angew. Chem. Int. Ed. 2018, 57, 9069–9072. [Google Scholar] [CrossRef]
  50. Fang, W.; Jiang, H.; Zheng, Y.; Zheng, H.; Liang, X.; Sun, Y.; Chen, C.; Xiang, H. A bilayer interface formed in high concentration electrolyte with SbF3 additive for long-cycle and high-rate sodium metal battery. J. Power Sources 2020, 455, 227956. [Google Scholar] [CrossRef]
  51. Wang, S.; Cai, W.; Sun, Z.; Huang, F.; Jie, Y.; Liu, Y.; Chen, Y.; Peng, B.; Cao, R.; Zhang, G.; et al. Stable cycling of Na metal anodes in a carbonate electrolyte. Chem. Commun. 2019, 55, 14375–14378. [Google Scholar] [CrossRef]
  52. Doi, K.; Yamada, Y.; Okoshi, M.; Ono, J.; Chou, C.-P.; Nakai, H.; Yamada, A. Reversible Sodium Metal Electrodes: Is Fluorine an Essential Interphasial Component? Angew. Chem. Int. Ed. 2019, 58, 8024–8028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Zheng, X.; Fu, H.; Hu, C.; Xu, H.; Huang, Y.; Wen, J.; Sun, H.; Luo, W.; Huang, Y. Toward a Stable Sodium Metal Anode in Carbonate Electrolyte: A Compact; Inorganic Alloy Interface. J. Phys. Chem. Lett. 2019, 10, 707–714. [Google Scholar] [CrossRef] [PubMed]
  54. Tian, H.; She, Z.W.; Yan, K.; Fu, Z.; Tang, P.; Lu, Y.; Zhang, R.; Legut, D.; Cui, Y.; Zhang, Q. Theoretical Investigation of 2D Layered Materials as Protective Films for Lithium and Sodium Metal Anodes. Adv. Energy Mater. 2017, 7, 1602528. [Google Scholar] [CrossRef]
  55. Wang, H.; Wang, C.; Matios, E.; Li, W. Critical Role of Ultrathin Graphene Films with Tunable Thickness in Enabling Highly Stable Sodium Metal Anodes. Nano Lett. 2017, 17, 6808–6815. [Google Scholar] [CrossRef]
  56. Kim, Y.-J.; Lee, H.; Noh, H.; Lee, J.; Kim, S.; Ryou, M.-H.; Lee, Y.M.; Kim, H.-T. Enhancing the Cycling Stability of Sodium Metal Electrode by Building an Inorganic/Organic Composite Protective Layer. ACS Appl. Mater. Interfaces 2017, 9, 6000–6006. [Google Scholar] [CrossRef]
  57. Zhao, Y.; Goncharova, L.V.; Zhang, Q.; Kaghazchi, P.; Sun, Q.; Lushington, A.; Wang, B.; Li, R.; Sun, X. Inorganic–Organic Coating via Molecular Layer Deposition Enables Long Life Sodium Metal Anode. Nano Lett. 2017, 17, 5653–5659. [Google Scholar] [CrossRef]
  58. Zhang, D.; Li, B.; Wang, S.; Yang, S. Simultaneous formation of artificial SEI film and 3D host for stable metallic sodium anodes. ACS Appl. Mater. Interfaces 2017, 9, 40265–40272. [Google Scholar] [CrossRef]
  59. Wang, A.; Hu, X.; Tang, H.; Zhang, C.; Liu, S.; Yang, Y.-W.; Yang, Q.-H.; Luo, J. Processable and Moldable Sodium Metal Anodes. Angew. Chem. Int. Ed. 2017, 56, 11921–11926. [Google Scholar] [CrossRef]
  60. Stevens, D.A.; Dahn, J.R. High Capacity Anode Materials for Rechargeable Sodium-Ion Batteries. J. Electrochem. Soc. 2000, 147, 1271–1273. [Google Scholar] [CrossRef]
  61. Barker, J.; Saidi, M.Y.; Swoyer, J.L. A Sodium-Ion Cell Based on the Fluorophosphate Compound NaVPO4F. Electrochem. Solid-State Lett. 2003, 6, A1–A4. [Google Scholar] [CrossRef]
  62. Irisarri, E.; Ponrouch, A.; Palacin, M.R. Review—Hard Carbon Negative Electrode Materials for Sodium-Ion Batteries. J. Electrochem. Soc. 2015, 162, A2476–A2482. [Google Scholar] [CrossRef]
  63. Ponrouch, A.; Goñi, A.R.; Palacín, M.R. High capacity hard carbon anodes for sodium ion batteries in additive free electrolyte. Electrochem. Commun. 2013, 27, 85–88. [Google Scholar] [CrossRef]
  64. Alptekin, H.; Au, H.; Olsson, E.; Cottom, J.; Jensen, A.C.; Headen, T.F.; Cai, Q.; Drew, A.J.; Crespo-Ribadeneyra, M.; Titirici, M.-M. Elucidation of the Solid Electrolyte Interphase Formation Mechanism in Micro-Mesoporous Hard-Carbon Anodes. Adv. Mater. Interfaces 2022, 9, 2101267. [Google Scholar] [CrossRef]
  65. Alptekin, H.; Au, H.; Jensen, A.; Olsson, E.; Goktas, M.; Headen, T.F.; Adelhelm, P.; Cai, Q.; Drew, A.J.; Titirici, M.-M. Sodium Storage Mechanism Investigations Through Structural Changes in Hard Carbons. ACS Appl. Energy Mater. 2020, 3, 9918–9927. [Google Scholar] [CrossRef]
  66. Komaba, S.; Murata, W.; Ishikawa, T.; Yabuuchi, N.; Ozeki, T.; Nakayama, T.; Ogata, A.; Gotoh, K.; Fujiwara, K. Electrochemical Na Insertion and Solid Electrolyte Interphase for Hard-Carbon Electrodes and Application to Na-Ion Batteries. Adv. Funct. Mater. 2011, 21, 3859–3867. [Google Scholar] [CrossRef]
  67. Muñoz-Márquez, M.A.; Zarrabeitia, M.; Passerini, S.; Rojo, T. Structure; Composition; Transport Properties; and Electrochemical Performance of the Electrode-Electrolyte Interphase in Non-Aqueous Na-Ion Batteries. Adv. Mater. Interfaces 2022, 9, 2101773. [Google Scholar] [CrossRef]
  68. Song, X.; Zhou, X.; Zhou, Y.; Deng, Y.; Meng, T.; Gao, A.; Nan, J.; Shu, D.; Yi, F. Reaction Mechanisms of Sodium-Ion Batteries under Various Charge and Discharge Conditions in a Three-Electrode Setup. ChemElectroChem 2018, 5, 2475–2481. [Google Scholar] [CrossRef]
  69. Eshetu, G.G.; Diemant, T.; Hekmatfar, M.; Grugeon, S.; Behm, R.J.; Laruelle, S.; Armand, M.; Passerini, S. Impact of the electrolyte salt anion on the solid electrolyte interphase formation in sodium ion batteries. Nano Energy 2019, 55, 327–340. [Google Scholar] [CrossRef]
  70. Fondard, J.; Irisarri, E.; Courrèges, C.; Palacin, M.R.; Ponrouch, A.; Dedryvère, R. SEI Composition on Hard Carbon in Na-Ion Batteries After Long Cycling: Influence of Salts (NaPF6; NaTFSI) and Additives (FEC; DMCF). J. Electrochem. Soc. 2020, 167, 070526. [Google Scholar] [CrossRef]
  71. Morikawa, Y.; Yamada, Y.; Doi, K.; Nishimura, S.; Yamada, A. Reversible and High-rate Hard Carbon Negative Electrodes in a Fluorine-free Sodium-salt Electrolyte. Electrochemistry 2020, 88, 151–156. [Google Scholar] [CrossRef]
  72. Che, H.; Chen, S.; Xie, Y.; Wang, H.; Amine, K.; Liao, X.-Z.; Ma, Z.-F. Electrolyte design strategies and research progress for room-temperature sodium-ion batteries. Energy Environ. Sci. 2017, 10, 1075–1101. [Google Scholar] [CrossRef]
  73. Bommier, C.; Ji, X. Electrolytes; SEI Formation; and Binders: A Review of Nonelectrode Factors for Sodium-Ion Battery Anodes. Small 2018, 14, 1703576. [Google Scholar] [CrossRef]
  74. Du, K.; Wang, C.; Subasinghe, L.U.; Gajella, S.R.; Law, M.; Rudola, A.; Balaya, P. A comprehensive study on the electrolyte; anode and cathode for developing commercial type non-flammable sodium-ion battery. Energy Storage Mater. 2020, 29, 287–299. [Google Scholar] [CrossRef]
  75. Zhao, D.; Lu, H.; Li, S.; Wang, P.; Fan, X. Boosting the cycling stability of hard carbon with NaODFB-based electrolyte at high temperature. Mater. Today Chem. 2022, 24, 100866. [Google Scholar] [CrossRef]
  76. Eshetu, G.G.; Grugeon, S.; Kim, H.; Jeong, S.; Wu, L.; Gachot, G.; Laruelle, S.; Armand, M.; Passerini, S. Comprehensive Insights into the Reactivity of Electrolytes Based on Sodium Ions. ChemSusChem 2016, 9, 462–471. [Google Scholar] [CrossRef]
  77. Ponrouch, A.; Dedryvère, R.; Monti, D.; Demet, A.E.; Mba, J.M.A.; Croguennec, L.; Masquelier, C.; Johansson, P.; Palacín, M.R. Towards high energy density sodium ion batteries through electrolyte optimization. Energy Environ. Sci. 2013, 6, 2361–2369. [Google Scholar] [CrossRef]
  78. Carboni, M.; Manzi, J.; Armstrong, A.R.; Billaud, J.; Brutti, S.; Younesi, R. Analysis of the Solid Electrolyte Interphase on Hard Carbon Electrodes in Sodium-Ion Batteries. ChemElectroChem 2019, 6, 1745–1753. [Google Scholar] [CrossRef] [Green Version]
  79. Ponrouch, A.; Monti, D.; Boschin, A.; Steen, B.; Johansson, P.; Palacín, M.R. Non-aqueous electrolytes for sodium-ion batteries. J. Mater. Chem. A 2015, 3, 22–42. [Google Scholar] [CrossRef]
  80. Bhawana, K.; Roy, A.; Chakrabarty, N.; Gautam, M.; Dutta, D.P.; Mitra, S. Sodium-ion batteries: Chemistry of biomass derived disordered carbon in carbonate and ether-based electrolytes. Electrochim. Acta 2022, 425, 140744. [Google Scholar] [CrossRef]
  81. Dahbi, M.; Nakano, T.; Yabuuchi, N.; Fujimura, S.; Chihara, K.; Son, J.-Y.; Cui, Y.-T.; Oji, H.; Komaba, S. Effect of Hexafluorophosphate and Fluoroethylene Carbonate on Electrochemical Performance and Surface Layer of Hard Carbon for Sodium-Ion Batteries. ChemElectroChem 2016, 3, 1856–1867. [Google Scholar] [CrossRef]
  82. Bai, P.; Han, X.; He, Y.; Xiong, P.; Zhao, Y.; Sun, J.; Xu, Y. Solid electrolyte interphase manipulation towards highly stable hard carbon anodes for sodium ion batteries. Energy Storage Mater. 2020, 25, 324–333. [Google Scholar] [CrossRef]
  83. Bouibes, A.; Takenaka, N.; Fujie, T.; Kubota, K.; Komaba, S.; Nagaoka, M. Concentration Effect of Fluoroethylene Carbonate on Formation of Solid Electrolyte Interphase Layer in Sodium-Ion Batteries. ACS Appl. Mater. Interfaces 2018, 10, 28525–28532. [Google Scholar] [CrossRef]
  84. Ponrouch, A.; Marchante, E.; Courty, M.; Tarascon, J.-M.; Palacín, M.R. In search of an optimized electrolyte for Na-ion batteries. Energy Environ. Sci. 2012, 5, 8572–8583. [Google Scholar] [CrossRef]
