Next Article in Journal
Interaction of Oxicam Derivatives with the Artificial Models of Biological Membranes—Calorimetric and Fluorescence Spectroscopic Study
Next Article in Special Issue
Development and Investigation of Hierarchically Structured Thin-Film Nanocomposite Membranes from Polyamide/Chitosan Succinate Embedded with a Metal-Organic Framework (Fe-BTC) for Pervaporation
Previous Article in Journal
Sunflower Oilcake as a Potential Source for the Development of Edible Membranes
Previous Article in Special Issue
Chemical and Physical Ionic Liquids in CO2 Capture System Using Membrane Vacuum Regeneration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of the Application of Modified Separators in Inhibiting the “shuttle effect” of Lithium–Sulfur Batteries

1
Hubei Provincial Key Laboratory of Green Materials for Light Industry, Hubei University of Technology, Wuhan 430068, China
2
Shandong Zhongsheng Pharmaceutical Equipment Co., Ltd., Yantai 264010, China
*
Author to whom correspondence should be addressed.
Membranes 2022, 12(8), 790; https://doi.org/10.3390/membranes12080790
Submission received: 23 July 2022 / Revised: 9 August 2022 / Accepted: 10 August 2022 / Published: 17 August 2022
(This article belongs to the Special Issue Mixed-Matrix Membranes and Polymeric Membranes)

Abstract

:
Lithium-sulfur batteries with high theoretical specific capacity and high energy density are considered to be one of the most promising energy storage devices. However, the “shuttle effect” caused by the soluble polysulphide intermediates migrating back and forth between the positive and negative electrodes significantly reduces the active substance content of the battery and hinders the commercial applications of lithium–sulfur batteries. The separator being far from the electrochemical reaction interface and in close contact with the electrode poses an important barrier to polysulfide shuttle. Therefore, the electrochemical performance including coulombic efficiency and cycle stability of lithium–sulfur batteries can be effectively improved by rationally designing the separator. In this paper, the research progress of the modification of lithium–sulfur battery separators is reviewed from the perspectives of adsorption effect, electrostatic effect, and steric hindrance effect, and a novel modification of the lithium–sulfur battery separator is prospected.

1. Introduction

Along with the development of mobile electronics and electric vehicles, a trend to develop energy storage devices with high efficiency and high specific energy has emerged [1,2,3]. Rechargeable batteries with long service life and high energy storage efficiency are attracting much attention from researchers. Among them, lithium–sulfur batteries (LSBs) have high theoretical specific capacity (1675 mAh g−1) and high energy density (2600 Wh kg−1) [4,5], and the cathode sulfur is low cost, abundant, and environmentally friendly. Therefore, LSBs have great development prospects [6]. The charge–discharge process of LSBs is a dissolution–deposition reaction [7]. During the discharge process, the cathode sulfur (S8) is electrochemically reduced to soluble long-chain lithium polysulfide Li2Sx (4 ≤ x ≤ 8) first, and then converted to insoluble short-chain Li2S2/Li2S. The charging reaction is the opposite of the discharging reaction. The solid Li2S2/Li2S is first converted to short-chain LiPS, which is then further oxidized to soluble long-chain Li2Sx (4 ≤ x ≤ 8), and finally to solid S8. The reaction is as follows [8]:
Negative electrode: 16 Li ⇌ 16 Li+ + 16 e
Positive electrode: S8 + 16 Li+ + 16 e ⇌ 8 Li2S
The “shuttle effect” of LSBs is known to be an important factor limiting their practical application [9,10,11,12]. The “shuttle effect” refers to the phenomenon that Li2Sx (4 ≤ x ≤ 8) produced by the positive electrode diffuses to the negative electrode during the charging and discharging process, and is reduced to solid Li2S2/Li2S on the negative electrode surface and attached to the negative electrode. It can cause irreversible loss of battery active material, increase battery internal resistance, and decrease the theoretical capacity of LSBs [13,14,15]. At present, researchers have proposed many methods to suppress the “shuttle effect” of LSBs, such as electrode modification, electrolyte optimization, and separator modification. Among them, the separator is far away from the electrochemical reaction interface and in close contact with the electrode, and more importantly, it is also the diffusion channel of polysulfides (LiPSs). Therefore, its structure and performance greatly affect the electrode reaction of LSBs and the shuttle of LiPSs between the two electrodes. However, currently, the most commercialized and widely used polyolefin separators (such as polyethylene (PE) and polypropylene (PP)) have large pore sizes and do not interact with LiPSs, so they cannot inhibit the shuttle of LiPSs between the two electrodes of LSBs [16,17]. Therefore, under the premise of ensuring the basic functions of the polyolefin-based separator, if it is modified to have the function of adsorbing and/or converting LiPSs, it will be able to inhibit the “shuttle effect” of LiPSs. [18,19,20]. From what is known in the literature [21,22,23,24], the current separator modifications of LSBs are mainly based on confining LiPSs to the cathode region of the battery during their migration to the anode. This can be done from three perspectives: (1) modification by adsorption effect, making the separator adsorbable and fixing LiPSs on the positive side; (2) modification by electrostatic effect, making the separator repulsive and inhibiting the migration of LiPSs; (3) modification by positioning barrier effect, reducing the pore size of the separator and keeping the LiPSs in the positive electrode region by means of physical barrier. These three methods are described below.

