You are on page 1of 11

Progress in Natural Science: Materials International xxx (xxxx) xxx

H O S T E D BY Contents lists available at ScienceDirect

Progress in Natural Science: Materials International


journal homepage: www.elsevier.com/locate/pnsmi

Research progress on electrolyte key salts for sodium-ion batteries


weimin Zhao a, *, Miao Wang a, Haichen Lin b, Kangwoon Kim b, Rongkai He c, Shijie Feng b,
Haodong Liu b, **
a
College of Chemical Engineering and Safety, Binzhou University, Binzhou, 256503, PR China
b
Center for Memory and Recording Research Building, University of California San Diego, 9500 Gilman Dr, San Diego, CA 92093, USA
c
STATE GRID Liaocheng PowerSupply Company, Liaocheng, 252000, PR China

A R T I C L E I N F O A B S T R A C T

Keywords: Sodium-ion batteries (SIBs) are considered potential successors to lithium-ion batteries in the fields of energy
Sodium ion batteries storage and low-speed vehicles, thanks to their advantages such as abundant raw material sources, high energy
Electrolyte density, and a wide operational temperature range. However, several scientific and engineering challenges still
Electrolyte salts
require attention in the development of sodium-ion batteries. Electrolyte salts, as a key component of sodium-ion
Electrochemical performance
battery electrolytes, play a critical role in battery performance. This paper provides a brief overview of the
research progress on different electrolyte salt systems in sodium-ion batteries. It discusses characteristics such as
ionic conductivity, electrochemical windows, electrochemical performance, and thermal safety in various solvent
systems. Furthermore, the paper summarizes a series of strategies for controlling electrolyte and electrode in-
terfaces, offering references for addressing the challenges in the mass production and application of sodium-ion
batteries.

1. Introduction Earth's crust is only about 0.002% [16,17], far from satisfying the
pressing requirements of the energy transformation and upgrade. In
The development of human society is closely intertwined with energy contrast, sodium's abundance in the Earth's crust is approximately 2.36%,
[1]. However, the extensive consumption of fossil fuels, such as oil, has and sodium salts can be extracted from seawater, ensuring evenly
led to a severe energy crisis and environmental pollution [2,3], signifi- distributed resources [18]. Furthermore, sodium-ion compounds are
cantly impacting economic and social development and the human living readily accessible, maintain price stability, and are cost-effective. Uti-
environment. Consequently, building a new, renewable, and clean en- lizing aluminum as the cathode current collector instead of the more
ergy infrastructure has emerged as a strategic objective for countries expensive copper foil reduces the overall cost of sodium-ion batteries by
worldwide. China, in particular, has set ambitious targets to peak carbon 30–50% compared to lithium-ion batteries [19,20]. Sodium-ion batteries
emissions before 2030 and achieve carbon neutrality by 2060 [4–6]. are widely regarded as the most promising electrochemical energy stor-
According to statistics from the carbon emission estimation database age devices for large-scale storage applications in the post-lithium era
Carbon Monitor, in 2021, the power and transportation sectors accoun- and have emerged as a leading direction in electrochemical energy en-
ted for a significant share of global carbon emissions, approximately gineering research and application development [21,22].
38.9% and 20.3%, respectively [7,8]. In light of the dual carbon policy,
adjusting the energy landscape through renewable and clean energy 2. Fundamentals of battery electrolytes
generation, such as wind and solar power [9–12], and employing elec-
trochemical energy storage and conversion devices for electricity storage, The working principle of sodium-ion batteries is akin to that of
can effectively mitigate carbon emissions [13,14]. Secondary recharge- lithium-ion batteries [23–25]. They were introduced almost simulta-
able chemical power sources play a pivotal role in this energy revolution. neously in the 1970s, primarily because sodium and lithium belong to the
Presently, the predominant secondary batteries are lithium-ion batteries, same main group in the periodic table and possess the same number of
relying on graphite anodes [15]. However, lithium's abundance in the outer electrons [26], signifying their similar chemical properties. A

* Corresponding author.
** Corresponding author.
E-mail address: zhaoweimin137@sina.com (Zhao).

https://doi.org/10.1016/j.pnsc.2024.03.003
Received 23 December 2023; Received in revised form 8 March 2024; Accepted 11 March 2024
Available online xxxx
1002-0071/© 2024 Published by Elsevier B.V. on behalf of Chinese Materials Research Society.

Please cite this article as: w. Zhao et al., Research progress on electrolyte key salts for sodium-ion batteries, Progress in Natural Science: Materials
International, https://doi.org/10.1016/j.pnsc.2024.03.003
Zhao et al. Progress in Natural Science: Materials International xxx (xxxx) xxx

sodium-ion battery consists of key components, including the cathode, Consequently, the type of anion in sodium salts significantly impacts
anode, separator, electrolyte, and current collector [27–29]. The struc- the ESW of the electrolyte. For instance, in a mixed solvent of
ture and performance of the cathode and anode materials govern the EC/DEC, the oxidation potentials follow this order: NaPF6 > NaClO4
battery's sodium storage capabilities. The separator, acting as an insu- > NaTFSI > NaFSI, with the PF6- ion having the lowest HOMO level
lator between the cathode and anode, prevents short circuits [30]. The (11.67 eV), making it less prone to electron loss and decomposition.
electrolyte serves as the medium for ion transport, while the current Conversely, the ClO 4 ion possesses an HOMO level at 7.89 eV,
collector facilitates electron transmission. During the charging process, suggesting lower chemical stability during oxidation. Nevertheless,
Naþ ions are extracted from the cathode, traverse through the electrolyte it's essential to understand that chemical stability is relative, and
and separator, and intercalate into the anode [31,32]. This leaves the there's no absolutely stable anion in SIBs. The anions in sodium salts
cathode in a high-potential state with less sodium, and the anode in a can participate in the formation of the solid electrolyte interface (SEI)
low-potential state with more sodium. The discharge process reverses [42]. For instance, decomposition products of PF 6 ions during
this, with Naþ ions de-intercalating from the anode, passing through the reduction, such as NaF, constitute the primary components of SEI. In
electrolyte and separator, and re-intercalating into the cathode material, summary, the anions in sodium salts can influence the chemical sta-
restoring the cathode to a state with more sodium. To maintain charge bility of the electrolyte in two ways: first, the HOMO level of the anion
balance, an equal number of electrons traverse the external circuit during constrains the highest electrochemical window of SIBs; second, the
both charging and discharging, shuttling between the cathode and anode LUMO level of the anion facilitates the formation of SEI to inhibit
alongside Naþ ions, resulting in oxidation and reduction reactions at the further electrolyte decomposition.
electrodes. The cathode material primarily supplies sodium ions and c. Ion Dissociation State and Migration Rate Issues: This encompasses
determines the battery's energy density. Presently, there are three pri- factors like the interaction between anions and cations, solvent
mary routes for cathode materials: layered transition metal oxides, dielectric constant, viscosity, freezing point, and their effects on Naþ
Prussian blue analogs, and polyanion compounds [33]. The anode ma- ions and anions. Sodium salts exhibit smaller diffusion rates, more
terial serves as the carrier for ions and electrons during sodium-ion intricate chemical reactions, and less stable solid electrolyte inter-
battery charge and discharge cycles, dictating energy storage and phase (SEI) films at the anode when compared to lithium salts [43].
release. Carbon-based materials are the preferred choice for the anode Furthermore, the solvation structures influenced by different sodium
[34]. Due to the instability of compounds formed when sodium inter- salt anions dictate the de-solvation kinetics of Naþ ions and the
calation into graphite layers, which is challenging with conventional evolution of the interface film. Notably, the weak interaction between
graphite materials. Currently, anode materials suitable for sodium-ion Na þ ions and PF 6 promotes sodium desolvation and storage kinetics.
batteries encompass amorphous carbon, metal compounds, and alloy Solvation structures involving PF 6 induce anion-prioritized decom-
materials. Because of the significant volume change and poor cycling position, resulting in a thin, inorganic-rich electrode-electrolyte
performance of alloy materials, and the low capacity of metal com- interface phase that enhances interface stability and limits excessive
pounds, hard carbon and soft carbon with amorphous structures are the electrolyte solvent decomposition [44]. This, in turn, ensures elec-
predominant anode materials [35]. Soft carbon can be fully graphitized trode stability and facilitates charge transfer kinetics.
at high temperatures, providing excellent conductivity. Hard carbon d. Compatibility with Cathode, Anode Materials, and current collectors:
boasts advantages like high sodium storage capacity, low sodium inter- for instance, anions such as FSI and TFSI have the potential to
calation potential, high specific capacity, and ease of synthesis. Its ca- corrode the aluminum current collector at high voltages. However,
pacity in sodium batteries (200- 450mAh g1) rivals that of graphite in this corrosion can be mitigated by incorporating additives, such as
lithium batteries (200- 450mAh g1), rendering it a more versatile oxalate, to passivate the current collector within the electrolyte.
choice. e. Influence of Anion Type and Concentration: Due to interactions be-
The electrolyte, often referred to as the 'lifeblood' of the battery, tween cations and anions and variations in anion sizes, different so-
serves as the conduit linking the positive and negative electrodes and dium salts in the same solvent exhibit discrepancies in conductivity.
facilitates ion conduction within the battery [36]. Notably, the electro- The concentration of sodium salts also significantly impacts electro-
lyte exerts a crucial influence on the performance of the electro- lyte conductivity [45]. Typically, in systems with lower sodium salt
de/electrolyte interface and significantly affects battery characteristics, concentrations, the electrolyte's conductivity gradually increases
including capacity, internal resistance, charge and discharge rate, oper- until it reaches a maximum value with increasing sodium salt con-
ating temperature, storage, and even the safety of lithium-ion batteries centration, after which it starts to decline [46]. For instance, in an
during overcharging and over-discharging. Typically, the electrolyte EC/DMC solvent, the conductivity of NaPF increases to a maximum of
comprises sodium salts, solvents, and additives. Among these, sodium 6.8 mS cm1 at a salt concentration of 0.6 mol L1, and then expe-
salts primarily supply Naþ ions in the electrolyte, thereby impacting the riences a slight decrease with the addition of more salts [45]. It's
power of battery, cycling performance, capacity, and safety. Sodium salts well-established that at lower salt concentrations, free solvent mole-
must meet fundamental criteria, including high solubility, robust thermal cules predominate. Consequently, adding more salt increases ionic
stability, cost-effectiveness, and environmental compatibility [37–39]. conductivity [45]. However, higher salt concentrations notably
Simultaneously, several other aspects necessitate consideration. elevate the viscosity of the electrolyte, thereby impacting the elec-
trochemical stability of the electrolyte with highly concentrated so-
a. Chemical Stability: It is widely recognized that even trace amounts of dium salts.
water in the electrolyte can react with PF 6 ions to generate Lewis
acids such as HF and POF3. During sodium deintercalation, these 3. Electrolytes salts
chemicals react with transition metal oxides, resulting in metal
dissolution and an accelerated rate of irreversible reactions within the The sodium salts commonly employed in sodium-ion battery elec-
battery. trolytes primarily include sodium hexa-fluorophosphate (NaPF6), sodium
b. Electrochemical Stability: Sodium salts dissolved in non-aqueous perchlorate (NaClO4), and bis-(trifluoromethanesulfonyl)imide sodium
solvents should possess a wide electrochemical window. To main- (NaTFSI) [47]. Among these, NaPF6 and NaClO4 are recognized as two
tain the thermodynamic stability of sodium-ion batteries, the elec- key salts with significant industrial application potential. NaPF6, owing
trolyte's electrochemical window (ESW) should exceed the oxidation- to its synthesis method resembling that of LiPF6 and its compatibility
reduction potentials of both the anode and cathode. ESW is defined as with lithium-ion battery production processes and equipment, has
the difference between the lowest unoccupied molecular orbital emerged as one of the leading electrolyte salts for sodium-ion batteries
(LUMO) and the highest occupied molecular orbital (HOMO) [40,41]. (SIBs). On the other hand, NaClO4 boasts advantages such as rapid ionic