  85. Purushotham, U.; Takenaka, N.; Nagaoka, M. Additive Effect of Fluoroethylene and Difluoroethylene Carbonates for the Solid Electrolyte Interphase Film Formation in Sodium-Ion Batteries: A Quantum Chemical Study. RSC Adv. 2016, 6, 65232–65242. [Google Scholar] [CrossRef]
  86. Cometto, C.; Yan, G.; Mariyappan, S.; Tarascon, J.-M. Means of Using Cyclic Voltammetry to Rapidly Design a Stable DMC-Based Electrolyte for Na-Ion Batteries. J. Electrochem. Soc. 2019, 166, A3723–A3730. [Google Scholar] [CrossRef]
  87. Zhang, Q.; Wang, Z.; Li, X.; Guo, H.; Peng, W.; Wang, J.; Yan, G. Comparative study of 1;3-propane sultone; prop-1-ene-1;3-sultone and ethylene sulfate as film-forming additives for sodium ion batteries. J. Power Sources 2022, 541, 231726. [Google Scholar] [CrossRef]
  88. Monti, D.; Ponrouch, A.; Palacín, M.R.; Johansson, P. Towards safer sodium-ion batteries via organic solvent/ionic liquid based hybrid electrolytes. J. Power Sources 2016, 324, 712–721. [Google Scholar] [CrossRef] [Green Version]
  89. Fukunaga, A.; Nohira, T.; Hagiwara, R.; Numata, K.; Itani, E.; Sakai, S.; Nitta, K. Performance validation of sodium-ion batteries using an ionic liquid electrolyte. J. Appl. Electrochem. 2016, 46, 487–496. [Google Scholar] [CrossRef] [Green Version]
  90. Benchakar, M.; Naéjus, R.; Damas, C.; Santos-Peña, J. Exploring the use of EMImFSI ionic liquid as additive or co-solvent for room temperature sodium ion battery electrolytes. Electrochim. Acta 2020, 330, 135193. [Google Scholar] [CrossRef]
  91. Lee, D.J.; Im, D.; Ryu, Y.-G.; Lee, S.; Yoon, J.; Lee, J.; Choi, W.; Jung, I.; Lee, S.; Doo, S.-G. Phosphorus derivatives as electrolyte additives for lithium-ion battery: The removal of O2 generated from lithium-rich layered oxide cathode. J. Power Sources 2013, 243, 831–835. [Google Scholar] [CrossRef]
  92. Yim, T.; Woo, S.-G.; Lim, S.H.; Cho, W.; Song, J.H.; Han, Y.-K.; Kim, Y.-J. 5V-Class High-Voltage Batteries with Over-Lithiated Oxide and a Multi-Functional Additive. J. Mater. Chem. A 2015, 3, 6157–6167. [Google Scholar] [CrossRef]
  93. Ledwoch, D.; Robinson, J.B.; Gastol, D.; Smith, K.; Shearing, P.; Brett, D.J.L.; Kendrick, E. Hard Carbon Composite Electrodes for Sodium-Ion Batteries with Nano-Zeolite and Carbon Black Additives. Batter. Supercaps 2021, 4, 163–172. [Google Scholar] [CrossRef]
  94. Zhang, X.; Liu, X.; Yang, C.; Li, N.; Ji, T.; Yan, K.; Zhu, B.; Yin, J.; Zhao, J.; Li, Y.A. V2O5-nanosheets-coated hard carbon fiber fabric as high-performance anode for sodium ion battery. Surf. Coat. Technol. 2019, 358, 661–666. [Google Scholar] [CrossRef]
  95. Lu, H.; Chen, X.; Jia, Y.; Chen, H.; Wang, Y.; Ai, X.; Yang, H.; Cao, Y. Engineering Al2O3 atomic layer deposition: Enhanced hard carbon-electrolyte interface towards practical sodium ion batteries. Nano Energy 2019, 64, 103903. [Google Scholar] [CrossRef]
  96. Li, Y.; Xu, S.; Wu, X.; Yu, J.; Wang, Y.; Hu, Y.-S.; Li, H.; Chen, L.; Huang, X. Amorphous monodispersed hard carbon micro-spherules derived from biomass as a high performance negative electrode material for sodium-ion batteries. J. Mater. Chem. A 2015, 3, 71–77. [Google Scholar] [CrossRef]
  97. Li, Q.; Zhu, Y.; Pinyi, Z.; Yuan, C.; Chen, M.; Wang, C. Commercial activated carbon as a novel precursor of the amorphous carbon for high performance sodium-ion batteries anode. Carbon 2018, 129, 85–94. [Google Scholar] [CrossRef]
  98. Ding, J.; Wang, H.; Li, Z.; Kohandehghan, A.; Cui, K.; Xu, Z.; Zahiri, B.; Tan, X.; Lotfabad, E.M.; Olsen, B.C.; et al. Carbon Nanosheet Frameworks Derived from Peat Moss as High Performance Sodium Ion Battery Anodes. ACS Nano 2013, 7, 11004–11015. [Google Scholar] [CrossRef]
  99. Bommier, C.; Luo, W.; Gao, W.-Y.; Greaney, A.; Ma, S.; Ji, X. Predicting capacity of hard carbon anodes in sodium-ion batteries using porosity measurements. Carbon 2014, 76, 165–174. [Google Scholar] [CrossRef]
  100. Xiao, L.; Lu, H.; Fang, Y.; Sushko, M.L.; Cao, Y.; Ai, X.; Yang, H.; Liu, J. Low-Defect and Low-Porosity Hard Carbon with High Coulombic Efficiency and High Capacity for Practical Sodium Ion Battery Anode. Adv. Energy Mater. 2018, 8, 1703238. [Google Scholar] [CrossRef]
  101. Wang, X.-K.; Shi, J.; Mi, L.-W.; Zhai, Y.-P.; Zhang, J.-Y.; Feng, X.-M.; Wu, Z.-J.; Chen, W.-H. Hierarchical porous hard carbon enables integral solid electrolyte interphase as robust anode for sodium-ion batteries. Rare Met. 2020, 39, 1053–1062. [Google Scholar] [CrossRef]
  102. Xie, H.; Wu, Z.; Wang, Z.; Qin, N.; Li, Y.; Cao, Y.; Lu, Z. Solid electrolyte interface stabilization via surface oxygen species functionalization in hard carbon for superior performance sodium-ion batteries. J. Mater. Chem. A 2020, 8, 3606–3612. [Google Scholar] [CrossRef]
  103. Evans, J.F.; Kuwana, T. Introduction of Functional Groups onto Carbon Electrodes via Treatment with Radio-Frequency Plasmas. Anal. Chem. 1979, 51, 358–365. [Google Scholar] [CrossRef]
  104. Besenhard, J.O.; Fritz, H.P. The Electrochemistry of Black Carbons. Angew. Chem. Int. Ed. Engl. 1983, 22, 950–975. [Google Scholar] [CrossRef]
  105. Lin, X.; Du, X.; Tsui, P.S.; Huang, J.-Q.; Tan, H.; Zhang, B. Exploring room- and low-temperature performance of hard carbon material in half and full Na-ion batteries. Electrochim. Acta 2019, 316, 60–68. [Google Scholar] [CrossRef]
  106. Rangom, Y.; Gaddam, R.R.; Duignan, T.T.; Zhao, X.S. Improvement of Hard Carbon Electrode Performance by Manipulating SEI Formation at High Charging Rates. ACS Appl. Mater. Interfaces 2019, 11, 34796–34804. [Google Scholar] [CrossRef]
  107. Liu, X.; Tan, Y.; Liu, T.; Wang, W.; Li, C.; Lu, J.; Sun, Y. A Simple Electrode-Level Chemical Presodiation Route by Solution Spraying to Improve the Energy Density of Sodium-Ion Batteries. Adv. Funct. Mater. 2019, 29, 1903795. [Google Scholar] [CrossRef]
  108. Thomas, P.; Billaud, D. Electrochemical insertion of sodium into hard carbons. Electrochim. Acta 2002, 47, 3303–3307. [Google Scholar] [CrossRef]
  109. Pan, Y.; Zhang, Y.; Parimalam, B.S.; Nguyen, C.C.; Wang, G.; Lucht, B.L. Investigation of the solid electrolyte interphase on hard carbon electrode for sodium ion batteries. J. Electroanalyt. Chem. 2017, 799, 181–186. [Google Scholar] [CrossRef]
  110. Thomas, P.; Ghanbaja, J.; Billaud, D. Electrochemical insertion of sodium in pitch-based carbon fibres in comparison with graphite in NaClO4–ethylene carbonate electrolyte. Electrochim. Acta 1999, 45, 423–430. [Google Scholar] [CrossRef]
  111. Schafzahl, L.; Ehmann, H.; Kriechbaum, M.; Sattelkow, J.; Ganner, T.; Plank, H.; Wilkening, M.; Freunberger, S.A. Long-Chain Li and Na Alkyl Carbonates as Solid Electrolyte Interphase Components: Structure; Ion Transport; and Mechanical Properties. Chem. Mater. 2018, 30, 3338–3345. [Google Scholar] [CrossRef]
  112. Bommier, C.; Leonard, D.; Jian, Z.; Stickle, W.F.; Greaney, P.A.; Ji, X. New Paradigms on the Nature of Solid Electrolyte Interphase Formation and Capacity Fading of Hard Carbon Anodes in Na-Ion Batteries. Adv. Mater. Interfaces 2016, 3, 1600449. [Google Scholar] [CrossRef]
  113. Skundin, A.M.; Kulova, T.L.; Yaroslavtsev, A.B. Sodium-ion Batteries (a Review). Russ. J. Electrochem. 2018, 54, 113–152. [Google Scholar] [CrossRef]
  114. Alcántara, R.; Lavela, P.; Ortiz, G.F.; Tirado, J.L. Carbon Microspheres Obtained from Resorcinol-Formaldehyde as High-Capacity Electrodes for Sodium-Ion Batteries. Electrochem. Solid-State Lett. 2005, 8, A222–A225. [Google Scholar] [CrossRef]
  115. Yan, X.; Liang, S.; Shi, H.; Hu, Y.; Liu, L.; Xu, Z. Nitrogen-enriched carbon nanofibers with tunable semi-ionic CAF bonds as a stable long cycle anode for sodium-ion batteries. J. Colloid Interface Sci. 2021, 583, 535–543. [Google Scholar] [CrossRef]
  116. Jin, J.; Shi, Z.; Wang, C. Electrochemical Performance of Electrospun carbon nanofibers as free-standing and binder-free anodes for Sodium-Ion and Lithium-Ion Batteries. Electrochim. Acta 2014, 141, 302–310. [Google Scholar] [CrossRef]
  117. Cao, Y.; Xiao, L.; Sushko, M.L.; Wang, W.; Schwenzer, B.; Xiao, J.; Nie, Z.; Saraf, L.V.; Yang, Z.; Liu, J. Sodium Ion Insertion in Hollow Carbon Nanowires for Battery Applications. Nano Lett. 2012, 12, 3783–3787. [Google Scholar] [CrossRef]
  118. Luo, W.; Schardt, J.; Bommier, C.; Wang, B.; Razink, J.; Simonsen, J.; Ji, X. Carbon nanofibers derived from cellulose nanofibers as a long-life anode material for rechargeable sodium-ion batteries. J. Mater. Chem. A 2013, 1, 10662–10666. [Google Scholar] [CrossRef]
  119. Cui, X.-Y.; Wang, Y.-J.; Wu, H.-D.; Lin, X.-D.; Tang, S.; Xu, P.; Liao, H.-G.; Zheng, M.-S.; Dong, Q.-F. A Carbon Foam with Sodiophilic Surface for Highly Reversible; Ultra-Long Cycle Sodium Metal Anode. Adv. Sci. 2021, 8, 2003178. [Google Scholar] [CrossRef]