2. Adsorption Effect Modification

2.1. Physical/Chemical Adsorption

The modification of the separator of LSBs by means of the physical adsorption effect generally refers to the use of the van der Waals force between LiPSs and the trapping material to capture LiPSs. Separator materials with high porosity and large specific surface area can provide more trapping sites, which are beneficial to inhibit the migration of LiPSs between the electrodes [25]. Therefore, attaching porous carbon materials such as carbon nanofibers, carbon flakes, microporous carbon nanofibers, reduced graphene oxide, etc. as a coating to the separator has been widely used in the separator modification of LSBs. The modification effects of different carbon materials are shown in Table 1. Generally speaking, the high sulfur loading of the cathode will increase more LiPSs in the electrolyte and result in a severer shuttle effect of LSBs. From Table 1, it can be seen that the improvement of using light mesoporous carbon as the coating to modify the separator is the most significant. At the high sulfur loading (3.5 mg cm−2), the LSBs with this separator have a mass specific capacity of 1021 mAh g−1 after 100 cycles at 0.5 C, with a capacity decay rate of only 0.081%. This may be because the large number of mesopores (12 nm) in the mesoporous carbon can tune the huge volume change of S8 during the lithiation process, thereby suppressing the loss of polysulfides, which in turn controls the shuttle of LiPSs and improves the electrochemical performance of LSBs.
Besides carbon materials, non-carbon materials can also be used as coatings to enhance the adsorption capacity of the separators of LiPSs. Deng et al. [16] prepared a polymetaphenylene isophthalamide (PMIN) separator by electrospinning. The addition of tetrabutylammonium chloride to the spinning solution produces a separator with high porosity. Using the dendritic nanofiber separator with good physical trapping effect, the LSBs exhibited a capacity decay rate of only 0.049% after 800 cycles at 0.5 C.
Chiu et al. [36] modified the PP separator of LSBs with a mixture of ethylene oxide (PEO) and lithium bis(trifluoromethanesulfonyl)imide (LiTFSI) as a coating. The experimental results show that the lithium-ion diffusion coefficient (9.6 × 10−9~3.0 × 10−8 cm2 s−1) of PEO/LiTFSI coating is significantly higher than that of pure PP separator (7.6 × 10−9~2.2 × 10−8 cm2 s−1). The LSBs with PEO/LiTFSI separator achieve an initial discharge capacity of 1212 mAh g−1 at 0.1 C and can still maintain a high reversible capacity of 534 mAh g−1 and a stable coulombic capacity after 200 cycles. This may be due to the fact that PEO can inhibit the diffusion of LiPSs as a gel polymer electrolyte, and the addition of LiTFSI salt enhances the ability of the lithium ion transfer of the PEO coating. The electrochemical efficiency and the synergistic effect of the PEO/LiTFSI coating are shown in Figure 1.