2
Zhao et al. Progress in Natural Science: Materials International xxx (xxxx) xxx

migration, robust chemical stability, impressive thermal stability, and electrolyte and materials and excessive solid electrolyte interphase (SEI)
cost-effectiveness, rendering it a preferred choice among electrolyte salts formation, ultimately deteriorating battery performance. To address
for SIBs. Sodium salts containing the sulfonyl fluoride group (NaTFSI, these challenges and further enhance battery performance, the addition
NaFTFSI, NaFSI, etc.) exhibit higher thermal stability and non-toxicity of certain film-forming additives to the electrolyte has proven effective.
but are seldom used independently due to the corrosive impact of their Komaba et al. [51], for instance, conducted battery tests with different
anions on aluminum foil current collectors. Moreover, when selecting additives in a 1 M NaClO4– PC electrolyte, including fluorinated ethylene
sodium salts for electrolytes, it is imperative to consider toxicity concerns carbonate (FEC), trans-difluoroethylene carbonate (DFEC), ethylene sul-
and potential harm to human health. Sodium salts based on AsF 6 and fonate (ES), and vinylene carbonate (VC). They observed that FEC could
SbF6 are rarely employed. In essence, the choice of sodium salts should significantly improve the cycle stability of the NaNi1/2Mn1/2O2||hard
encompass not only the intrinsic physical and chemical properties of the carbon full battery. This enhancement was attributed to FEC preferential
sodium salts themselves, including viscosity, conductivity, and thermal decomposition on the hard carbon surface, forming a dense SEI layer rich
stability but also their compatibility with electrodes when mixed with in NaF, which inhibited the reduction decomposition of PC and thus
solvents. This includes considerations of electrochemical stability and contributed to improved battery performance.
thermal stability under various operational conditions [48].The author's In the quest to suppress sodium metal dendrites, Li [52] developed an
review encompasses recent research on the application of both inorganic innovative ultra-thin (8 μm) external non-porous separator composed of
and organic electrolyte salts in sodium-ion batteries, while also offering a honeycomb structure of fibers. As shown in Fig. 1(a) schematic of the
insights into the anticipated trends in the use of sodium salts in SEI compositions formed in different electrolytes and a comparison of the
sodium-ion battery electrolytes. charge transfer energy barriers. This honeycomb-like fiber structure
exhibited an exceptional electrolyte absorption rate (376.7%) and
3.1. NaClO4 inherent polymer transmission capability, enabling unimpeded ion
transfer. They utilized a 1.0 M NaClO4-EC/PC electrolyte with 5.0 wt%
NaClO4 offers several advantages, including high ionic conductivity, FEC as an additive, employing polyethersulfone SPF and polyvinylidene
excellent compatibility with electrodes, and cost-effectiveness, which fluoride PVDF-HFP coated separators. This setup resulted in the assembly
have led to its extensive study as one of the earliest sodium salts. In 1980, of a high-performance NaNi1/3Fe1/3Mn1/3O2||hard carbon pouch bat-
Delmas et al. made a pivotal discovery, demonstrating that the P2-type tery, achieving a remarkable energy density of 229.4 Wh kg1 and 157.6
cathode material NaCoO2 could exhibit remarkable electrochemical Wh L1, along with excellent cycling stability. Additionally, Liu [53]
performance when used with a PC-based NaClO4 electrolyte [49]. Sub- employed ab initio molecular dynamics (AIMD) simulations and reactive
sequently, in 2000, Stevens employed a 1 M NaClO4- EC/DEC (3:7, V/V) molecular dynamics (RMD) to investigate the NaClO4– PC system with 2
electrolyte, achieving the reversible intercalation of Naþ ions in hard % FEC. This research elucidated the solvent decomposition process, the
carbon materials at room temperature [50]. However, it's important to formation of key SEI components (such as NaF, NaOH, and Na2CO3), and
note that the stability of the interfacial film formed by NaClO4- based potential reaction mechanisms. The study provided insights into the
electrolytes can be susceptible to the strong oxidizing nature of per- formation process of crucial components like OH and CO2 3 within the
chlorates. This can lead to the detachment of some interfacial films from SEI layer. The computational results confirmed that NaOH, NaF, and
the electrode surface, resulting in renewed contact between the Na2CO3 are indeed inorganic constituents of the SEI, aligning with

Fig. 1. (a)Schematic of the SEI compositions formed in different electrolytes and a comparison of the charge transfer energy barriers [52]. (b) (b-1)SEM images of the
(Ni0.5Mn0.5)(OH)2 precursor and (b-2) of the final Na(Ni0.5Mn0.5)O2 compound [54]. (c) Schematic of the SnC/NaNi0.5Mn0.5O2 full cell [54]. (d) Schematic illustration
of the sodium ion storage process in charging [55].