  120. Wang, Z.; Yang, H.; Liu, Y.; Bai, Y.; Chen, G.; Li, Y.; Wang, X.; Xu, H.; Wu, C.; Lu, J. Analysis of the Stable Interphase Responsible for the Excellent Electrochemical Performance of Graphite Electrodes in Sodium-Ion Batteries. Small 2020, 16, 2003268. [Google Scholar] [CrossRef]
  121. Wang, C.; Huang, J.; Li, Q.; Cao, L.; Li, J.; Kajiyoshi, K. Catalyzing carbon surface by Ni to improve initial coulombic efficiency of sodium-ion batteries. J. Energy Storage 2020, 32, 101868. [Google Scholar] [CrossRef]
  122. Cabello, M.; Chyrka, T.; Klee, R.; Aragón, M.J.; Bai, X.; Lavela, P.; Vasylchenko, G.M.; Alcántara, R.; Tirado, J.L.; Ortiz, G.F. Treasure Na-ion anode from trash coke by adept electrolyte selection. J. Power Sources 2017, 347, 127–135. [Google Scholar] [CrossRef]
  123. Cao, B.; Liu, H.; Xu, B.; Lei, Y.; Chen, X.; Song, H. Mesoporous soft carbon as an anode material for sodium ion batteries with superior rate and cycling performance. J. Mater. Chem. A 2016, 4, 6472–6478. [Google Scholar] [CrossRef]
  124. Cuesta, N.; Cameán, I.; Arenillas, A.; García, A.B. Exploring the application of carbon xerogels as anodes for sodium-ion batteries. Microporous Mesoporous Mater. 2020, 308, 110542. [Google Scholar] [CrossRef]
  125. Jo, C.; Park, Y.; Jeong, J.; Lee, K.T.; Lee, J. Structural effect on electrochemical performance of ordered porous carbon electrodes for Na-ion batteries. ACS Appl. Mater. Interfaces 2015, 7, 11748–11754. [Google Scholar] [CrossRef]
  126. Mehmood, A.; Ali, G.; Koyutürk, B.; Pampel, J.; Chung, K.Y.; Fellinger, T.-P. Nanoporous nitrogen doped carbons with enhanced capacity for sodium ion battery anodes. Energy Storage Mater. 2020, 28, 101–111. [Google Scholar] [CrossRef]
  127. Kim, H.; Hong, J.; Yoon, G.; Kim, H.; Park, K.-Y.; Park, M.-S.; Yoon, W.-S.; Kang, K. Sodium intercalation chemistry in graphite. Energy Environ. Sci. 2015, 8, 2963–2969. [Google Scholar] [CrossRef]
  128. Jache, B.; Adelhelm, P. Use of Graphite as a Highly Reversible Electrode with Superior Cycle Life for Sodium-Ion Batteries by Making Use of Co-Intercalation Phenomena. Angew. Chem. 2014, 126, 10333–10337. [Google Scholar] [CrossRef]
  129. Seidl, L.; Bucher, N.; Chu, E.; Hartung, S.; Martens, S.; Schneider, O.; Stimming, U. Intercalation of solvated Na-ions into graphite. Energy Environ. Sci. 2017, 10, 1631–1642. [Google Scholar] [CrossRef] [Green Version]
  130. Kim, H.; Hong, J.; Park, Y.-U.; Kim, J.; Hwang, I.; Kang, K. Sodium Storage Behavior in Natural Graphite using Ether-based Electrolyte Systems. Adv. Funct. Mater. 2015, 25, 534–541. [Google Scholar] [CrossRef]
  131. Goktas, M.; Bolli, C.; Berg, E.J.; Novák, P.; Pollok, K.; Langenhorst, F.; Roeder, M.V.; Lenchuk, O.; Mollenhauer, D.; Adelhelm, P. Graphite as Cointercalation Electrode for Sodium-Ion Batteries: Electrode Dynamics and the Missing Solid Electrolyte Interphase (SEI). Adv. Energy Mater. 2018, 8, 1702724. [Google Scholar] [CrossRef]
  132. Lin, Z.; Tan, X.; Ge, L.; Lin, Y.; Yang, W.; Lin, J.; Fu, H.; Ying, S.; Huang, Z. Ultrathin 2D hexagon CoP/N-doped carbon nanosheets for robust sodium storage. J. Alloys Comp. 2022, 921, 166075. [Google Scholar] [CrossRef]
  133. Li, J.; Qi, H.; Wang, Q.; Xu, Z.; Liu, Y.; Li, Q.; Kong, X.; Huang, J. Constructing graphene-like nanosheets on porous carbon framework for promoted rate performance of Li-ion and Na-ion storage. Electrochim. Acta 2018, 271, 92–102. [Google Scholar] [CrossRef]
  134. Wang, C.; Huang, J.; Li, J.; Cao, L.; Chen, Q.; Qian, C.; Chen, S. Revealing the effect of nanopores in biomass-derived carbon on its sodium-ion storage behavior. ChemElectroChem 2020, 7, 201–211. [Google Scholar] [CrossRef]
  135. Wang, C.; Huang, J.; Qi, H.; Cao, L.; Xu, Z.; Cheng, Y.; Zhao, X.; Li, J. Controlling pseudographtic domain dimension of dandelion derived biomass carbon for excellent sodium-ion storage. J. Power Sources 2017, 358, 85–92. [Google Scholar] [CrossRef]
  136. Wang, C.; Huang, J.; Li, J.; Cao, L.; Xi, Q.; Chen, S. Exposing inner defects of porous carbon sheets to enhance rate performance of sodium-ion batteries. J. Electroanalyt. Chem. 2020, 860, 113924. [Google Scholar] [CrossRef]
  137. Lin, Q.; Zhang, J.; Kong, D.; Cao, T.; Zhang, S.-W.; Chen, X.; Tao, Y.; Lv, W.; Kang, F.; Yang, Q.-H. Deactivating Defects in Graphenes with Al2O3 Nanoclusters to Produce Long-Life and High-Rate Sodium-Ion Batteries. Adv. Energy Mater. 2019, 9, 1803078. [Google Scholar] [CrossRef] [Green Version]
  138. Kang, J.; Kim, D.-Y.; Chae, S.-A.; Saito, N.; Choid, S.-Y.; Kim, K.-H. Maximization of sodium storage capacity of pure carbon material used in sodium-ion batteries. J. Mater. Chem. A 2019, 7, 16149–16160. [Google Scholar] [CrossRef]
  139. Zhang, S.-W.; Lv, W.; Qiu, D.; Cao, T.; Zhang, J.; Lin, Q.; Chen, X.; He, Y.-B.; Kang, F.; Yang, Q.-H. An ion-conducting SnS–SnS2 hybrid coating for commercial activated carbons enabling their use as high performance anodes for sodium-ion batteries. J. Mater. Chem. A 2019, 7, 10761–10768. [Google Scholar] [CrossRef]
  140. Kim, Y.; Kim, D.S.; Um, J.H.; Yoon, J.; Kim, J.M.; Kim, H.; Won-Sub Yoon, W.-S. Revisiting Solid Electrolyte Interphase on the Carbonaceous Electrodes Using Soft X-ray Absorption Spectroscopy. ACS Appl. Mater. Interfaces 2018, 10, 29992–29999. [Google Scholar] [CrossRef]
  141. Maibach, J.; Jeschull, F.; Brandell, D.; Edström, K.; Valvo, M. Surface Layer Evolution on Graphite During Electrochemical Sodium-tetraglyme Co-intercalation. ACS Appl. Mater. Interfaces 2017, 9, 12373–12381. [Google Scholar] [CrossRef]
  142. Doeff, M.M.; Ma, Y.; Visco, S.J.; De Jonghe, L.C. Electrochemical Insertion of Sodium into Carbon. J. Electrochem. Soc. 1993, 140, L169–L170. [Google Scholar] [CrossRef]
  143. Zhang, J.; Wang, D.-W.; Lv, W.; Zhang, S.; Liang, Q.; Zheng, D.; Kang, F.; Yang, Q.-H. Achieving Superb Sodium Storage Performance on Carbon Anodes through Ether-derived Solid Electrolyte Interphase. Energy Environ. Sci. 2017, 10, 370–376. [Google Scholar] [CrossRef]
  144. Ye, L.; Liao, M.; Zhao, T.; Sun, H.; Zhao, Y.; Sun, X.; Wang, B.; Peng, H. A Sodiophilic Interphase-Mediated; Dendrite-Free Anode with Ultrahigh Specific Capacity for Sodium-Metal Batteries. Angew. Chem. 2019, 131, 17210–17216. [Google Scholar] [CrossRef]
  145. Luo, X.-F.; Helal, A.S.; Hsieh, C.-T.; Li, J.; Chang, J.-K. Three-dimensional carbon framework anode improves sodiation–desodiation properties in ionic liquid electrolyte. Nano Energy 2018, 49, 515–522. [Google Scholar] [CrossRef]
  146. Darwiche, A.; Bodenes, L.; Madec, L.; Monconduit, L.; Martinez, H. Impact of the salts and solvents on the SEI formation in Sb/Na batteries: An XPS analysis. Electrochim. Acta 2016, 207, 284–292. [Google Scholar] [CrossRef]
  147. Darwiche, A.; Marino, C.; Sougrati, M.T.; Fraisse, B.; Stievano, L.; Monconduit, L. Better Cycling Performances of Bulk Sb in Na-Ion Batteries Compared to Li-Ion Systems: An Unexpected Electrochemical Mechanism. J. Am. Chem. Soc. 2012, 134, 20805–20811. [Google Scholar] [CrossRef] [PubMed]
  148. He, M.; Kravchyk, K.; Walter, M.; Kovalenko, M.V. Monodisperse Antimony Nanocrystals for High-Rate Li-ion and Na-ion Battery Anodes: Nano versus Bulk. Nano Lett. 2014, 14, 1255–1262. [Google Scholar] [CrossRef]
  149. Baggetto, L.; Ganesh, P.; Sun, C.-N.; Meisner, R.A.; Zawodzinski, T.A.; Veith, G.M. Intrinsic thermodynamic and kinetic properties of Sb electrodes for Li-ion and Na-ion batteries: Experiment and theory. J. Mater. Chem. A 2013, 1, 7985–7994. [Google Scholar] [CrossRef]
  150. Zhou, X.; Dai, Z.; Bao, J.; Guo, Y. Wet milled synthesis of an Sb/MWCNT nanocomposite for improved sodium storage. J. Mater. Chem. A 2013, 1, 13727–13731. [Google Scholar] [CrossRef]
  151. Ji, L.; Gu, M.; Shao, Y.; Li, X.; Engelhard, M.H.; Arey, B.W.; Wang, W.; Nie, Z.; Xiao, J.; Wang, C.; et al. Controlling SEI Formation on SnSb-Porous Carbon Nanofibers for Improved Na Ion Storage. Adv. Mater. 2014, 26, 2901–2908. [Google Scholar] [CrossRef]
  152. Zheng, X.-M.; You, J.-H.; Fan, J.-J.; Tu, G.-P.; Rong, W.-Q.; Li, W.-J.; Wang, Y.-X.; Tao, S.; Zhang, P.-Y.; Zhang, S.-Y.; et al. Electrodeposited binder-free Sb/NiSb anode of sodium-ion batteries with excellent cycle stability and rate capability and new insights into its reaction mechanism by operando XRD analysis. Nano Energy 2020, 77, 105123. [Google Scholar] [CrossRef]
  153. Bian, X.; Dong, Y.; Zhao, D.; Ma, X.; Qiu, M.; Xu, J.; Jiao, L.; Cheng, F.; Zhang, N. Microsized Antimony as a Stable Anode in Fluoroethylene Carbonate Containing Electrolytes for Rechargeable Lithium-/Sodium-Ion Batteries. ACS Appl. Mater. Interfaces 2020, 12, 3554–3562. [Google Scholar] [CrossRef] [PubMed]
  154. Williamson, G.A.; Hu, V.W.; Yoo, T.B.; Affandy, M.; Opie, C.; Paradis, E.K.; Holmberg, V.C. Temperature-Dependent Electrochemical Characteristics of Antimony Nanocrystal Alloying Electrodes for Na-Ion Batteries. ACS Appl. Energy Mater. 2019, 2, 6741–6750. [Google Scholar] [CrossRef]