Although the modification of polyolefin separators with carbon materials can inhibit the shuttle of LiPSs between the two electrodes of LSBs to a certain extent, the interaction force between non-polar carbon materials and polar LiPSs is weak. Therefore, inhibiting the diffusion of LiPSs in the LSBs only by physical adsorption is not satisfied [37]. Compared with physical adsorption, chemical adsorption, which relies on the formation of chemical bonding forces between LiPSs and the surface atoms of adsorbent materials, can make a higher selectivity and achieve a better immobilizing effect of LiPSs [38]. For example, doping electronegative heteroatoms (such as N, O, and S, etc.) in carbon coatings can help trap LiPSs by the formation of Li–X bonds (dipole–dipole interactions) between heteroatoms and LiPSs [39,40]. Hou et al. [41] analyzed the ability of various atomically doped graphene nanoribbons (GNRs) to adsorb sulfur and Li2Sx by density functional theory (DFT). It was found that the adsorption effect of N and O (binding energy −2.53~−2.56 eV)-doped GNRs on Li2S4 is much greater than that of B, F, S, P, and Cl (binding energy −1.93 eV), which is due to the existence of dipole–dipole electrostatic interactions (~1.95 eV). The relationship between element electronegativity and binding energy is shown in Figure 2.
Zeng et al. [42] designed a honeycomb-like N, P diatomic doped carbon (HNPC) modified separator, in which the HNPC coating can effectively anchor the LiPSs by forming N–Li and P–S bonds. When using the acetylene black-sulfur composite as the cathode with the sulfur content of 79.7%, the LSBs showed excellent long-cycle stability with a capacity decay rate of only 0.06% per cycle after 900 cycles at 1 C. Figure 3 shows the cycle performance of separators modified by different coatings. It can be seen that the chemical confinement of LiPSs by diatomic doping is higher than that of single-atom doping systems.
Besides polar atoms, the doping of polar functional groups can also realize the chemisorption of the separator to LiPSs. Pei et al. [43] used polydopamine/polyethylene oxide (PDA/PEI) to modify polyolefin separators and achieved desired results. The PDA/PEI-modified separator made the initial discharge capacity of LSBs 1250 mAh g−1 at 0.2 C and it remained900 mAh g−1 after 100 cycles. This could be attributed to the formation of numerous C=N bonds between the amine groups in PEI and dopamine, which inhibited the accumulation of dopamine oligomers on the separator through hydrogen bonds or π−π bonds, thereby making the separator hydrophilic, and the enriched N and O functional groups of the PDA/PEI coating effectively adsorbed LiPSs in the electrolyte.
In addition, the modification of PP separator with metal materials, metal oxides, and metal sulfides, the exposed metal (M) sites of which can form S–M bonds with LiPSs, are also beneficial to improve its adsorption capacity to LiPSs [44]. Tao et al. [12] calculated by DFT that the binding energy of metal oxides and LiPSs (from −1.54 to −7.12 eV), is higher than that of heteroatoms and LiPSs (from −2.53 to −2.56 eV). Liu et al. [45] designed graphene oxide (GO) coatings doped with TiO2 nanoparticles, as shown in Figure 4a. The Ti–O–C bond formed between TiO2 and GO and the wrinkled sheet structure of GO make the combination of TiO2 and GO tight, which helps to enhance the electrical conductivity of the separator. Meanwhile, the S–Ti–O and Ti–S bonds formed between TiO2 and S enhanced the ability of the separator to adsorb LiPSs. The LSBs coated with TiO2/GO separator can retain a reversible capacity of 843.4 mAh g−1 after 100 cycles at 0.2 C (Figure 4b).