3
Zhao et al. Progress in Natural Science: Materials International xxx (xxxx) xxx

experimental findings. a gel-state polymer solid polymer electrolyte (SPE) membrane. XRD
In addition to research on additives, optimizing solvent composition spectra of M3STFSI and the SPE membrane demonstrated efficient Na þ
represents a crucial avenue of investigation. Alcantara et al. [56] con- ion diffusion in the ionic liquid and polymer membrane. By utilizing a
ducted a study using hard carbon as the negative electrode and NaClO4 as ratio of 8.51:0.25:0.26 mM, they achieved high ionic transference
the electrolyte salt. They compared the electrochemical performance of numbers (~0.26) and substantial ionic conductivity (~0.58 mS cm1),
hard carbon in different solvents or solvent combinations, including coupled with an electrochemical stability window of 4.25 V. After 70
EC/DMC (1:1, v/v), dimethyl ether (DME), tetrahydrofuran (THF), and cycles at 0.1C, the discharge capacity remained at 75 mAh g1, with a
EC/THF (1:1, v/v). Their findings revealed that THF could effectively coulombic efficiency exceeding 99%. These results offer valuable guid-
enhance the electrochemical performance of the hard carbon negative ance for the potential commercialization of low-cost and high-safety
electrode. It is important to note that these solvent systems all exhibited batteries [60].
an initial capacity decline trend. Furthermore, Oh et al. [54] reported the In summary, NaClO4 presents several advantages, including its
characteristics and performance of a sodium-ion battery based on a Sn–C affordability and impressive electrochemical and air stability, which
anode and a Na(Ni0.5Mn0.5)O2 cathode especially in the micrometer scale make it convenient for transport and storage. Consequently, it is often
spherical morphology optimized by Na(Ni0.5Mn0.5)O2, it can lead to a preferred in the development of new electrolyte formulations. However,
continuous increase in the vibrational density, about 2.5 g cm3. High its high oxidizing properties raise safety concerns, posing a challenge to
vibrational density is a very popular characteristic for battery electrodes, its widespread commercial use. Moreover, anhydrous NaClO4, being a
as it reflects high volumetric energy density. The SEM images in Fig. 1 potentially explosive compound, is subject to regulatory controls. Future
reveal that the co-precipitated Na(Ni0.5Mn0.5)O2 hydroxide precursor research on NaClO4 should prioritize enhancing the safety of commercial
powder exhibits a spherical morphology (Fig.b-1) that persisted also in batteries utilizing this electrolyte system.
the final Na(Ni0.5Mn0.5)O2 compound obtained by sodiation at the high
temperature of 820  C (Fig.b-2). Fig. 1(c) shows the scheme of the Sn– 3.2. NaPF6
C/NaClO4 PC-FEC/Na(Ni0.5Mn0.5)O2 sodium-ion battery developed in
this work. They replaced PC with EMC as a solvent and incorporated 2 % In a study by Eshetu [61], NaPF6 exhibits minimal mass loss even at
FEC as a film-forming additive, with NaClO4 as the electrolyte salt. This temperatures up to 300  C, demonstrating its high thermal stability.
electrolyte demonstrated an impressive electrochemical stability window Nevertheless, studies have shown that NaPF6 is highly hygroscopic and
of up to 5.6 V and an ionic conductivity of up to 6.0 mS cm1. Despite readily undergoes hydrolysis to produce NaF, HF, and other phosphate
NaClO4 thermal stability at 310  C, it often contains crystalline water, species (such as POF3). Remarkably, this hydrolysis can still take place in
resulting in higher water content in the electrolyte system. Additionally, battery-grade electrolytes with <20 ppm of trace water, highlighting the
its intrinsic oxygen content contributes to its strong oxidizing properties, highly hygroscopic nature of NaPF6. Consequently, this means that the
posing safety concerns and hindering its practical commercial-scale use. quality of commercially purchased NaPF6 may be variable, and it de-
To address these issues, researchers have shown significant interest in pends on the degree of hydrolysis that has occurred (either through
gel electrolytes and ionic liquid-like electrolytes based on NaClO4. Wu storage or as a result of the synthetic route used to make it). This property
et al. [55] reported a highly porous PVdF- HFP gel polymer electrolyte allows NaPF6 to be compatible with various positive and negative elec-
using 1 M NaClO4, with a thermal stability temperature of up to 130  C trode materials [62].
(matching that of the commercial separator Celgard-2730). This elec- Kim [63] used NaPF6- DME as an electrolyte to test the electro-
trolyte exhibited a room temperature ionic conductivity of 0.60 mS cm1, chemical performance of graphite anodes in SIB systems, maintaining a
a broader electrochemical window (4.6 V), and excellent mechanical stable capacity of 150mAh g1 after 2500 cycles. Among the sodium salts
properties, including a tensile strength of 7.6 MPa. As shown in Fig. 1(d), (NaPF6, NaCF3SO3, and NaClO4) with three different anions used in DME
schematic illustration of the sodium ion storage process in charging. solvents in the alloy negative electrode, only NaPF6 demonstrated stable
Goodenough [57] designed a sodium-ion battery featuring Sb as the cycling performance, while NaCF3SO3 and NaClO4 proved incompatible
anode, Na3V2(PO4)3 as the cathode, and a cross-linked PMMA composite with the alloy negative electrode. Lucht [64] utilized NaPF6 as the
gel polymer as the electrolyte. This gel polymer employed NaClO4 as the electrolyte salt and EC/DEC (1:2, v/v) as the solvent, resulting in an
electrolyte salt and achieved an electrochemical window of 4.8 V, a room initial coulombic capacity of 162 mAh g1 for hard carbon materials
temperature ionic conductivity of up to 6.2 mS cm1, and an assembled within the 0.05–2 V charge-discharge range. By employing attenuated
battery with impressive discharge capacities of 106.8 mAh g1 at 0.1C total reflection infrared spectroscopy (IR-ATR) and X-ray photoelectron
and 61.1 mAh g1 at 10C, showcasing its potential for large-scale spectroscopy (XPS), the study revealed that the primary components of
applications. the SEI film formed on the hard carbon anode included NaF, along with
Additionally, Lai et al. [58] demonstrated a polymer electrolyte minor quantities of alkyl carbonates and Na2CO3. This suggests that the
comprising NaClO4 as the electrolyte salt, silane-modified β- Al2O3 SEI components and reaction mechanisms in sodium-ion batteries share
powder, and a crosslinked PVDF- HFP matrix. This electrolyte design fundamental similarities with those in lithium-ion batteries [64]. As
aimed to suppress side reactions at the electrode/electrolyte shown in Fig. 2, the sodium salt anion participates in the solvation
inorganic-organic interface, thereby enhancing electrochemical perfor- structure, where PF 6 preferentially decomposes to form a stable inor-
mance. Results indicated that the composite electrolyte maintained an ganic compound-rich position-electrolyte interface phase (CEI), inhibit-
ionic conductivity of 1.37 mS cm1 at 20  C, with a sodium ion trans- ing solvation and ensuring interfacial stability. Furthermore, it's worth
ference number reaching 0.424. Furthermore, the sodium symmetric cell noting that the interaction between Naþ and PF 6 is significantly weaker
based on this composite electrolyte displayed robust cyclic performance, than the interaction between Naþ and ClO 4 , which promotes the
exceeding 800h of stable cycling at 0.5 mA cm2 without significant de-solvation of Naþ, improves charge transfer kinetics [63]. The
voltage polarization. This underscored its effective protective role for the Na4Fe3(PO4)2P2O7 (NFPP)||Na battery, using NaPF6 electrolyte, exhibi-
sodium anode, with the assembled Na3V2(PO4)3||hard carbon battery ted a high capacity retention rate of 81.2 % after 800 cycles at 4.3 V.
retaining 92.2% of its capacity after 1000 cycles at 3  C, providing a Additionally, it achieved a high specific capacity of 70 mAh g1 at a high
valuable strategy for constructing highly stable sodium-ion batteries. rate of 20C.
To mitigate the exothermic reactions associated with highly volatile In order to construct a stable electrode/electrolyte interface and
organic solvents and address dendrite formation on the negative elec- enhance the cycling performance of the battery, Che [65] incorporated
trode, Guzm an et al. [59] synthesized an ionic liquid, M3STFSI, through three additives (FEC, PST, DTD) into a 1 M NaPF6-PC/EMC-based elec-
ion exchange within a trimethylthorium-bis(trifluoromethylsulfonyl) trolyte. The combination of these three additives resulted in the forma-
imide/sodium perchlorate/polyepoxide ternary system. They prepared tion of a SEI layer rich in organic components and a dense CEI layer at the

4
Zhao et al. Progress in Natural Science: Materials International xxx (xxxx) xxx

Fig. 2. Electrochemical performance of NFPP based batteries: (a) initial charge and discharge curve of NFPP/Na battery, (b) cycle performance of NFPP/Na battery,
(c, d) corresponding charge and discharge curve, (e) rate performance, (f) cycle performance of NFPP/hard carbon battery, and (g, h) Corresponding charge-discharge
curve (1C ¼ 129 mAh g1).

Fig. 3. Electrochemical performance of Na2Ti3O7@C/Na half-cells containing different electrolytes:(a) cycle curve,(b) first charge and discharge curve,(c) rate
performance,(d) radar diagram.

5
Zhao et al. Progress in Natural Science: Materials International xxx (xxxx) xxx

electrode interface, contributing to improved cycling stability. The CEI center such as the carbon atom of the ether solvent, leading to the
generated by NaPF6-E was more uniform and much thinner (4 nm) than breakage of chemical bonds. As illustrated in Fig. 4(d), schematic dia-
the CEI generated by NaClO4-E, indicating that the PF6-anion-induced gram of decomposition pathways of ether solvent (DEGDME). Tarascon
CEI layer remained relatively stable even at 4.3 V cutoff voltage [72] used 1.0 M NaPF6-DEGDME as the electrolyte. They achieved a
(Fig. 4(a), (a-1)). In contrast, the CEI generated by NaClO4-E is uneven specific capacity of 768.0 mAh g1 after 100 cycles for micron-sized Sn
and thicker (>10 nm), with a large uncovered area on the cathode sur- electrodes, with a capacity retention rate of 88.0 %. The authors attrib-
face (Fig. 4(b), (b-1)), which may be due to severe interfacial reactions uted the improved cycling stability of the battery to the evolution of
occurring on the cathode surface. Fig. 4(c) shows the action mechanism micron-sized Bi into a porous structure after cycling in the ether elec-
of NaPF6 and NaClO4 electrolytes. Ponrouch et al. [66] systematically trolyte. This structural transformation was found to be related to the
studied the viscosity, ionic conductivity, electrochemistry, and thermal contact wetting angle between the ether electrolyte and the Bi electrode.
stability of the electrolyte after the combination of different solvents (PC, Seh [73] reported a highly reversible sodium deposition process in
EC, DMC, DME, DEC, Diglyme, and Triglyme) or their mixtures with ether electrolytes, with a long cycle life of over 300 cycles at a current
different sodium salts (NaClO4, NaPF6, and NaTFSI). The results showed density of 0.5 mA cm2 and an average coulombic efficiency of 99.9 %.
that the binary EC:PC solvent mixture is considered the best solvent Further cryo-EM and XPS results suggested that the high reversibility of
formulation for sodium-ion batteries. The combination of NaClO4 or metallic sodium was attributed to the formation of a stable and thin SEI.
NaPF6 with binary EC:PC solvents promotes the formation of a stable SEI Qin et al. [74] conducted a comparison of the electrochemical perfor-
layer, leading to excellent reversible capacity and good magnification mance between a 1 M NaPF6-DGME electrolyte and a typical carbonate
performance of the hard carbon electrode. In the study of co-intercalation electrolyte (1 M NaPF6 in PC with 5.0 wt% FEC as the additive), using the
behavior with 1 M NaPF6 in PC or FEC for various CFx materials, Chen high-voltage poly-anion material Na3V2(PO4)2F3 as the cathode. When
[67] found that 1 M NaPF6 in FEC is conducive to solvent using the 1 M NaPF6-DGME electrolyte, they observed a capacity reten-
co-intercalation. Among them, F-CF0.88 cathodes exhibited a tion rate as high as 96.2% after 300 cycles at 0.5C, which significantly
discharge-specific capacity of 1094 mAh g1 for the first time, main- surpassed the 77.5 % observed with 1 M NaPF6 PC-FEC. Furthermore, the
taining a stable rechargeable specific capacity of 344 mAh g1. This capacity retention rate remained at 95.0 % after 500 cycles at 1C and
improved rechargeable behavior is attributed to solvent co-intercalation, reached 98.8 % at 20C after 1000 cycles, demonstrating excellent cycling
mainly due to the larger solvation energy of fluoroethylene carbonate stability. The authors attribute the higher cycling efficiency in the NaPF6-
(FEC), making it easier to embed. This study offers new insights and a ether electrolyte to the absence of large agglomerates in the material
deeper understanding of solvent co-intercalation in CFx as a Na-ion particles, resulting in reduced electrochemical polarization.
rechargeable battery. In contrast, the poor cycling performance in the carbonate electrolyte
Moreover, Yan [70] prepared a composite electrolyte formulation by results from the high solubility of Na2CO3 produced by oxidation, which
adding additives VC, butanedinitrile (SN), 1,3-propanesultone (PS), and fails to stabilize the positive electrode-electrolyte interface. To effectively
sodium difluoroborate (NaDFOB) into the baseline electrolyte (1.0 M mitigate side reactions and the continuous decomposition of the elec-
NaPF6-EC/PC). These additives synergistically formed stable SEI and CEI trolyte caused by active sites (defects, pores, and functional groups) on
layers on the hard carbon anode and Na3V2(PO4)2F3 cathode, enhancing the surface of the hard carbon anode, Yu [69] engineered an amorphous
cycling performance at 55  C. Using EIS spectroscopy, XPS, and theo- Al2O3 layer on the hard carbon (HC) surface. This layer suppressed the
retical calculations, they confirmed the key role of NaF and sulfate de- continuous decomposition of NaPF6 in carbonate solvents, reducing the
posits at the interface. Considering the special process of accumulation of NaF. As a result, it formed a thinner, denser solid elec-
oxidation-reduction in ether-based electrolytes, Xu [71] designed a trolyte interface, effectively reducing interface resistance. This straight-
NaPF6-DME electrolyte with a sodium salt concentration of 2 M. This forward and effective surface engineering method significantly increased
design aimed to reduce the co-intercalation voltage of the graphite the initial coulombic efficiency of the hard carbon anode, raising it from
anode, resulting in a higher average working voltage (3.1 V) and a long 64.7 % to 81.1 %. Furthermore, the optimized HC- Al2O3 exhibited a
cycle life (1000 cycles) for the Na1.5VPO4.8F0.7||graphite batteries. The higher reversible capacity, reaching 321.5 mAh g1 at 50.0 mA g1,
combination of ether electrolytes and DOL presents a promising solution along with excellent cycling stability, retaining 89.3 % of its capacity
for achieving high-performance full cells [68]. For ether solvents, the after 2000 cycles at 1.0 mA g1. Fig. 4(e) shows the schematic diagram of
electrons generated during the discharge process attack an electrophilic how homotype heterogeneous HC-Al2O3 material improves ICE.