  155. Song, J.; Xiao, D.; Jia, H.; Zhu, G.; Engelhard, M.; Xiao, B.; Feng, S.; Li, D.; Reed, D.; Sprenkle, V.L.; et al. A comparative study of pomegranate Sb@C yolk–shell microspheres as Li and Na-ion battery anodes. Nanoscale 2019, 11, 348–355. [Google Scholar] [CrossRef] [PubMed]
  156. Li, P.; Guo, X.; Wang, S.; Zang, R.; Li, X.; Man, Z.; Li, P.; Liu, S.; Wu, Y.; Wang, G. Two-dimensional Sb@TiO2−x nanoplates as a high-performance anode material for sodium-ion batteries. J. Mater. Chem. A 2019, 7, 2553–2559. [Google Scholar] [CrossRef]
  157. Bodenes, L.; Darwiche, A.; Monconduit, L.; Martinez, H. The Solid Electrolyte Interphase a key parameter of the high performance of Sb in sodium-ion batteries: Comparative X-ray Photoelectron Spectroscopy study of Sb/Na-ion and Sb/Li-ion batteries. J. Power Sources 2015, 273, 14–24. [Google Scholar] [CrossRef]
  158. Qian, J.; Chen, Y.; Wu, L.; Cao, Y.; Ai, X.; Yang, H. High capacity Na-storage and superior cyclability of nanocomposite Sb/C anode for Na-ion batteries. Chem. Commun. 2012, 48, 7070–7072. [Google Scholar] [CrossRef] [PubMed]
  159. Wang, M.; Yang, Z.; Wang, J.; Li, W.; Gu, L.; Yu, Y. Sb Nanoparticles Encapsulated in a Reticular Amorphous Carbon Network for Enhanced Sodium Storage. Small 2015, 11, 5381–5387. [Google Scholar] [CrossRef]
  160. Ko, Y.N.; Kang, Y.C. Electrochemical properties of ultrafine Sb nanocrystals embedded in carbon microspheres for use as Na-ion battery anode materials. Chem. Commun. 2014, 50, 12322–12324. [Google Scholar] [CrossRef]
  161. Zhu, Y.; Han, X.; Xu, Y.; Liu, Y.; Zheng, S.; Xu, K.; Hu, L.; Wang, C. Electrospun Sb/C Fibers for a Stable and Fast Sodium-Ion Battery Anode. ACS Nano 2013, 7, 6378–6386. [Google Scholar] [CrossRef]
  162. Huang, J.; Guo, X.; Du, X.; Lin, X.; Huang, J.-Q.; Tan, H.; Zhu, Y.; Zhang, B. Nanostructures of solid electrolyte interphases and their consequences for microsized Sn anodes in sodium ion batteries. Energy Environ. Sci. 2019, 12, 1550–1557. [Google Scholar] [CrossRef]
  163. Song, M.; Wang, C.; Du, D.; Li, F.; Chen, J. A high-energy-density sodium-ion full battery based on tin anode. Sci. China Chem. 2019, 62, 616–621; [Google Scholar] [CrossRef]
  164. Prosini, P.P.; Carewska, M.; Cento, C.; Tarquini, G.; Maroni, F.; Birrozzi, A.; Nobili, F. Tin-Decorated Reduced Graphene Oxide and NaLi0.2Ni0.25Mn0.75Oδ as Electrode Materials for Sodium-Ion Batteries. Materials 2019, 12, 1074. [Google Scholar] [CrossRef] [Green Version]
  165. Zhang, B.; Rousse, G.; Foix, D.; Dugas, R.; Corte, D.A.D.; Tarascon, J.-M. Microsized Sn as Advanced Anodes in Glyme-Based Electrolyte for Na-Ion Batteries. Adv. Mater. 2016, 28, 9824–9830. [Google Scholar] [CrossRef] [Green Version]
  166. Baggetto, L.; Ganesh, P.; Meisner, R.P.; Unocic, R.R.; Jumas, J.-C.; Bridges, C.A.; Veith, G.M. Characterization of sodium ion electrochemical reaction with tin anodes: Experiment and theory. J. Power Sources 2013, 234, 48–59. [Google Scholar] [CrossRef]
  167. Komaba, S.; Matsuura, Y.; Ishikawa, T.; Yabuuchi, N.; Murata, W.; Kuze, S. Redox reaction of Sn-polyacrylate electrodes in aprotic Na cell. Electrochem. Commun. 2012, 21, 65–68. [Google Scholar] [CrossRef]
  168. Wang, Z.; Dong, K.; Wang, D.; Luo, S.; Liu, Y.; Wang, Q.; Zhang, Y.; Hao, A.; Shid, C.; Zhao, N. A nanosized SnSb alloy confined in N-doped 3D porous carbon coupled with ether-based electrolytes toward high-performance potassium-ion batteries. J. Mater. Chem. A 2019, 7, 14309–14318. [Google Scholar] [CrossRef]
  169. Yuan, H.; Ma, F.; Wei, X.; Lan, J.-L.; Liu, Y.; Yu, Y.; Yang, X.; Park, H.S. Ionic-Conducting and Robust Multilayered Solid Electrolyte Interphases for Greatly Improved Rate and Cycling Capabilities of Sodium Ion Full Cells. Adv. Energy Mater. 2020, 10, 2001418. [Google Scholar] [CrossRef]
  170. Yang, H.; Xu, R.; Yao, Y.; Ye, S.; Zhou, X.; Yu, Y. Multicore–Shell Bi@N-doped Carbon Nanospheres for High Power Density and Long Cycle Life Sodium- and Potassium-Ion Anodes. Adv. Funct. Mater. 2019, 29, 1809195. [Google Scholar] [CrossRef]
  171. Long, H.; Yin, X.; Wang, X.; Zhao, Y.; Yan, L. Bismuth nanorods confined in hollow carbon structures for high performance sodium- and potassium-ion batteries. J. Energy Chem. 2022, 67, 787–796. [Google Scholar] [CrossRef]
  172. Xiong, P.; Bai, P.; Li, A.; Li, B.; Cheng, M.; Chen, Y.; Huang, S.; Jiang, Q.; Bu, X.-H.; Xu, Y. Bismuth Nanoparticle@Carbon Composite Anodes for Ultralong Cycle Life and High-Rate Sodium-Ion Batteries. Adv. Mater. 2019, 31, 1904771. [Google Scholar] [CrossRef] [PubMed]
  173. Cheng, X.; Li, D.; Wu, Y.; Xu, R.; Yu, Y. Bismuth Nanospheres Embedding in Three-Dimensional (3D) Porous Graphene Frameworks as Highly Performance Anodes for Sodium and Potassium-Ion Batteries. J. Mater. Chem. A 2019, 7, 4913–4921. [Google Scholar] [CrossRef]
  174. Xue, P.; Wang, N.; Fang, Z.; Lu, Z.; Xu, X.; Wang, L.; Du, Y.; Ren, X.; Bai, Z.; Dou, S.; et al. Rayleigh-Instability-Induced Bismuth Nanorod@Nitrogen-Doped Carbon Nanotubes as A Long Cycling and High Rate Anode for Sodium-Ion Batteries. Nano Lett. 2019, 19, 1998–2004. [Google Scholar] [CrossRef] [PubMed]
  175. Qiu, J.; Li, S.; Su, X.; Wang, Y.; Xua, L.; Yuan, S.; Li, H.; Zhang, S. Bismuth nano-spheres encapsulated in porous carbon network for robust and fast sodium storage. Chem. Eng. J. 2017, 320, 300–307. [Google Scholar] [CrossRef]
  176. Weadock, N.; Varongchayakul, N.; Wan, J.; Lee, S.; Seog, J.; Hun, L. Determination of mechanical properties of the SEI in sodium ion batteries via colloidal probe microscopy. Nano Energy 2013, 2, 713–719. [Google Scholar] [CrossRef]
  177. Li, X.; Steirer, K.X.; Krishna, L.; Xiao, C.; Fink, K.; Santhanagopalan, S. Electrochemical Properties and Challenges of Type II Silicon Clathrate Anode in Sodium Ion Batteries. J. Electrochem. Soc. 2019, 166, A3051–A3058. [Google Scholar] [CrossRef]
  178. Xu, Y.; Swaans, E.; Basak, S.; Zandbergen, H.W.; Borsa, D.M.; Mulder, F.M. Reversible Na-Ion Uptake in Si Nanoparticles. Adv. Energy Mater. 2016, 6, 1501436. [Google Scholar] [CrossRef]
  179. Gavrilin, I.M.; Smolyaninov, V.A.; Dronov, A.A.; Gavrilov, S.A.; Trifonov, A.Y.; Kulova, T.L.; Kuz’mina, A.A.; Skundin, A.M. Electrochemical insertion of sodium into nanostructured materials based on germanium. Mendeleev Commun. 2018, 28, 659–660. [Google Scholar] [CrossRef]
  180. Lebedev, E.A.; Gavrilin, I.M.; Kudryashova, Y.O.; Martynova, I.K.; Volkov, R.L.; Kulova, T.L.; Skundin, A.M.; Borgardt, N.I.; Gavrilov, S.A. Effect of Vinylene Carbonate Electrolyte Additive on the Process of Insertion/Extraction of Na into Ge Microrods Formed by Electrodeposition. Batteries 2022, 8, 109. [Google Scholar] [CrossRef]
  181. Oh, S.-M.; Myung, S.-T.; Yoon, C.S.; Lu, J.; Hassoun, J.; Scrosati, B.; Amine, K.; Sun, Y.-K. Advanced Na[Ni0.25Fe0.5Mn0.25]O2/C–Fe3O4 Sodium-Ion Batteries Using EMS Electrolyte for Energy Storage. Nano Lett. 2014, 14, 1620–1626. [Google Scholar] [CrossRef]
  182. Ahuja, V.; Vengarathody, R.; Singh, S.; Senguttuvan, P. Realization of high cycle life bismuth oxychloride Na-ion anode in glyme-based electrolyte. J. Power Sources 2022, 529, 231227. [Google Scholar] [CrossRef]
  183. Zhang, Y.; Lu, S.; Wang, M.-Q.; Niu, Y.; Liu, S.; Li, Y.; Wu, X.; Bao, S.-J.; Xu, M. Bismuth oxychloride ultrathin nanoplates as an anode material for sodium-ion batteries. Mater. Lett. 2016, 178, 44–47. [Google Scholar] [CrossRef]
  184. Bai, Z.; Lv, X.; Liu, D.-H.; Dai, D.; Gu, J.; Yang, L.; Chen, Z. Two-Dimensional NiO@C-N Nanosheets Composite as a Superior Low-Temperature Anode Material for Advanced Lithium-/Sodium-Ion Batteries. ChemElectroChem 2020, 7, 3616–3622. [Google Scholar] [CrossRef]
  185. Cha, C.; Mohajernia, S.; Nguyen, N.T.; Mazare, A.; Denisov, N.; Hwang, I.; Schmuki, P. Li+ Pre-Insertion Leads to Formation of Solid Electrolyte Interface on TiO2 Nanotubes That Enables High-Performance Anodes for Sodium Ion Batteries. Adv. Energy Mater. 2020, 10, 1903448. [Google Scholar] [CrossRef] [Green Version]
  186. Kong, H.; Wu, Y.; Hong, W.; Yan, C.; Zhao, Y.; Chen, G. Structure-designed synthesis of Cu-doped Co3O4@N-doped carbon with interior void space for optimizing alkali-ion storage. Energy Storage Mater. 2020, 24, 610–617. [Google Scholar] [CrossRef]
  187. Rubio, S.; Maça, R.R.; Aragón, M.J.; Cabello, M.; Castillo-Rodríguez, M.; Lavela, P.; Tirado, J.L.; Etacheri, V.; Ortiz, G.F. Superior electrochemical performance of TiO2 sodium-ion battery anodes in diglyme-based electrolyte solution. J. Power Sources 2019, 432, 82–91. [Google Scholar] [CrossRef]
  188. Wang, B.; Zhang, X.; Liu, X.; Wang, G.; Wang, H.; Bai, J. Rational design of Fe3O4@C yolk-shell nanorods constituting a stable anode for high-performance Li/Na-ion batteries. J. Colloid Interface Sci. 2018, 528, 225–236. [Google Scholar] [CrossRef]