Fig. 4. (a, b) TEM images of the cycled NFPP cathode of NFPP/Na cells: (a, a-1) NaPF6-E, (b, b-1) NaClO4-E [65]. (c) Schematic diagrams of the mechanism of action of
different electrolytes [65]. (d) Schematic diagram of decomposition pathways of (d) ether solvent (DEGDME) [68]. (e) The schematic diagram of Al2O3 layers
improving ICE of HC electrode. HC, hard carbon; ICE, initial Coulombic efficiency [69].

6
Zhao et al. Progress in Natural Science: Materials International xxx (xxxx) xxx

Homotype refers to both HC and Al2O3 as amorphous materials, while the Song et al. [77] developed a 1.2 M NaFSI electrolyte blend, incor-
disorder degree of HC is significantly higher than that of Al2O3 porating DME and BTFE to create a locally high salt concentration elec-
In order to enhance the wide temperature range operation of sodium- trolyte formulation. This formulation effectively suppressed the
ion batteries, Li [75] utilized Na4Fe3(PO4)2P2O7 (NFPP) as the cathode formation of rock salt on the surface of NaNi0.68Mn0.22Co0.1O2. They
and natural graphite (NG) as the anode. They employed a 1.0 M NaPF6 assembled a NaNi0.68Mn0.22Co0.1O2||hard carbon full cell with a high
electrolyte dissolved in an ether-based solvent to create a low-cost load exceeding 2.5 mAh cm2, and this cell maintained 82% of its ca-
NFPP||NG battery. This battery achieved a discharge specific capacity pacity after 450 cycles. Yan et al. [78] demonstrated an electrolyte based
of 153.2 mAh g1 at 0.2 A g1, maintained a discharge capacity of on NaFSI with triethyl phosphate (TEP) and 1,1,2,2-tetrafluoroethyl-2,2,
90.8mAh g1 even at a discharge current density of 10 A g1, and 3,3-tetrafluoropropyl ether (TTE) as solvents. They assembled a
exhibited a capacity retention rate of 60%. Notably, during the charging NaCu1/9Ni2/9Fe1/3Mn1/3O2||hard carbon full cell. Cryo-EM analysis
process, it took only 32.6 s to reach a 60% capacity. Even after 1000 confirmed the presence of an inorganically rich solid electrolyte interface
cycles at room temperature, the battery still exhibited an outstanding (SEI), which reduced side reactions between hard carbon and the elec-
capacity retention rate of 81%. The excellent performance of the NFPP|| trolyte. Additionally, a thin but robust cathode-electrolyte interface (CEI)
NG battery across a wide temperature range was attributed to the high inhibited the dissolution of transition metals and surface reconstruction.
boiling point of 162.1  C and low freezing point of 72.0  C of the NaPF6 As a result, the full cell exhibited high safety and a long cycle life, with a
diglycol ether electrolyte. This, coupled with the solvent co-intercalation capacity retention rate of 82.5 % after 200 cycles. More recently, Yang
process at the anode and the high Naþ diffusion rate of the NFPP elec- et al. [79] proposed a similar electrolyte formula, combining NaTFSI with
trode material, allowed the battery to achieve discharge capacities of trimethyl phosphate (TMP), BTFE, and VC for Prussian blue||hard carbon
148.7 and 130.3 mAh g1 at 20  C and 40  C, respectively. pouch cells. Leveraging hydrogen bond regulation, they achieved the
Remarkably, even at the ultra-low temperature of 70.0  C, the NFPP|| formation of an optimized SEI. This allowed the full cell to maintain a
NG battery operated effectively, maintaining a discharge capacity of 61.9 high capacity retention rate of over 85 % after 120 cycles.
mAh g1. Furthermore, the high boiling point of the electrolyte and the In typical sodium-ion electrolytes, there are three main sodium ion
excellent thermal stability of the electrode materials enabled the entire solvation structures present, including solvent-separated ion pairs, con-
battery to operate at elevated temperatures, reaching as high as 130.0  C. tact ion pairs, and aggregates. However, in high salt concentration
At this temperature, the battery maintained a discharge capacity of 129.8 electrolytes, the predominant structures are aggregates and contact ion
mAh g1 and a coulombic efficiency of 93.0 %. Therefore, the full battery pairs. As a result, most solvents are coordinated, and anions are likely to
can operate in a wide temperature range from 70.0  C to 130.0  C [76]. coordinate with two or more sodium ions, leading to the formation of an
Darren demonstrated a new synthesis method for NaPF6. This anion-based solid electrolyte interface (SEI). The high solubility of NaFSI
involved the reaction of ammonium hexafluorophosphate with sodium and NaTFSI in carbonate/ether solvents has opened up possibilities for
metal in THF solvent, resulting in the production of electrolyte salt free designing and implementing high concentration electrolyte or locally
from common impurities found in commercial NaPF6. The NaPF6 salt was high concentration electrolyte systems in sodium-ion battery applica-
dissolved in a binary mixture of ethylene carbonate: diethyl carbonate tions. Patra et al. [80] reported a 3.0 M NaFSI-EC/PC electrolyte with an
(EC:DEC ¼ 1:1,v/v) solvent. It's worth noting that when the concentra- ionic conductivity of 6.3 mS cm1, a viscosity of 22 cP, and relatively low
tion of NaPF6 reached up to 3.5 M, the electrolyte exhibited a very high flammability. Notably, this medium concentration electrolyte facilitated
viscosity, a gel-like appearance, and contained undissolved NaPF6. The the formation of an organic-inorganic composite component on the hard
maximum electrolyte conductivity, measured at 358  C, was obtained carbon electrode. The initial discharge coulombic efficiency and average
with a 1 M electrolyte solution, reaching 19.0 mS cm1. This conductivity coulombic efficiency of the hard carbon anode reached 85.2% and
value was higher than those observed for 0.5 M, 2.0 M, and 3.0 M >99.9%, respectively. Even after 500 cycles, the capacity loss was only
electrolyte solutions, which recorded conductivities of 8.2, 16.1, and 5.0%. Compared to the thicker SEI layer (~20 nm) formed in the 1.0 M
11.0 mS cm1, respectively. Electrolyte conductivity and the aging of the NaPF6-EC/PC electrolyte, the SEI thickness remained at ~5 nm.
native SEI were measured in Na–Na symmetric cells within the concen- Hwang et al. [81] utilized a molten salt system to enhance the solu-
tration range of 1.0–3.0 M using electrochemical impedance spectros- bility of sodium salts in organic solvents and introduced a composite
copy. For ultraconcentrated (>2.0 M) electrolytes, the interface electrolyte salt consisting of 5.0 M NaFTFSI-NaFSI-NaOTF in PC as the
impedance was approximately 300 W (6.3 mS cm1), while for 1.0 M solvent. This innovative electrolyte effectively mitigated the corrosion of
electrolyte, it measured around 500 W (3.8 mS cm1) at the same storage the aluminum current collector and exhibited a wide electrochemical
time.Notably, the interface stabilization is slower for the 1.0 M electro- stability window. When assembled in a NaCrO2||hard carbon full cell, it
lyte, and this phenomenon may be attributed to differences in SEI com- maintained an impressive capacity retention rate of 86.9% after 100
positions, structures, or dissolution rates. cycles at 1.0C. In a parallel study, Wen et al. [82] conducted a compar-
ative analysis to assess the compatibility of NaCF3SO3 and NaClO4 as
3.3. Imides electrolyte salts with a Na2Ti3O7 anode. Their findings indicated a sig-
nificant difference in first-cycle coulombic efficiency. Na2Ti3O7@C||Na
Compared to traditional inorganic sodium salts like NaClO4 and NaPF6, cells using a NaClO4-based electrolyte achieved a first-cycle coulombic
sodium salts containing the fluorosulfonyl group (NaTFSI, NaFTFSI, NaFSI, efficiency of 55.5 %, while Na2Ti3O7@C||Na cells using a NaCF3SO3--
etc.) offer higher thermal stability and non-toxicity. However, they are based electrolyte achieved a substantially higher first-cycle coulombic
rarely used alone as sodium salts, primarily due to their anions' corrosive efficiency of 76.5 %, as illustrated in Fig. 3. The authors attributed this
effect on aluminum foil current collectors and the high cost of synthesizing improvement in coulombic efficiency to the decomposition of NaCF3SO3
these types of sodium salts. As mentioned before, the electrolyte, as a in carbonate solvents. This decomposition produced additional
crucial component of the battery, significantly influences battery perfor- sulfur-containing compounds that exhibited favorable sodium conduc-
mance. Even in the same organic solvent system, when the concentration tivity and stability, ultimately enhancing the interface transport prop-
of the electrolyte salt reaches a certain level (3.0 M), referred to as high erties of the material and inhibiting the increase in solid electrolyte
concentration electrolytes (HCE), the electrolyte exhibits unique properties interface (SEI) film thickness. As a result, this contributed to improved
not typically found in traditional low-concentration electrolytes. High cycling and rate performance of the battery.
concentration electrolytes have demonstrated excellent electrochemical Fan et al. [83] utilized a NaCF3SO3-DEGDME-based electrolyte to
performance, including reduced side reactions, an expanded working effectively enhance the first-cycle coulombic efficiency to 96.2% and
temperature range, a broader electrochemical window, improved rate achieve a long cycle specific capacity of 341.7 mAh g1 at 0.2 A g1 for
performance, and enhanced cycling stability. 250 cycles when using NiCo2S4 as a sodium-ion battery anode material.

7
Zhao et al. Progress in Natural Science: Materials International xxx (xxxx) xxx

Their research, complemented by techniques like DFT, in situ XRD, and phosphate (TEP). These solvents have higher polarity compared to car-
other characterization methods, substantiated the optimization impact of bonates or ethers, which enhances the solubility of NaBOB. In NMP,
NaCF3SO3 on the formation process of sulfide intermediates. Further- NaBOB can achieve a dissociation of 0.66 M, resulting in an ionic con-
more, Li et al. [84] conducted a comparative study to evaluate the so- ductivity of 8.83 mS cm1. Although there is limited literature on the
dium storage performance of high surface area carbon materials (HSAC) utilization of NaBOB in batteries, it has demonstrated an initial
in various electrolytes. Their findings revealed that the NaCF3SO3-- coulombic efficiency of up to 82 % when used with hard carbon anodes.
DEGDME electrolyte demonstrated the highest initial cycle coulombic During cycling voltammetry on TMP-coated aluminum foil with NaBOB
efficiency at 69.28 % and delivered a reversible capacity of 283 mAh g1. as the electrolyte salt, the reduction potential is approximately 1.5 V vs.
This particular electrolyte exhibited relatively superior electrochemical Naþ/Na, and the oxidation potential is around 3.0 V. Since NaBOB de-
performance when compared to others. composes at the positive electrode to form a passivation layer, it can
Wang et al. [85] discovered that the positive electrode, NNMO, extend the electrolyte window to as high as 4.0 V. In a full cell comprising
exhibited superior performance in a 1.0 M NaPF6 diethylene glycol hard carbon and Prussian white, a capacity retention rate of 57.0% was
dimethyl ether (G2) - based electrolyte. In this G2 electrolyte, NNMO achieved after 1000 cycles. When using TEP, a capacity retention rate of
achieved a reversible capacity of 97.5 mAh g1 at 2000 mA g1, 80 % was achieved after 2000 cycles. However, it's important to note that
compared to only 48.9 mAh g1 in a PC-based electrolyte. The discharge the high viscosity of TEP can lead to significant power decay and reduced
capacity remained at 140 mAh g1 under a 50 mA g1 discharge current, rate performance in the battery.
with a retention rate of 94.6 % after 100 cycles. The G2 electrolyte also Similar to NaBOB, other salts like NaBSB have been explored for their
demonstrated excellent rate performance in the 1.0V to 4.0 V charge and solubility and conductivity characteristics. For NaBSB, it was found to
discharge range. This enhanced performance is attributed to the forma- have solubility exceeding 1.0 mol kg1 in GBL, but its ionic conductivity
tion of a dense and stable cathode-electrolyte interface (CEI) layer rich in was relatively low at 2 mS cm1 at 1 mol kg1. In a button cell with
Na–F and C–O species, facilitating Naþ diffusion and transport channels chromium oxide as the cathode and metallic Na, the cell exhibited a
on the NNMO positive electrode during cycling. higher charging cutoff voltage during constant current cycling [88].
Another noteworthy salt, sodium tetraphenylborate (NaBPh4), has
3.4. Other-type sodium salts gained attention due to its strong anti-reduction properties. NaBPh4 is
highly soluble in ether solvents, with an ionic conductivity of up to 6.0
Boron-based anion sodium salts often exhibit superior electro- mS cm1 at 0.5 M. When used in a full cell with hard carbon and
chemical and thermal stability, while effectively improving the film- Na2V3O7, it demonstrated good cycling stability and high-rate perfor-
forming properties of the electrolyte on the electrode surface, resulting mance. However, its oxidative stability is limited to 3.4 V, which may
in enhanced cycling stability. Notably, NaBF4 demonstrates high ionic pose challenges with certain cathode materials. Cyan-based sodium salts,
conductivity at 0.9833 mS cm1 and finds extensive use in ionic liquids such as NaTIM, NaPCPI, and NaTCP, show promise in terms of solubility
[86]. In a study by Du and colleagues, the performance of two different and ionic conductivity [89]. These cyan-substituted anions have delo-
electrolytes was investigated in Na2Ti3O7@C half cells: a 1.0 M NaBF4 - calized negative charges, leading to higher solubility and conductivity.
diglyme electrolyte and a 1.0 M NaClO4-EC/PC electrolyte (1:1, v/v) For example, the conductivity of NaTIM and NaTCP in EC/DMC at 0.75
[86]. The results revealed that the battery employing the 1.0 M NaBF4- mol kg1 reached 16.53 and 13.53 mS cm1, respectively, surpassing the
diglyme electrolyte exhibited a first-cycle coulombic efficiency of 73.0% 10.2 mS cm1 of NaClO4 in PC. Additionally, these cyan-based salts
and lower SEI film impedance. Importantly, this electrolyte formulation exhibit excellent thermal stability, remaining stable at temperatures
also proved suitable for use with hard carbon anodes. exceeding 500  C. In terms of electrochemical stability, NaTCP and
In comparison to NaBF4, NaODFB stands out as a novel chelating NaTIM show promise with oxidative stability exceeding 4 V when used in
sodium salt, featuring an anion comprising half BOB and half BF 4 , as a PC solvent, making them attractive for various applications.
exemplified by B(C2O4)F 2 . Research indicates that in carbonate elec- Bitner-Michalska et al. [90] investigated the use of Na0.68CoO2 as a
trolytes, NaODFB forms a dense and smooth SEI film on the cycled so- cathode material in an electrolyte consisting of 0.75 M NaTDI-EC/DMC
dium anode, effectively mitigating sodium dendrite growth. NaODFB (1:1) at C/10 for 50 cycles of constant current charge and discharge.
boasts several outstanding properties, including compatibility with The initial capacity was 80.68 mAh g1, and after 50 cycles, the capacity
common SIB solvents, a wider electrochemical window than NaClO4 and was well-maintained, with a discharge capacity of 63.0 mAh g1 on the
NaPF 6 based electrolytes, impressive rate and cycling performance in 50th cycle (more than 78.0% of the first cycle). This research suggests
Na0.44MnO2||Na half cells compared to commercially available sodium that NaTDI is compatible with the positive electrode material and can be
salts, and excellent chemical stability, with no production of harmful successfully applied as an electrolyte in sodium-ion batter-
gases upon exposure to air or reaction with trace water [86]. ies.Plewa-Marczewska et al. [91] examined electrolytes containing new
The electrochemical performance of NaODFB-based electrolytes was sodium salts and found that both NaTDI and NaPDI electrolytes exhibited
investigated in NVP cathode materials. Specifically, a NaODFB-based excellent ionic conductivity. The highest conductivity was observed at
diglyme electrolyte was employed in NVP||HC button cells. The use of concentrations of 0.5 M and 1.0 M in the tested temperature range. For
1.0 M NaODFB diglyme electrolyte contributed to the cycling stability of example, at 20  C, the ionic conductivity was 4.0 mS cm1 for both
NVP||Na half cells, resulting in a specific capacity of 95 mA h g1 and a NaTDI (0.5 M, 3.71 mS cm1; 1.0 M, 3.78 mS cm1) and NaPDI (0.5 M,
capacity retention rate of 86.0 % after 100 cycles at room temperature at 3.79 mS cm1; 1.0 M, 3.83 mS cm1). NaTDI exhibited higher conduc-
0.5C. This improvement can be attributed to the formation of a thinner tivity than the most conductive concentration (0.5 M and 1.0 M) at
CEI film on the NVP material surface after cycling in the 1.0 M NaODFB temperatures above 40  C. However, 1.5 M samples showed lower con-
diglyme electrolyte, along with a low B2O3 content in the battery, ductivity for both salts.
ensuring favorable cycling stability. It's worth noting that NaBOB salts Bitner-Michalska and Michalczewski et al. [92] observed that the
exhibit desirable thermal stability (>300  C) and chemical stability in Sb2O3/C composite anode achieved an initial capacity of 584.0 mAh g1
air, making them effective in passivating aluminum current collectors. at C/10 constant-current charge and discharge in 0.75 M NaTDI EC/DMC
However, NaBOB is susceptible to hydrolysis with water, resulting in the electrolyte. Ether solvents, which have been rarely used in lithium-ion
production of oxalic acid and other insoluble compounds. Additionally, batteries, have gained traction in sodium-ion batteries due to their ad-
NaBOB has extremely low solubility in most organic solvents [87]. vantages. They form thinner SEI films, exhibit solvent intercalation ef-
Regarding the practical application of NaBOB, its use is currently fects, and facilitate rapid sodium ion migration. Ether-based solvents
restricted to amide solvents like N-methyl-2-pyrrolidone (NMP) and enable highly reversible embedding and release of sodium ions, resulting
phosphate esters such as trimethyl phosphate (TMP) and triethyl in denser, thinner, and more stable SEI films. This significantly reduces