  189. Philippe, B.; Valvo, M.; Lindgren, F.; Rensmo, H.; Edström, K. Investigation of the Electrode/Electrolyte Interface of Fe2O3 Composite Electrodes: Li vs Na Batteries. Chem. Mater. 2014, 26, 5028–5041. [Google Scholar] [CrossRef]
  190. Komaba, S.; Mikumo, T.; Yabuuchi, N.; Ogata, A.; Yoshida, H.; Yamada, Y. Electrochemical Insertion of Li and Na Ions into Nanocrystalline Fe3O4 and α-Fe2O3 for Rechargeable Batteries. J. Electrochem. Soc. 2010, 157, A60–A65. [Google Scholar] [CrossRef]
  191. Valvo, M.; Lindgren, F.; Lafont, U.; Björefors, F.; Edström, K. Towards more sustainable negative electrodes in Na-ion batteries via nanostructured iron oxide. J. Power Sources 2014, 245, 967–978. [Google Scholar] [CrossRef]
  192. Kim, K.-T.; Ali, G.; Chung, K.Y.; Yoon, C.S.; Yashiro, H.; Sun, Y.-K.; Lu, J.; Amine, K.; Myung, S.-T. Anatase Titania Nanorods as an Intercalation Anode Material for Rechargeable Sodium Batteries. Nano Lett. 2014, 14, 416–422. [Google Scholar] [CrossRef] [PubMed]
  193. Zhou, M.; Xu, Y.; Wang, C.; Li, Q.; Xiang, J.; Liang, L.; Wu, M.; Zhao, H.; Lei, Y. Amorphous TiO2 Inverse Opal Anode for High-rate Sodium Ion Batteries. Nano Energy 2017, 31, 514–524. [Google Scholar] [CrossRef]
  194. Bresser, D.; Oschmann, B.; Tahir, M.N.; Mueller, F.; Lieberwirth, I.; Tremel, W.; Zentel, R.; Passerini, S. Carbon-Coated Anatase TiO2 Nanotubes for Li and Na-Ion Anodes. J. Electrochem. Soc. 2015, 162, A3013–A3020. [Google Scholar] [CrossRef]
  195. Oh, S.-M.; Hwang, J.-Y.; Yoon, C.S.; Lu, J.; Amine, K.; Belharouak, I.; Sun, Y.-K. High Electrochemical Performances of Microsphere C-TiO2 Anode for Sodium-Ion Battery. ACS Appl. Mater. Interfaces 2014, 6, 11295–11301. [Google Scholar] [CrossRef] [PubMed]
  196. Xu, Y.; Lotfabad, E.M.; Wang, H.; Farbod, B.; Xu, Z.; Kohandehghan, A.; Mitlin, D. Nanocrystalline anatase TiO2: A new anode material for rechargeable sodium ion batteries. Chem. Commun. 2013, 49, 8973–8975. [Google Scholar] [CrossRef] [PubMed]
  197. Xiong, H.; Slater, M.D.; Balasubramanian, M.; Johnson, C.S.; Rajh, T. Amorphous TiO2 Nanotube Anode for Rechargeable Sodium Ion Batteries. J. Phys. Chem. Lett. 2011, 2, 2560–2565. [Google Scholar] [CrossRef]
  198. Bi, Z.; Paranthaman, M.P.; Menchhofer, P.A.; Dehoff, R.R.; Bridges, C.A.; Chi, M.; Guo, B.; Sun, X.-G.; Dai, S. Self-organized amorphous TiO2 nanotube arrays on porous Ti foam for rechargeable lithium and sodium ion batteries. J. Power Sources 2013, 222, 461–466. [Google Scholar] [CrossRef]
  199. Wu, L.; Bresser, D.; Buchholz, D.; Passerini, S. Nanocrystalline TiO2(B) as Anode Material for Sodium-Ion Batteries. J. Electrochem. Soc. 2015, 162, A3052–A3058. [Google Scholar] [CrossRef] [Green Version]
  200. Wu, L.; Buchholz, D.; Bresser, D.; Chagas, L.G.; Passerini, S. Anatase TiO2 nanoparticles for high power sodium-ion anodes. J. Power Sources 2014, 251, 379–385. [Google Scholar] [CrossRef]
  201. Yan, Z.; Liu, L.; Tan, J.; Zhou, Q.; Huang, Z.; Xia, D.; Shu, H.; Yang, X.; Wang, X. One-pot synthesis of bicrystalline titanium dioxide spheres with a core–shell structure as anode materials for lithium and sodium ion batteries. J. Power Sources 2014, 269, 37–45. [Google Scholar] [CrossRef]
  202. Hwang, J.-Y.; Myung, S.-T.; Lee, J.-H.; Abouimrane, A.; Belharouak, I.; Sun, Y.-K. Ultrafast sodium storage in anataseTiO2 nanoparticles embedded on carbon nanotubes. Nano Energy 2015, 16, 218–226. [Google Scholar] [CrossRef]
  203. Pérez-Flores, J.C.; Baehtz, C.; Kuhn, A.; García-Alvarado, F. Hollandite-type TiO2: A new negative electrode material for sodium-ion batteries. J. Mater. Chem. A 2014, 2, 1825–1833. [Google Scholar] [CrossRef]
  204. Prutsch, D.; Wilkening, M.; Hanzu, I. Long Cycle-Life Na-ion Anodes Based on Amorphous Titania Nanotubes—Interfaces and Diffusion. ACS Appl. Mater. Interfaces 2015, 7, 25757–25769. [Google Scholar] [CrossRef]
  205. Usui, H.; Yoshioka, S.; Wasada, K.; Shimizu, M.; Sakaguchi, H. Nb-Doped Rutile TiO2: A Potential Anode Material for Na-Ion Battery. ACS Appl. Mater. Interfaces 2015, 7, 6567–6573. [Google Scholar] [CrossRef] [PubMed]
  206. Lee, J.; Chen, Y.-M.; Zhu, Y.; Vogt, B.D. Fabrication of porous carbon/TiO2 composites through polymerization induced phase separation and use as an anode for Na-ion batteries. ACS Appl. Mater. Interfaces 2014, 6, 21011–21018. [Google Scholar] [CrossRef]
  207. Wen, J.-W.; Zhang, D.-W.; Zang, Y.; Sun, X.; Cheng, B.; Ding, C.-X.; Yu, Y.; Chen, C.-H. Li and Na storage behavior of bowl-like hollow Co3O4 microspheres as an anode material for lithium-ion and sodium-ion batteries. Electrochim. Acta 2014, 132, 193–199. [Google Scholar] [CrossRef]
  208. Jian, Z.; Liu, P.; Li, F.; Chen, M.; Zhou, H. Monodispersed hierarchical Co3O4 spheres intertwined with carbon nanotubes for use as anode materials in sodium-ion batteries. J. Mater. Chem. A 2014, 2, 13805–13809. [Google Scholar] [CrossRef]
  209. Klavetter, K.C.; Garcia, S.; Dahal, N.; Snider, J.L.; de Souza, J.P.; Cell, T.H.; Cassara, M.A.; Heller, A.; Humphrey, S.M.; Mullins, C.B. Li- and Na-reduction products of meso-Co3O4 form high-rate; stably cycling battery anode materials. J. Mater. Chem. A 2014, 2, 14209–14221. [Google Scholar] [CrossRef]
  210. Rahman, M.M.; Glushenkov, A.M.; Ramireddy, T.; Chen, Y. Electrochemical investigation of sodium reactivity with nanostructured Co3O4 for sodium-ion batteries. Chem. Commun. 2014, 50, 5057–5060. [Google Scholar] [CrossRef] [Green Version]
  211. Rahman, M.M.; Sultana, I.; Chen, Z.; Srikanth, M.; Li, L.H.; Dai, X.J.; Chen, Y. Ex-situ electrochemical sodiation/desodiation observation of Co3O4 anchored carbon nanotubes: A high performance sodium-ion battery anode produced by a pulsed plasma in liquid. Nanoscale 2015, 7, 13088–13095. [Google Scholar] [CrossRef]
  212. Liu, Y.; Cheng, Z.; Sun, H.; Arandiyan, H.; Li, J.; Ahmad, M. Mesoporous Co3O4 sheets/3D graphene networks nanohybrids for high-performance sodium-ion battery anode. J. Power Sources 2015, 273, 878–884. [Google Scholar] [CrossRef]
  213. Sun, W.; Rui, X.; Zhu, J.; Yu, L.; Zhang, Y.; Xu, Z.; Madhavi, S.; Yan, Q. Ultrathin nickel oxide nanosheets for enhanced sodium and lithium storage. J. Power Sources 2015, 274, 755–761. [Google Scholar] [CrossRef]
  214. Sun, Q.; Ren, Q.-Q.; Li, H.; Fu, Z.-W. High capacity Sb2O4 thin film electrodes for rechargeable sodium battery. Electrochem. Commun. 2011, 13, 1462–1464. [Google Scholar] [CrossRef]
  215. Zhou, X.; Liu, X.; Xu, Y.; Liu, Y.; Dai, Z.; Bao, J. An SbOx/Reduced Graphene Oxide Composite as a High-Rate Anode Material for Sodium-Ion Batteries. J. Phys. Chem. C 2014, 118, 23527–23534. [Google Scholar] [CrossRef]
  216. Hu, M.; Jiang, Y.; Sun, W.; Wang, H.; Jin, C.; Yan, M. Reversible Conversion-Alloying of Sb2O3 as a High-Capacity; High-Rate; and Durable Anode for Sodium Ion Batteries. ACS Appl. Mater. Interfaces 2014, 6, 19449–19455. [Google Scholar] [CrossRef] [PubMed]
  217. Li, N.; Liao, S.; Sun, Y.; Song, H.W.; Wang, C.X. Uniformly dispersed self-assembled growth of Sb2O3/Sb@graphene nanocomposites on a 3D carbon sheet network for high Na-storage capacity and excellent stability. J. Mater. Chem. A 2015, 3, 5820–5828. [Google Scholar] [CrossRef]
  218. Kim, H.; Lim, E.; Jo, C.; Yoon, G.; Hwang, J.; Jeong, S.; Lee, J.; Kang, K. Ordered-mesoporous Nb2O5/carbon composite as a sodium insertion material. Nano Energy 2015, 16, 62–70. [Google Scholar] [CrossRef]
  219. Zhang, Y.; Xie, J.; Zhang, S.; Zhu, P.; Cao, G.; Zhao, X. Ultrafine tin oxide on reduced graphene oxide as high-performance anode for sodium-ion batteries. Electrochim. Acta 2015, 151, 8–15. [Google Scholar] [CrossRef]
  220. Wang, Y.; Su, D.; Wang, C.; Wang, G. SnO2@MWCNT nanocomposite as a high capacity anode material for sodium-ion batteries. Electrochem. Commun 2013, 29, 8–11. [Google Scholar] [CrossRef]
  221. Su, D.; Ahn, H.-Y.; Wang, G. SnO2@graphene nanocomposites as anode materials for Na-ion batteries with superior electrochemical performance. Chem. Commun. 2013, 49, 3131–3133. [Google Scholar] [CrossRef]
  222. Shimizu, M.; Usui, H.; Sakaguchi, H. Electrochemical Na-insertion/extraction properties of SnO thick-film electrodes prepared by gas-deposition. J. Power Sources 2014, 248, 378–382. [Google Scholar] [CrossRef] [Green Version]
  223. Su, D.; Xie, X.; Wang, G. Hierarchical Mesoporous SnO Microspheres as High Capacity Anode Materials for Sodium-Ion Batteries. Chem. Eur. J. 2014, 20, 3192–3197. [Google Scholar] [CrossRef] [PubMed]
  224. Wang, Y.-X.; Lim, Y.-G.; Park, M.-S.; Chou, S.-L.; Kim, J.H.; Liu, H.-K.; Dou, S.-X.; Kim, Y.-J. Ultrafine SnO2 nanoparticle loading onto reduced graphene oxide as anodes for sodium-ion batteries with superior rate and cycling performances. J. Mater. Chem. A 2014, 2, 529–534. [Google Scholar] [CrossRef]
  225. Lu, Y.C.; Ma, C.; Alvarado, J.; Kidera, T.; Dimov, N.; Meng, Y.S.; Okada, S. Electrochemical properties of tin oxide anodes for sodium-ion batteries. J. Power Sources 2015, 284, 287–295. [Google Scholar] [CrossRef]