8
Zhao et al. Progress in Natural Science: Materials International xxx (xxxx) xxx

sodium ion diffusion lengths, enhances sodium ion transmission effi- efficiency of up to 93.8 % in the first cycle when using SPE. In contrast,
ciency, and prevents additional electrolyte decomposition on the nega- the control groups showed lower reversible capacities (89 and 97 mAh
tive electrode surface. As a result, when using ether-based electrolytes, g1), initial coulomb efficiencies (84.8 and 88.4%), and higher polari-
graphite negative electrodes can provide high coulomb efficiency, stable zation (88 mV and 91 mV). Batteries utilizing SPE (FSI–Al2O3-AQ)
capacity of 100 mAh g1, and long cycling performance up to 1000 cycles maintained high and stable coulomb efficiency (~100%) even up to the
[93]. Reber et al. [94] employed Na(FSI)0.71(FTFSI)0.29 electrolytes at 2000th cycle. The reversible capacity of NVP|FSI–Al2O3-AQ||Na at
concentrations of 20.0 M, 25.0 M, 30.0 M, or 35.0 M in NVPOF||NTP and various rates (0.1C, 0.2C, 0.5C, 1.0C, 2.0C, and 5.0C) was 110.0, 103.7,
NaMnHCF||NTP full battery configurations. Among these, 30.0 M and 100.6, 95.2, 91.5, and 76.8 mAh g1, respectively, while the other two
35.0 M concentrations exhibited superior capacity retention rates. The control SPEs exhibited poor performance. The Al2O3 suspension in
hybrid electrolyte, utilizing the highly fluid TFSI anion, achieved different solutions is centrifuged and washed six times with the corre-
coulomb efficiencies of 98–99% (NTP-NVPOF) and 96–98 % sponding solvent to filter out the soluble salt, and then dried. TEM images
(NTP-NaMnHCF). This innovative electrolyte enabled the stable opera- of the resulting powder (Fig. 5(e)) provided visual evidence that the
tion of a 2V sodium-ion battery, featuring the NaTi2(PO4)3/Na2Mn Al2O3 powder did react with NaFSI aqueous solution according to the
[Fe(CN)6] electrode pair, at 1C for 300 cycles. The battery demonstrated obvious morphological changes of Al2O3 in FSI-AQ. Completely trans-
a coulomb efficiency exceeding 99.5 % and an energy density up to 77 formed from the original nanorods >20 nm in length (Al2O3– P) to
Wh kg1. nanoparticles less than 10 nm. The crystal plane spacing of the outer
Liu et al. [95] reported that solid-state sodium batteries employing a layer of the particle is about 0.2027 nm (Fig. 5(f)). The XPS spectrum of
PEO/NaFSI–Al2O3-AQ electrolyte and Na3V2(PO4)3 as the active material Al2O3–FSI-AQ is shown in Fig. 5(g), and the fitted Al2p spectrum shows
demonstrated excellent reversibility, high rate capability, long-term two peaks associated with Al–O (73.8 eV) and Al– F(75 eV). Sun et al.
cycle stability, and high coulomb efficiency. Comparative experiments [96] studied the interface characteristics of carbon anode in sodium ion
evaluating the electrochemical performance of SPE (FSI1-Al2O3-AQ) battery (nib) when it was in contact with super concentrated sodium
were conducted, with control groups using SPE (FSI1-Al2O3-AN) and (fluoromethane) sulfonimide (NaFSI)/N-methyl-n-propyl pyridinium
TFSI-Al2O3-AQ. In Fig. 5, the NVP|FSI–Al2O3-AQ||Na coin battery, (fluoromethane) sulfonimide (C3mpyrFSI). In order to more fundamen-
charged and discharged at 0.1C between 2.5 and 3.8 V at 80  C, exhibited tally understand the effect of the SEI layer on the interfacial dynamics,
a highly reversible specific capacity of 110 mAh g1 and a coulomb temperature dependent EIS measurements were performed to reveal the

Fig. 5. Electrochemical performance of the cells NVP|SPEs|Na with FSI - Al2O3- AQ, FSI - Al2O3- AN, or TFSI-Al2O3-AQ as he electrolyte at 80  C. (a) First charge and
discharge curves at the 0.1C rate. (b) Coulombic efficiencies at the 1C rate. (c) Typical charge and discharge curves at the 1C rate. (d) Rate performance from 0.1C to
5C. (e) Optical photographs and TEM images of the Al2O3 particles filtrated out from NaFSI aqueous solution (Al2O3–FSI-AQ), NaFSI acetonitrile solution (Al2O3–FSI-
AN), and NaTFSI aqueous solution (Al2O3-TFSI-AQ) after stirring at 55  C for 12 h followed by washing six times [95]. (f) High resolution TEM image of Al2O3–FSI-AQ
[95]. (g) XPS spectra of Al2p and F1s of the Al2O3–FSI-AQ [95].(h) Comparison of new sodium salt NaCF3SO3 and NaClO4 at the Na2Ti3O7/electrolyte interface layer.
(i)Schematic illustration of Na solvation and diffusion across the SEI layer and CMK bulk material, respectively [96]. -