  226. Górka, J.; Baggetto, L.; Keum, J.K.; Mahurin, S.M.; Mayes, R.T.; Dai, S.; Veith, G.M. The electrochemical reactions of SnO2 with Li and Na: A study using thin films and mesoporous carbons. J. Power Sources 2015, 284, 1–9. [Google Scholar] [CrossRef] [Green Version]
  227. Liu, Y.; Fang, X.; Ge, M.; Rong, J.; Shen, C.; Zhang, A.; Enaya, H.A.; Zhou, C. SnO2 coated carbon cloth with surface modification as Na-ion battery anode. Nano Energy 2015, 16, 399–407. [Google Scholar] [CrossRef]
  228. Cheng, Y.; Huang, J.; Li, J.; Xu, Z.; Cao, L.; Ouyang, H.; Yan, J.; Qi, H. SnO2/super P nanocomposites as anode materials for Na-ion batteries with enhanced electrochemical performance. J. Alloys Compd. 2016, 658, 234–240. [Google Scholar] [CrossRef]
  229. Yuan, H.; Ma, F.; Wei, X.; Jia, S.; Kang, P.; Yu, Y.; Yang, X.; Lan, J.-L. Engineering fluoride-rich interphase and inhibiting grain coarsening for highly reversible bismuth sulfide anode for sodium storage. Mater. Today Energy 2022, 28, 101084. [Google Scholar] [CrossRef]
  230. Chen, L.; Song, K.; Shi, J.; Zhang, J.; Mi, L.; Chen, W.; Liu, C.; Shen, C. PAANa-induced ductile SEI of bare micro-sized FeS enables high sodium-ion storage performance. Sci. China Mater. 2021, 64, 105–114. [Google Scholar] [CrossRef]
  231. Zhang, C.; Wang, F.; Han, F.; Wu, H.; Zhang, F.; Zhang, G.; Liu, J. Improved Electrochemical Performance of Sodium/Potassium-Ion Batteries in Ether-Based Electrolyte: Cases Study of MoS2@C and Fe7S8@C Anodes. Adv. Mater. Interfaces 2020, 7, 2000486. [Google Scholar] [CrossRef]
  232. Wang, T.; Yang, K.; Shi, J.; Zhou, S.; Mi, L.; Li, H.; Chen, W. Simple synthesis of sandwich-like SnSe2/rGO as high initial coulombic efficiency and high stability anode for sodium-ion batteries. J. Energy Chem. 2020, 46, 71–77. [Google Scholar] [CrossRef]
  233. Ihsan-Ul-Haq, M.; Huang, H.; Wu, J.; Cui, J.; Yao, S.; Chong, W.G.; Huang, B.; Kim, J.-K. Thin solid electrolyte interface on chemically bonded Sb2Te3/CNT composite anodes for high performance sodium ion full cells. Nano Energy 2020, 71, 104613. [Google Scholar] [CrossRef]
  234. Xu, S.; Gao, X.; Hua, Y.; Neville, A.; Wang, Y.; Zhang, K. Rapid deposition of WS2 platelet thin films as additive-free anode for sodium ion batteries with superior volumetric capacity. Energy Storage Mater. 2020, 26, 534–542. [Google Scholar] [CrossRef]
  235. Yang, K.; Zhang, X.; Song, K.; Zhang, J.; Liu, C.; Mi, L.; Wang, Y.; Chen, W. Se–C bond and reversible SEI in facile synthesized SnSe2⸦3D carbon induced stable anode for sodium-ion batteries. Electrochim. Acta 2020, 337, 135783. [Google Scholar] [CrossRef]
  236. Chang, X.; Ma, Y.; Yang, M.; Xing, T.; Tang, L.; Chen, T.; Guo, Q.; Zhu, X.; Liu, J.; Xia, H. In-situ solid-state growth of N; S codoped carbon nanotubes encapsulating metal sulfides for high-efficient-stable sodium ion storage. Energy Storage Mater. 2019, 23, 358–366. [Google Scholar] [CrossRef]
  237. Liu, Q.; Gao, J.; Cao, C.; Yin, G.; Jiang, Z.; Ge, M.; Xiao, X.; Lee, W.-K.; Wang, J. Insights into enhanced sodium ion storage mechanism in Fe3S4: The coupling of surface chemistry; microstructural regulation and 3D electronic transport. Nano Energy 2019, 62, 384–392. [Google Scholar] [CrossRef]
  238. Li, X.; Zhang, W.; Feng, Y.; Li, W.; Peng, P.; Yao, J.; Li, M.; Jiang, C. Ultrafine CoSe nano-crystallites confined in leaf-like N-doped carbon for long-cyclic and fast sodium ion storage. Electrochim. Acta 2019, 294, 173–182. [Google Scholar] [CrossRef]
  239. Dong, C.; Wu, L.; He, Y.; Zhou, Y.; Sun, X.; Du, W.; Sun, X.; Xu, L.; Jiang, F. Willow-Leaf-Like ZnSe@N-Doped Carbon Nanoarchitecture as a Stable and High-Performance Anode Material for Sodium-Ion and Potassium-Ion Batteries. Small 2020, 16, 2004580. [Google Scholar] [CrossRef]
  240. Shi, Z.-T.; Kang, W.; Xu, J.; Sun, L.-L.; Wu, C.; Wang, L.; Yu, Y.-Q.; Yu, D.Y.W.; Zhang, W.; Lee, C.-S. In Situ Carbon-Doped Mo(Se0.85S0.15)2 Hierarchical Nanotubes as Stable Anodes for High-Performance Sodium-Ion Batteries. Small 2015, 11, 5667–5674. [Google Scholar] [CrossRef]
  241. Ahmed, B.; Anjum, D.H.; Hedhili, M.N.; Alshareef, H.N. Mechanistic Insight into the Stability of HfO2–Coated MoS2 Nanosheet Anodes for Sodium Ion Batteries. Small 2015, 11, 4341–4350. [Google Scholar] [CrossRef]
  242. Xiong, X.; Wang, G.; Lin, Y.; Wang, Y.; Ou, X.; Zheng, F.; Yang, C.; Wang, J.-H.; Liu, M. Enhancing Sodium Ion Battery Performance by Strongly Binding Nanostructured Sb2S3 on Sulfur-Doped Graphene Sheets. ACS Nano 2016, 10, 10953–10959. [Google Scholar] [CrossRef] [PubMed]
  243. Zhou, T.; Pang, W.K.; Zhang, C.; Yang, J.; Chen, Z.; Liu, H.K.; Guo, Z. Enhanced Sodium-Ion Battery Performance by Structural Phase Transition from Two-Dimensional Hexagonal-SnS2 to Orthorhombic-SnS. ACS Nano 2014, 8, 8323–8333. [Google Scholar] [CrossRef] [PubMed]
  244. Jang, J.Y.; Lee, Y.; Kim, Y.; Lee, J.; Lee, S.-M.; Lee, K.T.; Choi, N.-S. Interfacial architectures based on a binary additive combination for high-performance Sn4P3 anodes in sodium-ion batteries. J. Mater. Chem. A 2015, 3, 8332–8338. [Google Scholar] [CrossRef]
  245. Li, W.; Chou, S.-L.; Wang, J.-Z.; Kim, J.H.; Liu, H.-K.; Dou, S.-X. Sn4+xP3@Amorphous Sn-P Composites as Anodes for Sodium-Ion Batteries with Low Cost; High Capacity; Long Life; and Superior Rate Capability. Adv. Mater. 2014, 26, 4037–4042. [Google Scholar] [CrossRef]
  246. Nazarian-Samani, M.; Nazarian-Samani, M.; Haghighat-Shishavan, S.; Kim, K.-B. Predelithiation-driven ultrastable Na-ion battery performance using Si; P-rich ternary M-Si-P anodes. Energy Storage Mater. 2022, 49, 421–432. [Google Scholar] [CrossRef]
  247. Li, M.; Feng, N.; Liu, M.; Cong, Z.; Sun, J.; Du, C.; Liu, Q.; Pu, X.; Hua, W. Hierarchically porous carbon/red phosphorus composite for high-capacity sodium-ion battery anode. Sci. Bull. 2018, 63, 982–989. [Google Scholar] [CrossRef] [Green Version]
  248. Kaushik, S.; Matsumoto, K.; Orikasa, Y.; Katayama, M.; Inada, Y.; Sato, Y.; Gotoh, K.; Ando, H.; Hagiwara, R. Vanadium diphosphide as a negative electrode material for sodium secondary batteries. J. Power Sources 2021, 483, 229182. [Google Scholar] [CrossRef]
  249. Zhang, J.; Zhang, K.; Yang, J.; Lau, V.W.; Lee, G.; Park, M.; Kang, Y.-M. Engineering Solid Electrolyte Interphase on Red Phosphorus for Long-Term and High-Capacity Sodium Storage. Chem. Mater. 2020, 32, 448–458. [Google Scholar] [CrossRef]
  250. Liu, W.; Yuan, X.; Yu, X. A core–shell structure of polydopamine-coated phosphorus–carbon nanotube composite for high-performance sodium-ion batteries. Nanoscale 2018, 10, 16675–16682. [Google Scholar] [CrossRef]
  251. Liu, S.; Feng, J.; Bian, X.; Liu, J.; Xu, H.; An, Y. A controlled red phosphorus@Ni–P core@shell nanostructure as an ultralong cycle-life and superior high-rate anode for sodium-ion batteries. Energy Environ. Sci. 2017, 10, 1222–1233. [Google Scholar] [CrossRef]
  252. Muñoz-Márquez, M.A.; Zarrabeitia, M.; Castillo-Martínez, E.; Eguía-Barrio, A.; Rojo, T.; Casas-Cabanas, M. Composition and Evolution of the Solid-Electrolyte Interphase in Na2Ti3O7 Electrodes for Na-Ion Batteries: XPS and Auger Parameter Analysis. ACS Appl. Mater. Interfaces 2015, 7, 7801–7808. [Google Scholar] [CrossRef] [PubMed]
  253. Sarkar, A.; Manohar, C.V.; Mitra, S. A simple approach to minimize the first cycle irreversible loss of sodium titanate anode towards the development of sodium-ion battery. Nano Energy 2020, 70, 104520. [Google Scholar] [CrossRef]
  254. Gangaja, B.; Nair, S.; Santhanagopalan, D. Solvent-controlled solid-electrolyte interphase layer composition of a high performance Li4Ti5O12 anode for Na-ion battery applications. Sustain. Energy Fuels 2019, 3, 2490–2498. [Google Scholar] [CrossRef]
  255. Slamet, P.; Widayatno, W.B.; Subhan, A.; Prihandoko, B. Synthesis of Na2Ti3O7–based anode for sodium-ion battery using solid state reaction method. IOP Conf. Series Mater. Sci. Eng. 2018, 432, 012058. [Google Scholar] [CrossRef]
  256. Qi, Y.; Gao, W.; Wang, H.; Liu, D.; Deng, J.; Guo, B.; Bao, S.; Xu, M. Na3(TiOPO4)2F microspheres as a long-life anode for Na-ion batteries. Chem. Eng. J. 2020, 402, 126118. [Google Scholar] [CrossRef]
  257. Ma, J.; Zhang, Z.; Mentbayeva, A.; Yuan, G.; Wang, B.; Wang, H.; Wang, G. Enhanced electrochemical performance of hollow heterostructured carbon encapsulated zinc metastanate microcube composite for lithium-ion and sodium-ion batteries. Electrochim. Acta 2019, 312, 31–44. [Google Scholar] [CrossRef]
  258. Galceran, M.; Rikarte, J.; Zarrabeitia, M.; Pujol, M.C.; Aguilό, M.; Casas-Cabanas, M. Investigation of NaTiOPO4 as Anode for Sodium-Ion Batteries: A Solid Electrolyte Interphase Free Material? ACS Appl. Energy Mater. 2019, 2, 1923–1931. [Google Scholar] [CrossRef]
  259. Brennhagen, A.; Cavallo, C.; Wragg, D.S.; Vajeeston, P.; Sjåstad, A.O.; Koposov, A.Y.; Fjellvåg, H. Operando XRD studies on Bi2MoO6 as anode material for Na-ion batteries. Nanotechnology 2022, 33, 185402. [Google Scholar] [CrossRef]
  260. van Dinter, J.; Grantz, D.; Bitter, A.; Bensch, W. A Combined Sodium Intercalation and Copper Extrusion Mechanism in the Thiophosphate Family: CuCrP2S6 as Anode Material in Sodium-Ion Batteries. ChemElectroChem 2022, 9, e202200018. [Google Scholar] [CrossRef]
  261. Wang, C.; Wang, L.; Li, F.; Cheng, F.; Chen, J. Bulk Bismuth as a High-Capacity and Ultralong Cycle-Life Anode for Sodium-Ion Batteries by Coupling with Glyme-Based Electrolytes. Adv. Mater. 2017, 29, 1702212. [Google Scholar] [CrossRef]
  262. Chen, H.-C.; Patra, J.; Lee, S.-W.; Tseng, C.-J.; Wu, T.-Y.; Lind, M.-H.; Chang, J.-K. Electrochemical Na+ storage properties of SnO2/graphene anodes in carbonate-based and ionic liquid electrolytes. J. Mater. Chem. A 2017, 5, 13776–13784. [Google Scholar] [CrossRef]
  263. Li, C.-Y.; Patra, J.; Yang, C.-H.; Tseng, C.-M.; Majumder, S.B.; Dong, Q.-F.; Chang, J.-K. Electrolyte Optimization for Enhancing Electrochemical Performance of Antimony Sulfide/Graphene Anodes for Sodium-Ion Batteries—Carbonate-Based and Ionic Liquid Electrolytes. ACS Sustain. Chem. Eng. 2017, 5, 8269–8276. [Google Scholar] [CrossRef]
  264. Tang, J.; Kye, D.K.; Pol, V.G. Ultrasound-assisted synthesis of sodium powder as electrode additive to improve cycling performance of sodium-ion batteries. J. Power Sources 2018, 396, 476–482. [Google Scholar] [CrossRef]
  265. Matsumoto, K.; Hwang, J.; Kaushik, S.; Chen, C.-Y.; Hagiwara, R. Advances in sodium secondary batteries utilizing ionic liquid electrolytes. Energy Environ. Sci. 2019, 12, 3247–3287. [Google Scholar] [CrossRef] [Green Version]
  266. Wang, C.-H.; Yang, C.-H.; Chang, J.-K. Suitability of ionic liquid electrolytes for roomtemperature sodium-ion battery applications. Chem. Commun. 2016, 52, 10890–10893. [Google Scholar] [CrossRef]
  267. Sangster, J.; Pelton, A.D. The Na-Si (Sodium-Silicon) System. J. Phase Equil. 1992, 13, 67–69. [Google Scholar] [CrossRef]
  268. Ellis, L.D.; Wilkes, B.N.; Hatchard, T.D.; Obrovac, M.N. In Situ XRD Study of Silicon; Lead and Bismuth Negative Electrodes in Nonaqueous Sodium Cells. J. Electrochem. Soc. 2014, 161, A416–A421. [Google Scholar] [CrossRef]
  269. Bian, H.; Zhang, J.; Yuen, M.-F.; Kang, W.; Zhan, Y.; Yu, D.Y.W.; Xu, Z.; Li, Y.Y. Anodic nanoporous SnO2 grown on Cu foils as superior binder-free Na-ion battery anodes. J. Power Sources 2016, 307, 634–640. [Google Scholar] [CrossRef]
  270. An, C.; Yuan, Y.; Zhang, B.; Tang, L.; Xiao, B.; He, Z.; Zheng, J.; Lu, J. Graphene Wrapped FeSe2 Nano-Microspheres with High Pseudocapacitive Contribution for Enhanced Na-Ion Storage. Adv. Energy Mater. 2019, 9, 1900356. [Google Scholar] [CrossRef]
  271. Kulova, T.L.; Skundin, A.M. The Use of Phosphorus in Sodium-Ion Batteries (A Review). Russ. J. Electrochem. 2020, 56, 1–17. [Google Scholar] [CrossRef]
  272. Liu, S.; Zhang, H.; Xu, L.; Ma, L.; Chen, X. Solvothermal preparation of tin phosphide as a long-life anode for advanced lithium and sodium ion batteries. J. Power Sources 2016, 304, 346–353. [Google Scholar] [CrossRef]
  273. Kulova, T.L.; Kudryashova, Y.O.; Kuz’mina, A.A.; Skundin, A.M.; Stenina, I.A.; Chekannikov, A.A.; Yaroslavtsev, A.B.; Libich, J. Study of degradation of Na2Ti3O7-based electrode during cycling. J. Solid State Electrochem. 2019, 23, 455–463. [Google Scholar] [CrossRef]
  274. Wang, Y.; Zhao, F.; Qian, Y.; Ji, H. High-Performance P2-Na0.70Mn0.80Co0.15Zr0.05O2 Cathode for Sodium-Ion Batteries. ACS Appl. Mater. Interfaces 2018, 10, 42380–42386. [Google Scholar] [CrossRef] [PubMed]
  275. Ren, Q.-Q.; Yu, F.-D.; Zheng, L.-L.; Yin, B.-S.; Wang, Z.-B.; Ke, K. Spinel (Ni0.4Co0.4Mn0.2)3O4 nanoparticles as conversion-type anodes for Li- and Na-ion batteries. Ceram. Int. 2019, 45, 7552–7559. [Google Scholar] [CrossRef]
  276. Garg, U.; Rexhausen, W.; Smith, N.; Harris, J.; Qu, D.; Guptasarma, P. Crystal structure stabilization; electrochemical properties; and morphology of P2-type Na0.67Mn0.625Fe0.25Ni0.125O2 for Na-ion battery cathodes. J. Power Sources 2019, 431, 105–113. [Google Scholar] [CrossRef]
  277. Wang, L.; Yan, C.; Wang, Z.; Zhuang, Q. Facile synthesis of Na2FexFe1−x(SO4)2(OH)x material as a cathode for sodium-ion batteries. Ionics 2019, 25, 2509–2518. [Google Scholar] [CrossRef]
  278. Jin, Y.; Le, P.M.; Gao, P.; Xu, Y.; Xiao, B.; Engelhard, M.H.; Cao, X.; Vo, T.D.; Hu, J.; Zhong, L.; et al. Low-solvation electrolytes for high-voltage sodium-ion batteries. Nat. Energy 2022, 7, 718–725. [Google Scholar] [CrossRef]
  279. Cheng, Z.; Mao, Y.; Dong, Q.; Jin, F.; Shen, Y.; Chen, L. Fluoroethylene Carbonate as an Additive for Sodium-Ion Batteries: Effect on the Sodium Cathode. Acta Phys.-Chim. Sin. 2019, 35, 868–875. [Google Scholar] [CrossRef]
  280. Fu, H.; Xia, M.; Qi, R.; Liang, X.; Zhao, M.; Zhang, Z.; Lu, X.; Cao, G. Improved rate performance of Prussian blue cathode materials for sodium ion batteries induced by ion-conductive solid-electrolyte interphase layer. J. Power Sources 2018, 399, 42–48. [Google Scholar] [CrossRef]
  281. Manohar, C.V.; Forsyth, M.; MacFarlane, D.R.; Mitra, S. Role of N-Propyl-N-Methyl Pyrrolidinium bis(trifluoromethanesulfonyl)imide as an Electrolyte Additive in Sodium Battery Electrochemistry. Energy Technol. 2018, 6, 2232–2237. [Google Scholar] [CrossRef]
  282. Wu, F.; Zhu, N.; Bai, Y.; Li, Y.; Wang, Z.; Ni, Q.; Wang, H.; Wu, C. Unveil the mechanism of solid electrolyte interphase on Na3V2(PO4)3 formed by a novel NaPF6/BMITFSI ionic liquid electrolyte. Nano Energy 2018, 51, 524–532. [Google Scholar] [CrossRef]
Figure 1. Schematic of SEI functioning.
Figure 1. Schematic of SEI functioning.
Energies 15 08615 g001
Figure 2. Depth variation in the composition of SEIs formed in different electrolytes. Electrolyte composition: (a)—1.7 M NaFSI in DME, (b)—5.2 M NaFSI in DME, (c)—2.1 M NaFSI in DME:BTFE (1:2).chematic of SEI functioning.
Figure 2. Depth variation in the composition of SEIs formed in different electrolytes. Electrolyte composition: (a)—1.7 M NaFSI in DME, (b)—5.2 M NaFSI in DME, (c)—2.1 M NaFSI in DME:BTFE (1:2).chematic of SEI functioning.
Energies 15 08615 g002
Figure 3. SEI effect on dendrite formation in various electrolytes. (a) a common electrolyte (1 M NaClO4 in EC–PC mixture), (b) high concentration imide electrolyte (4 M NaFSI in DME), (c) high concentration imide electrolyte with the addition of 1% SbF3.
Figure 3. SEI effect on dendrite formation in various electrolytes. (a) a common electrolyte (1 M NaClO4 in EC–PC mixture), (b) high concentration imide electrolyte (4 M NaFSI in DME), (c) high concentration imide electrolyte with the addition of 1% SbF3.
Energies 15 08615 g003
Figure 4. Galvanostatic charge (Na insertion) and discharge (Na extraction) curves of hard carbon electrodes and a scheme for determining reversible (Qrev) and irreversible (Qirr) capacity.
Figure 4. Galvanostatic charge (Na insertion) and discharge (Na extraction) curves of hard carbon electrodes and a scheme for determining reversible (Qrev) and irreversible (Qirr) capacity.
Energies 15 08615 g004
Figure 5. (a) Variation in reversible capacities for hard-carbon electrodes in (a) EC, (b) PC, and (c) BC solution containing 1 mol/dm3 NaClO4 tested at 25 mA/g. (b) Reversible capacities for hard-carbon electrodes in (a) EC:DMC (1:1), (b) EC:EMC (1:1), and (c) EC:DEC (1:1) solution containing 1 mol/dm3 NaClO4 at 25 mA/g. (Reprinted with permission from [66]. Copyright 2011, Wiley Online Library).
Figure 5. (a) Variation in reversible capacities for hard-carbon electrodes in (a) EC, (b) PC, and (c) BC solution containing 1 mol/dm3 NaClO4 tested at 25 mA/g. (b) Reversible capacities for hard-carbon electrodes in (a) EC:DMC (1:1), (b) EC:EMC (1:1), and (c) EC:DEC (1:1) solution containing 1 mol/dm3 NaClO4 at 25 mA/g. (Reprinted with permission from [66]. Copyright 2011, Wiley Online Library).
Energies 15 08615 g005
Figure 6. Voltage profile of the 1st cycle and the concentration of the SEI species obtained by fitting C 1s and F 1s photoelectron lines of hard carbon negative electrode tested in (a) carbonate ester-based (1 m NaPF6 in EC:PC) and (b) ionic-liquid (0.35 m Na[FSI] in [Pyr14][FSI]) electrolyte. (Reprinted with permission from [67]. Copyright 2022, Wiley Online Library, Open Access).
Figure 6. Voltage profile of the 1st cycle and the concentration of the SEI species obtained by fitting C 1s and F 1s photoelectron lines of hard carbon negative electrode tested in (a) carbonate ester-based (1 m NaPF6 in EC:PC) and (b) ionic-liquid (0.35 m Na[FSI] in [Pyr14][FSI]) electrolyte. (Reprinted with permission from [67]. Copyright 2022, Wiley Online Library, Open Access).