9
Zhao et al. Progress in Natural Science: Materials International xxx (xxxx) xxx

Naþ diffusion dynamics. The weak binding energy associated with the [9] C. Wang, F. Wang, Structural decomposition analysis of carbon emissions and policy
recommendations for energy sustainability in Xinjiang, Sustainability 7 (6) (2015)
high coordination number 5 in the 50 mol% NaFSI system can promote
7548–7567.
the Na ion jumping process, resulting in high Naþ mobility, and also [10] K. Ooka, T. Kajiwara, S. Aihara, et al., Control of rotating magnetic flux distribution
facilitate the desolvation ability, as shown in Fig. 5(i). A long cycle life of in a transformer model core by laser irradiation, IEEE Trans. Magn. 50 (4) (2014)
the cell, up to 3500 cycles with 320 mAh g1, is achieved, which is 1–4.
[11] X. Liu, R.C. Bansal, Optimizing combustion process by adaptive tuning technology
contributed by the anion-derived inorganic solid electrolyte interphase based on integrated genetic algorithm and computational fluid dynamics, Energy
(SEI) layer on the anode [97]. Convers. Manag. 56 (2012) 53–62.
[12] X. Su, C. Gao, M. Cheng, et al., Controllable synthesis of Ni (OH) 2/Co (OH) 2
hollow nanohexagons wrapped in reduced graphene oxide for supercapacitors, RSC
4. Conclusion advances 6 (99) (2016) 97172–97179.
[13] Y. Tang, S. Chen, S. Mu, et al., Synthesis of capsule-like porous hollow nanonickel
Over the past few decades, substantial strides have been taken in cobalt sulfides via cation exchange based on the Kirkendall effect for high-
performance supercapacitors, ACS Appl. Mater. Interfaces 8 (15) (2016)
lithium-ion and sodium-ion battery research, leading to their growing 9721–9732.
market share in the energy storage sector. However, the development [14] H.F. Wang, C. Tang, Q. Zhang, A review of precious-metal-free bifunctional oxygen
and enhancement of sodium-ion battery electrolyte systems have pre- electrocatalysts: rational design and applications in Zn air batteries, Adv. Funct.
Mater. 28 (46) (2018) 1803329.
dominantly centered on two sodium salt systems, NaClO4 and NaPF6, [15] L.H. Lai, S.T. Shiue, Effects of acetylene/ammonia mixtures on the properties of
with other sodium salts primarily serving as additives to improve elec- carbon films prepared by thermal chemical vapor deposition, Surf. Coating.
trolyte performance. To achieve superior cycling performance in sodium- Technol. 215 (2013) 161–169.
[16] L.A. Powell, R.L. Tease, Determination of calcium, magnesium, strontium, and
ion battery (SIB) systems, further comprehensive investigations are
silicon in brines by graphite furnace atomic absorption spectrometry, Anal. Chem.
warranted in the following areas: 1) The application of theoretical cal- 54 (13) (1982) 2154–2158.
culations, in-situ spectroscopy, and similar techniques to delve deeper [17] Z. Jian, L. Zhao, H. Pan, et al., Carbon coated Na3V2 (PO4) 3 as novel electrode
into the solvation structure of electrolyte systems and the compatibility material for sodium ion batteries, Electrochem. Commun. 14 (1) (2012) 86–89.
[18] A. Brennhagen, C. Cavallo, D.S. Wragg, et al., Understanding the (De) sodiation
of the electrode/electrolyte interface; 2) The creation of novel, highly mechanisms in Na-based batteries through operando X-ray methods, Batteries &
stable sodium salts and electrolyte additives to expand the electro- Supercaps 4 (7) (2021) 1039–1063.
chemical stability window of the electrolyte, mitigate excessive electro- [19] N. Yabuuchi, K. Kubota, M. Dahbi, et al., Research development on sodium-ion
batteries, Chemical reviews 114 (23) (2014) 11636–11682.
lyte decomposition, and adapt to high-voltage cathode materials, thereby [20] Y. Zhao, Q. Fu, D. Wang, et al., Co9S8@ carbon yolk-shell nanocages as a high
further boosting the energy density of sodium-ion battery systems; 3) performance direct conversion anode material for sodium ion batteries, Energy
Streamlining the synthesis methods for sodium salts to reduce production Storage Mater. 18 (2019) 51–58.
[21] Fangxi Xie, L. Zhang, C. Ye, et al., The application of hollow structured anodes for
costs of NaPF6, NaClO4, NaTFSI, NaFSI, and others, ultimately enhancing sodium-ion batteries: from simple to complex systems, Adv. Mater. 31 (38) (2019)
the cost-effectiveness of sodium-ion batteries. Addressing these chal- 1800492.
lenges will propel sodium-ion battery technology forward and enhance [22] W.J. Li, Q.R. Yang, S.L. Chou, et al., Cobalt phosphide as a new anode material for
sodium storage, J. Power Sources 294 (2015) 627–632.
its competitiveness within the energy storage landscape. [23] Y. Tang, S. Zheng, Y. Xu, et al., Advanced batteries based on manganese dioxide and
its composites, Energy Storage Mater. 12 (2018) 284–309.
Declaration of competing interest [24] Z. Tong, Y. Qi, J. Zhao, et al., One-step synthesis of carbon-coated Na 3 (VOPO 4) 2
F using biomass as a reducing agent and their electrochemical properties, Waste and
Biomass Valorization 11 (2020) 2201–2209.
The authors declare no competing financial interest. [25] Y. Liu, X. Liu, T. Wang, et al., Research and application progress on key materials
for sodium-ion batteries, Sustain. Energy Fuels 1 (5) (2017) 986–1006.
Data availability [26] J. Ye, J. Zang, Z. Tian, et al., Sulfur and nitrogen co-doped hollow carbon spheres
for sodium-ion batteries with superior cyclic and rate performance, J. Mater. Chem.
A 4 (34) (2016) 13223–13227.
Data will be made available on request. [27] D.E. Demirocak, S.S. Srinivasan, E.K. Stefanakos, A review on nanocomposite
materials for rechargeable Li-ion batteries, Applied sciences 7 (7) (2017) 731.
[28] T. Dolker, D. Pant, Bioremediation of metals from lithium-ion battery (LIB) waste,
Acknowledgments Waste Bioremediation (2018) 265–278.
[29] C. Liu, R. Masse, X. Nan, et al., A promising cathode for Li-ion batteries: Li3V2 (PO4)
This work was financially supported by Shandong Provincial Natural 3, Energy Storage Mater. 4 (2016) 15–58.
[30] J. Alessi, E. Beebe, C. Carlson, et al., A hollow cathode ion source for production of
Science Foundation (Grant No. ZR202011050003) primary ions for the BNL electron beam ion source, Rev. Sci. Instrum. 85 (2) (2014).
[31] S. Li, Y.F. Dong, D.D. Wang, et al., Hierarchical nanowires for high-performance
References electrochemical energy storage, Frontiers of Physics 9 (2014) 303–322.
[32] H.Y. Park, K.P. Singh, D.S. Yang, et al., Simple approach to advanced binder-free
nitrogen-doped graphene electrode for lithium batteries, Rsc Advances 5 (5) (2015)
[1] H. Han Shengjuan, Jingping Zhu, Research on the dynamic relationship of the
3881–3887.
Energy-Economy-Environment (3E) system-based on an empirical analysis of China,
[33] S. Guo, Q. Li, P. Liu, et al., Environmentally stable interface of layered oxide
Energy Proc. 5 (2011) 2397–2404.
cathodes for sodium-ion batteries, Nat. Commun. 8 (1) (2017) 135.
[2] R. Lin, J. Cheng, Z. Yang, et al., Enhanced energy recovery from cassava ethanol
[34] L.S. Roselin, R.S. Juang, C.T. Hsieh, et al., Recent advances and perspectives of
wastewater through sequential dark hydrogen, photo hydrogen and methane
carbon-based nanostructures as anode materials for Li-ion batteries, Materials 12
fermentation combined with ammonium removal, Bioresour. Technol. 214 (2016)
(8) (2019) 1229.
686–691.
[35] W. Li, M. Yue, H. Guo, et al., Rational design of MnS nanoparticles anchored on N,
[3] M. Ohnishi, A. Takeoka, S. Nakano, et al., Advanced photovoltaic technologies and
S-codoped carbon matrix as anode for lithium-ion batteries, Prog. Nat. Sci.: Mater.
residential applications, Renew. Energy 6 (3) (1995) 275–282.
Int. 31 (5) (2021) 649–655.
[4] Y. Liu, Q. Sun, B. Yue, et al., Elucidating solid electrolyte interphase formation in
[36] Y. Huang, L. Zhao, L. Li, et al., Electrolytes and electrolyte/electrode interfaces in
sodium-based batteries: key reductive reactions and inorganic composition,
sodium-ion batteries: from scientific research to practical application, Advanced
J. Mater. Chem. A 27 (11) (2023) 14640–14645.
materials 31 (21) (2019) 1808393.
[5] X. Li, J. Zhang, X. Guo, et al., An ultrathin nonporous polymer separator regulates
[37] T. Gao, B. Wang, L. Wang, et al., LiAlCl4⋅3SO2 as a high conductive, non-flammable
Na transfer toward dendrite-free sodium storage batteries, Adv. Mater. 35 (15)
and inorganic non-aqueous liquid electrolyte for lithium ion batteries, Electrochim.
(2023) 2203547.
Acta 286 (2018) 77–85.
[6] G.L. Xu, R. Amine, A. Abouimrane, et al., Challenges in developing electrodes,
[38] B. Zhang, C. Liu, Y. Liu, A novel one-step approach to synthesize fluorescent carbon
electrolytes, and diagnostics tools to understand and advance sodium-ion batteries,
nanoparticles, EuropeanJournalofInorganic Chemistry 28 (2010) 4411–4414.
Adv. Energy Mater. 8 (14) (2018) 1702403.
[39] B. Wang, W. Al Abdulla, D. Wang, et al., A three-dimensional porous LiFePO 4
[7] X. Shi, Z. Yang, Y. Liu, et al., Strength and corrosion properties of Portland cement
cathode material modified with a nitrogen-doped graphene aerogel for high-power
mortar and concrete with mineral admixtures, Construct. Build. Mater. 25 (8)
lithium ion batteries, Energy Environ. Sci. 8 (3) (2015) 869–875.
(2011) 3245–3256.
[40] J. Wang, J. Zhao, L. Ma, et al., First-principles study of structural evolution of
[8] C. Luederitz, E. Brink, F. Gralla, et al., A review of urban ecosystem services: six key
medium-sized SiN clusters (41 N 50) within stuffed fullerene cages, The European
challenges for future research, Ecosyst. Serv. 14 (2015) 98–112.
Physical Journal D 45 (2) (2007) 289–294.

10
Zhao et al. Progress in Natural Science: Materials International xxx (xxxx) xxx

[41] I. Dance, E. Wenger, H. Harris, Towards rational syntheses of the elusive [70] W. Zhang, Y. Zhang, J. Zhou, et al., Optimizing electrochemical performance in
metallocarbohedrenes: density functional prescriptions for electronic and geometric sodium-ion batteries using O3-type Na0. 90Cu0.22Fe0.30Mn0.48O2 and hard carbon,
structures, Chem.–Eur. J. 8 (15) (2002) 3497–3511. J. Electrochem. Soc. 170 (7) (2023) 070518.
[42] M. Yue, H. Guo, Z. Yuan, et al., Cobalt coordination modified double-schiff-base [71] H. Xu, Q. Yan, W. Yao, et al., Mainstream optimization strategies for cathode
with low solubility as a long-cycle life anode for sodium-ion batteries, Prog. Nat. materials of sodium-ion batteries, Small Structures 3 (4) (2022) 2100217.
Sci.: Mater. Int. 32 (4) (2022) 510–516. [72] W. Li, S.L. Chou, J.Z. Wang, et al., Sn4þxP3@ amorphous Sn-P composites as anodes
[43] C. Lin, L. Ouyang, R. Hu, et al., Synthesis of amorphous SeP2/C composite by for sodium-ion batteries with low cost, high capacity, long life, and superior rate
plasma assisted ball milling for high-performance anode materials of lithium and capability, Advanced materials 26 (24) (2014) 4037–4042.
sodium-ion batteries, Prog. Nat. Sci.: Mater. Int. 31 (4) (2021) 567–574. [73] Z.W. Seh, J. Sun, Y. Sun, et al., A highly reversible room-temperature sodium metal
[44] J. Chen, Y. Gao, C. Li, et al., Interface modification in high voltage spinel lithium- anode, ACS Cent. Sci. 1 (8) (2015) 449–455.
ion battery by using N-methylpyrrole as an electrolyte additive, Electrochim. Acta [74] B. Qin, M. Zarrabeitia, A. Hoefling, et al., A unique polymer-inorganic cathode-
178 (2015) 127–133. electrolyte-interphase (CEI) boosts high-performance Na3V2(PO4)2F3 batteries in
[45] H. Che, S. Chen, Y. Xie, et al., Electrolyte design strategies and research progress for ether electrolytes, J. Power Sources 560 (2023) 232630.
room-temperature sodium-ion batteries, Energy Environ. Sci. 10 (5) (2017) [75] Z. Li, Y. Zhang, Y. Wang, High-power and low-cost sodium-ion batteries with a wide
1075–1101. operation temperature from -70 C to 130 C, Smart Mat 4 (5) (2023) e1191.
[46] K.H. Kamarudin, M.I.N. Mohamad Isa, Ionic conductivity via quantum mechanical [76] Y. Liu, W. Qin, D. Zhang, et al., Effect of Naþ in situ doping on LiFePO4/C cathode
tunneling in NH4NO3 doped carboxymethyl cellulose solid biopolymer electrolytes, material for lithium-ion batteries, Prog. Nat. Sci.: Mater. Int. 31 (1) (2021) 14–18.
Adv. Mater. Res. 1107 (2015) 236–241. [77] X. Song, T. Meng, Y. Deng, et al., The effects of the functional electrolyte additive on
[47] G.G. Eshetu, T. Diemant, M. Hekmatfar, et al., Impact of the electrolyte salt anion on the cathode material Na0.76Ni0.3Fe0.4Mn0.3O2 for sodium-ion batteries, Electrochim.
the solid electrolyte interphase formation in sodium ion batteries, Nano Energy 55 Acta 281 (2018) 370–377.
(2019) 327–340. [78] G. Yan, K. Reeves, D. Foix, et al., A new electrolyte formulation for securing high
[48] R. Martin, J.P. Catinat, J.R. Caille, et al., Polymer Composition, Polymer Membrane temperature cycling and storage performances of Na-ion batteries, Adv. Energy
Comprising the Polymer Composition, Process for Preparing it and Fuel Cell Mater. 9 (41) (2019) 1901431.
Comprising the Membrane: U.S. Patent Application, 2011-1-6, 12/919,072[P]. [79] H. Yang, J. Hwang, Y. Tonouchi, et al., Sodium difluorophosphate: facile synthesis,
[49] C. Delmas, J.J. Braconnier, C. Fouassier, et al., Electrochemical intercalation of structure, and electrochemical behavior as an additive for sodium-ion batteries,
sodium in NaxCoO2 bronzes, Solid State Ionics 3–4 (1981) 165–169. J. Mater. Chem. A 9 (6) (2021) 3637–3647.
[50] D.A. Stevens, J.R. Dahn, The mechanisms of lithium and sodium insertion in carbon [80] J. Patra, P.C. Rath, C. Li, et al., A water-soluble NaCMC/NaPAA binder for
materials, J. Electrochem. Soc. 148 (2001) A803–A811. exceptional improvement of sodium-ion batteries with an SnO2-ordered
[51] S. Komaba, W. Murata, T. Ishikawa, et al., Electrochemical Na insertion and solid mesoporous carbon anode, ChemSusChem 11 (22) (2018) 3923–3931.
electrolyte interphase for hard-carbon electrodes and application to Na-Ion [81] J.J. Fan, P. Dai, C.G. Shi, et al., Synergistic dual-additive electrolyte for interphase
batteries, Adv. Funct. Mater. 21 (20) (2011) 3859–3867. modification to boost cyclability of layered cathode for sodium ion batteries, Adv.
[52] K. Li, J. Zhang, D. Lin, et al., Evolution of the electrochemical interface in sodium Funct. Mater. 31 (17) (2021) 2010500.
ion batteries with ether electrolytes, Nat. Commun. 10 (1) (2019) 725. [82] S. Wen, X. Li, J. Zhang, et al., Effects of sodium salts on compatibility between
[53] J. Zhao, C. Liu, W. Guo, et al., Prediction on the light-assisted exfoliation of Na2Ti3O7@ C anode and electrolyte for sodium-ion batteries, J. Alloys Compd. 930
multilayered arsenene by the photo-isomerization of azobenzene, Nanoscale 9 (21) (2023) 167380.
(2017) 7006–7011. [83] S. Fan, H. Liu, Ye Huang, et al., Electrolyte engineering on performance
[54] S.M. Oh, S.T. Myung, M.W. Jang, et al., An advanced sodium-ion rechargeable enhancement of NiCo2S4 anode for sodium storage, Small 19 (26) (2023) 2300188.
battery based on a tin–carbon anode and a layered oxide framework cathode, Phys. [84] Y. Li, S. Chen, S. Xu, et al., Impact of electrolyte salts on Na storage performance for
Chem. Chem. Phys. 15 (11) (2013) 3827–3833. high-surface-area carbon anodes, ACS Appl. Mater. Interfaces 13 (41) (2021)
[55] S. Wu, X. Lu, K. Zhang, et al., Nitrogen/phosphorus dual-doped hard carbon anode 48745–48752.
with high initial coulombic efficiency for superior sodium storage, Batteries & [85] C. Wang, K. Wang, M. Ren, et al., Interfacial chemistry enables highly reversible Na
Supercaps 6 (1) (2023) e202200427. extraction/intercalation in layered-oxide cathode materials, Chin. J. Chem. 41 (15)
[56] R. Alcantara, P. Lavela, G.F. Ortiz, et al., Carbon microspheres obtained from (2023) 1791–1796.
resorcinol-formaldehyde as high-capacity electrodes for sodium-ion batteries, [86] J. Wang, H. Lu, J. Zhang, et al., Improved interfacial property of Na3V2(PO4)3@C
Electrochem. Solid State Lett. 8 (4) (2005) A222–A225. cathode: application of NaODFB-based ether electrolyte in sodium-ion batteries,
[57] J.B. Goodenough, Y. Kim, Challenges for rechargeable Li batteries, Chem. Mater. 22 Journal of Electrochemical Energy Conversion and Storage 20 (1) (2023) 011005.
(3) (2010) 587–603. [87] P. Yu, X. Feng, et al., The evolution of local structure of Mo6S8 during Liþ
[58] H. Lai, Y. Lu, W. Zha, et al., In situ generated composite gel polymer electrolyte with electrochemical storage studied by in-situ tender X-ray absorption spectroscopy,
crosslinking structure for dendrite-free and high-performance sodium metal Prog. Nat. Sci.: Mater. Int. 32 (6) (2022) 739–744.
batteries, Energy Storage Mater. 54 (2023) 478–487. [88] V.A. Diamant, S.M. Malovanyy, K.D. Pershina, et al., Electrochemical properties of
[59] J. Guzm an-Torres, A.G. Sanchez-Valdez, L.L. Garza-Tovar, et al., Solid polymer sodium bis [salicylato (2-)]-borate-γ-butyrolactone electrolytes in sodium battery,
electrolyte membranes of trimethylsulfonium bis (trifluoromethylsulfonyl) imide/ Mater. Today: Proc. 6 (2019) 86–94.
NaClO4/PEO for Na-ion batteries, Polym. Bull. 81 (3) (2024) 2465–2480. [89] G. Hernandez, R. Mogensen, R. Younesi, et al., Fluorine-free electrolytes for lithium
[60] Jie Xiong, T.X. Lei, D.M. Fu, et al., Data driven discovery of an analytic formula for and sodium batteries, Batteries & Supercaps 5 (6) (2022) e202100373.
the life prediction of Lithium-ion batteries, Prog. Nat. Sci.: Mater. Int. 32 (6) (2022) [90] A. Bitner-Michalska, A. Krzto _
n-Maziopa, G. Zukowska, et al., Liquid electrolytes
793–799. containing new tailored salts for sodium-ion batteries, Electrochim. Acta 222
[61] G.G. Eshetu, S. Grugeon, H. Kim, et al., Comprehensive insights into the reactivity (2016) 108–115.
of electrolytes based on sodium ions, ChemSusChem 9 (5) (2016) 462–471. [91] A. Plewa-Marczewska, T. Trzeciak, A. Bitner, et al., New tailored sodium salts for
[62] G. Hern andez, R. Mogensen, R. Younesi, et al., Fluorine-free electrolytes for lithium battery applications, Chem. Mater. 26 (17) (2014) 4908–4914.
and sodium batteries, Batteries & Supercaps 5 (6) (2022) e202100373. [92] A. Bitner-Michalska, K. Michalczewski, J. Zdunek, et al., Microwave Plasma
[63] Y. Kim, Y. Kim, A. Choi, et al., Tin phosphide as a promising anode material for Na- Chemical Vapor Deposition of SbxOy/C negative electrodes and their compatibility
ion batteries, Advanced materials 26 (24) (2014) 4139–4144. with lithium and sodium Hückel salts-based, tailored electrolytes, Electrochim. Acta
[64] Y. Pan, Y. Zhang, B.S. Parimalam, et al., Investigation of the solid electrolyte 210 (2016) 395–400.
interphase on hard carbon electrode for sodium ion batteries, J. Electroanal. Chem. [93] Y. Li, F. Wu, Y. Li, et al., Ether-based electrolytes for sodium ion batteries, Chem.
799 (15) (2017) 181–186. Soc. Rev. 51 (11) (2022) 4484–4536.
[65] F. Cheng, M. Cao, Q. Li, et al., Electrolyte salts for sodium-ion batteries: NaPF6 or [94] D. Reber, R. Grissa, M. Becker, et al., Anion selection criteria for water-in-salt
NaClO4? ACS Nano 17 (18) (2023) 18608–18615. electrolytes, Adv. Energy Mater. 11 (5) (2021) 2002913.
[66] A. Ponrouch, E. Marchante, M. Courty, et al., In search of an optimized electrolyte [95] L. Liu, X. Qi, S. Yin, et al., In situ formation of a stable interface in solid-state
for Na-ion batteries, Energy Environ. Sci. 5 (9) (2012) 8572–8583. batteries, ACS Energy Lett. 4 (7) (2019) 1650–1657.
[67] P. Chen, X. Niu, L. Wang, Improving the electrochemical performance of fluorinated [96] J. Sun, L.A. O'Dell, M. Armand, et al., Anion-derived solid-electrolyte interphase
carbons by solvent Co–intercalation for rechargeable Na ion battery cathodes[C]// enables long life Na-ion batteries using superconcentrated ionic liquid electrolytes,
Journal of physics: conference series, IOP Publishing 2499 (1) (2023) 012017. ACS Energy Lett. 6 (7) (2021) 2481–2490.
[68] Y. Li, F. Wu, Y. Li, et al., Ether-based electrolytes for sodium ion batteries, Chem. [97] Q. He, H. Chen, X. Chen, et al., Tea-derived sustainable materials, Adv. Funct.
Soc. Rev. 51 (11) (2022) 4484–4536. Mater. (2024) 2310226.
[69] C. Yu, Y. Li, H. Ren, et al., Engineering homotype heterojunctions in hard carbon to
induce stable solid electrolyte interfaces for sodium-ion batteries, Carbon Energy 5
(1) (2023) e220.

11

You might also like