Energies 15 08615 g006
Figure 7. Electrochemical performances of the carbon electrodes with and without presodiation. (a) The voltage-capacity curves of the pristine and presodiated carbon electrode for the first charge/discharge cycle. (b) TEM images of presodiated carbon after immersed in a liquid electrolyte before battery cycling. (Reprinted with permission from [107]. Copyright 2019, Wiley Online Library).
Figure 7. Electrochemical performances of the carbon electrodes with and without presodiation. (a) The voltage-capacity curves of the pristine and presodiated carbon electrode for the first charge/discharge cycle. (b) TEM images of presodiated carbon after immersed in a liquid electrolyte before battery cycling. (Reprinted with permission from [107]. Copyright 2019, Wiley Online Library).
Energies 15 08615 g007
Figure 8. (a) Cycling performance (a, b) and Coulombic efficiency (c, d) at a cycling rate of 0.2 C in FEC-containing (a, c) and FEC-free (b, d) NaPF6/EC/DEC electrolytes NaPF6/EC/DEC electrolyte of porous CNF-SnSb electrodes. (Reprinted with permission from [151]. Copyright 2014, Wiley Online Library). (b) Performance of Sn electrode in 1 M NaPF6/DGME electrolyte: voltage profiles with their rate capability (inset), where 1 C equal to 847 mA/g. (Reprinted with permission from [165]. Copyright 2016, Wiley Online Library).
Figure 8. (a) Cycling performance (a, b) and Coulombic efficiency (c, d) at a cycling rate of 0.2 C in FEC-containing (a, c) and FEC-free (b, d) NaPF6/EC/DEC electrolytes NaPF6/EC/DEC electrolyte of porous CNF-SnSb electrodes. (Reprinted with permission from [151]. Copyright 2014, Wiley Online Library). (b) Performance of Sn electrode in 1 M NaPF6/DGME electrolyte: voltage profiles with their rate capability (inset), where 1 C equal to 847 mA/g. (Reprinted with permission from [165]. Copyright 2016, Wiley Online Library).
Energies 15 08615 g008
Figure 9. (a) Galvanostatic charge/discharge curves of Bi@N-C for first three cycles at 1 A/g. (Reprinted with permission from [170]. Copyright 2019, Wiley Online Library). (b) Rate capability at different current rates of Si NP electrodes. (Reprinted with permission from [178]. Copyright 2016, Wiley Online Library).
Figure 9. (a) Galvanostatic charge/discharge curves of Bi@N-C for first three cycles at 1 A/g. (Reprinted with permission from [170]. Copyright 2019, Wiley Online Library). (b) Rate capability at different current rates of Si NP electrodes. (Reprinted with permission from [178]. Copyright 2016, Wiley Online Library).
Energies 15 08615 g009
Figure 10. Change in the discharge capacity of Ge nanorods at galvanostatic cycling with a current density of 125 mA/g in VC-free electrolyte (black squares) and electrolyte with 2% VC (red circles). (Reprinted with permission from [180]. Copyright 2022, MDPI, Open Access).
Figure 10. Change in the discharge capacity of Ge nanorods at galvanostatic cycling with a current density of 125 mA/g in VC-free electrolyte (black squares) and electrolyte with 2% VC (red circles). (Reprinted with permission from [180]. Copyright 2022, MDPI, Open Access).
Energies 15 08615 g010
Figure 11. (a) Rate performance of MoS2@C in DME and EC/DEC electrolytes. (Reprinted with permission from [231]. Copyright 2020, Wiley Online Library). (b) The rate capability of P-FeSe2 and B-FeSe2/G at various current densities. (Reprinted with permission from [270]. Copyright 2019, Wiley Online Library).
Figure 11. (a) Rate performance of MoS2@C in DME and EC/DEC electrolytes. (Reprinted with permission from [231]. Copyright 2020, Wiley Online Library). (b) The rate capability of P-FeSe2 and B-FeSe2/G at various current densities. (Reprinted with permission from [270]. Copyright 2019, Wiley Online Library).
Energies 15 08615 g011
Table 1. SEI composition and irreversible capacity at the first cycle.
Table 1. SEI composition and irreversible capacity at the first cycle.
Electrode MaterialElectrolyteSEI CompositionQc/Qa
1st Cycle
Ref.
HC1 M NaClO4 in EC:DEC (3:7)-350/300[60]
HC1 M NaPF6 in DEC:EC (1:1)NaF, Na2CO3, NaxPFy, ROCO2Na-[69]
1 M NaClO4 in DEC:EC (1:1)NaCl, Na2CO3, NaxPFy, ROCO2Na, PEO-
1 M NaTFSI in DEC:EC (1:1)NaF, ROCO3Na, RONa, PEO-
1 M NaFTFSI in DEC:EC (1:1)NaF, ROCO3Na, RONa, PEO-
1 M NaFSI in DEC:EC (1:1)NaF, Na2SO4, Na2S, ROCO3Na, RONa, PEO-
HC1 M NaPF6 in EC:DMC (1:1) + 3% FECNaF, Na2CO3, R–COONa, NaO2CO-C2H4-OCO2Na320/220[70]
1 M NaTFSI in EC:DMC (1:1) +3% FECNa2CO3, NaF, NaO2CO-C2H4-OCO2Na275/175
HC0.5 M NaBPh4 in DMENaF275/261[71]
0.5 M NaPF6 in DMENaF276/261
0.5 M NaFSA in DMENa2S270/244
0.5 M NaTFSA in DMENaF234/207
1 M NaPF6 in EC:DEC (1:1)NaF271/251
HC1M NaClO4 PC:EC (1:1)Na2CO3, (CH2OCO2Na)2, Na2CO3,
R–OCO2Na
275/190[74]
1 M NaBF4 in tetraglymeR–ONa, -C-O-C-, CH3OCH2CH2O-,
O–CH2
320/278
HC1 M NaODFB in DMERCH2ONa; groups C=O/B–O, groups B–F; B–O, Na2CO3290/175[75]
1 M NaPF6 in DMENaF; Na2CO3; RCH2ONa265/250
HC1 M NaClO4 in EC:PC:DMC (0.45:0.45:0.1)Na2CO3, ROCO2Na, R–CH2–OCO2Na, PEO560/360[77,79]
HC 1 M NaTFSI in PC + 3% FECNa2CO3, NaF, sodium organic salts570/220[78]
HC1M NaClO4 in EC:PC (1:1) + 5% FECPEO; Na2CO3; NaF; groups C–F910/301[80]
1M NaClO4 in TEGDMESodium alkoxides; Na2CO3; Na2CH2R1050/363
HC1 M NaPF6 in PC + 0.5% FECNa2CO3, NaF, Na-O-(C=O)-O-CH2-R,
Na-O-(C=O)-O-R, Na2CO3, NaF
280/240
1 M NaPF6 in PC:EC + 0.5% FECNa-O-(C=O)-O-CH2-R,
Na-O-(C=O)-O-R, NaxPFy, NaxPOyFz
280/240
HC1 M NaPF6 in DME + 0.5% VCNa2CO3, RCO3Na, NaF,
-OCO2CH=CH)n-
210/200
1 M NaPF6 in DEGDME + 0.5% VC
HC1 M NaClO4 in EC:DEC (1:1)Na2CO3, RONa, ROCO2Na, (CH2–CH2–O–)n400/325[101]
HC1 M NaPF6 in EC:DEC (1:2)NaF, Na2CO3, NaxPFyOz335/235[109]
Graphite1 M NaPF6 in EC:DEC (1:1)CH3CH2OCO2Na, (CH2OCO2Na)2), Na2CO3, NaF120/30[120]
1 M NaPF6 in diglymeCH3OCH2CH2ONa, CH3CH2OCH2CH2ONa, Na2CO3, NaF260/160
Sb1 M NaClO4 in PCNa2CO3, NaF, alkylcarbonates714/550[146]
1 M NaPF6 in EC:DMCNa2CO3, NaF, alkylcarbonates740/620
1 M NaPF6 in PC + 5% FECNa2CO3, NaF, alkylcarbonates735/580
Sn4P31 M NaClO4 in EC:PC (1:1) + 5% FECNa2CO3, NaF720/560[244]
1 M NaClO4 in EC:PC (1:1) + 5% FEC + 0.5% in TMSPNa2CO3, NaF, SiF880/680
Na2Ti3O71 M NaClO4) in EC:PC (1:1)Na2CO3, NaCO3R, NaF, NaCl, NaOR, PEO, poly(ethylene oxide)s385/170[252]
FeS1 M NaCF3SO3 in diglymeNa2CO3, Na2CO2R, NaF600/500[230]
MoS2@C1 M NaPF6 in EC:DEC (1:1)RONa, Na2CO3, ROCOONa, RCH2ONa500/450[231]
Na2Ti3O71 NaClO4 in EC:PC + 1 % FECNaF, Na2O, Na2CO3, NaHCO3, ROCO2Na250/200[253]
SnSe21 M LiPF6 in EC:DMC (1: 1)NaF, Na2CO3, polyethers, Se-O750/525[235]
TiO2 1 M NaClO4 in EC:PC (1:1)Na2CO3, ROCO2Na-/200[185]
Ni3S2@NS-CNTs1.0 M NaSO3CF3 in diglymeNaF, Na2CO3, Na2SO3, organic polyethers,
C-Fx components
470/435[236]
TiO21 M NaPF6 EC:DEC (1:1) + 3% VCPolycarbonates, Alkyl carbonates350/160[187]
1 M NaPF6 in diglymePolyethers320/150
Fe3S41.0 M NaClO4 in DMC:EC (1:1)RCH2ONa, Na2CO3, NaOH600/530[237]
NaTiOPO41 M NaClO4 in EC:PC (1:1)Hydrocarbons, alkyl-carbonates, carbonates, NaF, ethers110/80[258]
Li4Ti5O121 M NaPF6 in diglymeRCH2ONa, ROCO3Na, NaF, Na2CO3, R–CO3300/160[254]
1 M NaPF6 in EC:DMC (3:7)NaF, CFx, Li2CO3275/140
Sn1 M NaPF6 in diglymeRCH2ONa, Na2CO3, NaOH, Na2O, NaF850/780[165]
Sn films 1 M NaClO4 in PCNa2CO3, NaCl900/800[166]
Graphite1 M NaFSI in TEGDMENaF, polyethers160/110[141]
Fe2O3NaClO4 in EC:DEC (2:1)NaOH, Na2CO3, alkyl carbonates ROCO2Li-[189]
Sb1 M NaClO4 in PC + 5% FECNaOH, Na2O, NaF, NaCl770/600[157]
rGO1 M sodium triflate (NaOTf) in (EC-DEC (1:1))Na2CO3, Na2CO2R, polyesters, RSO3Na, NaF1200/500[143]
1 M sodium triflate (NaOTf) in diglymCF3, Na2CO3, Na2CO2R, polyesters, RSO3Na, NaF700/650
Bi1M NaPF6 in diglymeR-COO-Na, RCH2ONa, Na2CO3, sodium alkycarbonates, polyesters420/400[261]
meso-porous Co3O41 M NaPF6 in EC:DEC + 5% FECR-CH2-OCO2Na, R-CH2-OCO2Na, Na2CO3, NaOH, NaF790/750[212]
1 M NaPF6 in FEC:DECR-CH2-OCO2Na, R-CH2-OCO2Na, Na2CO3, NaOH, NaF820/750
Na3V2(PO4)30.25 M NaPF6-incorporated in 1-butyl-3-methylimidazolium bis(trifluoromethanesulfonyl) imide (BMITFSI IL)NaF, NaOH, Na2SO4, Na2S2O7108/135[282]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kulova, T.L.; Skundin, A.M. Electrode/Electrolyte Interphases of Sodium-Ion Batteries. Energies 2022, 15, 8615. https://doi.org/10.3390/en15228615

AMA Style

Kulova TL, Skundin AM. Electrode/Electrolyte Interphases of Sodium-Ion Batteries. Energies. 2022; 15(22):8615. https://doi.org/10.3390/en15228615

Chicago/Turabian Style

Kulova, Tatiana L., and Alexander M. Skundin. 2022. "Electrode/Electrolyte Interphases of Sodium-Ion Batteries" Energies 15, no. 22: 8615. https://doi.org/10.3390/en15228615

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