2.2. Catalytic Adsorption

Physical/chemical adsorption can suppress the “shuttle effect” of LiPSs, but under the conditions of high sulfur loading (>5 mg cm−2) or long cycling, the increase of LiPSs dissolved in the electrolyte will lead to slow reaction kinetics. Moreover, the limited adsorption sites on the separator are not enough to adsorb LiPSs in the electrolyte fully, thus resulting in a low specific capacity and short cycle life of the battery. Therefore, the ability of the separator to suppress the “shuttle effect” may be reduced or even lost [46,47]. The introduction of catalytic substances on the surface of the separator can not only reduce the energy barrier of the conversion of LiPSs and insoluble Li2S2/Li2S, accelerating the electrochemical reaction of LSBs, but also transfer the adsorbed LiPSs to the redox reaction in time. Thus, the content of the active substances of the battery is kept to the greatest extent, and the cycle performance of LSB is significantly improved [14].
Inherent defects in the metal oxide structure can provide more active sites for trapping and transforming LiPSs [48]. Lv et al. [49] loaded bimetallic NiCo2O4 nanoparticles onto reduced graphene oxide as a catalytic coating for the separator. During the electrode reaction process, the oxidized metal ions (Ni3+ and Co3+) combine with LiPSs to form Ni–S and Co–S bonds, which anchor the LiPSs to the NiCo2O4@rGO surface to suppress the “shuttle effect” of LBSs. Furthermore, DFT calculations show that the NiCo2O4 (100) crystal surface has a low Li-ion diffusion barrier (0.15 eV compared to 0.293 eV for carbon materials), which helps to improve the migration rate of lithium ions and promote lithium-ion–electron coupling, thereby accelerating the conversion of LiPSs to Li2S2/Li2S. Using a core-shell structure material (encapsulated sulfur in carbon) with a sulfur content of 70 wt% as the cathode and under the condition of a high sulfur loading of 6 mg cm−2, the LBSs showed the capacity loss rate only 0.02% after 400 cycles at a current density of 1 mA cm−2, and at the same time, there was an excellent areal capacity of 7.1 mAh cm−2.
Adding other metals to monometallic compounds can also improve the reaction kinetics of LiPS transformation on the surface of metal compounds [50]. Zhang et al. [51] introduced Mo atoms into Ni3N and used a composite of multi-walled carbon nanotubes and Ni0.2Mo0.8N to modify the separator. During the electrochemical reaction process, the Mo atoms in the Ni0.2Mo0.8N are etched and leached out by LiPSs, leaving a large number of vacancies around Ni (Figure 5a). The vacancies formed can accelerate the charge transfer and the LiPS conversion. The potential of the Ni3N and Mo2N surfaces in the coating is higher than lithium negative electrode, thus creating an electric field between the separator and the lithium anode (Figure 5b), which promotes the directional movement of sulfur and lithium ions and inhibits the shuttling of LiPSs to some extent. The LSBs with this separator can achieve an initial battery capacity of 1097.2 mAh g−1 at a high rate of 5 C.
Due to the abundant pore structures and catalytic sites, metal–organic frameworks (MOFs) composed of organic ligands and transition metal ions have been widely used for catalytic modification of the separators of LBSs. The pore structure in MOFs can promote the adequate contact between the separator and the electrolyte and limits the diffusion of LiPSs, while the metal central ions can chemisorb and catalyze LiPSs [41]. Hong et al. [52] doped cerium-based metal-organic frameworks (Ce-MOFs) into CNTs to form Ce-MOFs/CNT separator coatings. The large specific surface area of Ce-MOFs and the ligated unsaturated Ce (IV) cluster nodes can rapidly adsorb LiPSs and accelerate the conversion between LiPSs and Li2S2/Li2S, which is shown in Figure 6a. As seen, under the high sulfur loading of 6 mg cm−2, the initial specific capacity of LSBs was 1021.8 mAh g−1 at 1 C, which slowly decreased to 838.8 mAh g−1 after 800 cycles with a decay rate of only 0.022% per cycle, as shown in Figure 6b. The coulombic efficiency of the LBSs was close to 100%, which suggesting its excellent cycle performance.
In addition to transition metal elements, rare-earth elements have also been used in the catalytic modification of the separators of LBSs. Peng et al. [53] prepared a novel pp separator coated with Eu2O3/Ketjen black (Eu2O3/KB). The exposed (222) plane of Eu2O3 exhibits strong polarity-polarity interaction with LiPSs, which is beneficial to confine the transfer of LiPSs. The adsorption energy of Eu2O3 with LiPSs is shown in Figure 7. Apart from that, oxygen vacancies in Eu2O3 can provide more catalytic sites for regulating Li2S precipitation. Therefore, with high crystal face index, Eu2O3 can lower the energy barrier for Li2S nucleation on it and promote the conversion of LiPSs to Li2S. The LBSs with Eu2O3/KB modified separator exhibited excellent cycling stability and rate capability. Its capacity decay rate achieved only 0.05% per cycle during 500 cycles at 1 C.
Some non-metallic materials with catalytic activity have also been used in the separator of LBSs to promote the conversion of LiPSs. For example, Wang et al. [54] used red phosphorus (RP) as a coating to enhance the sulfur reaction kinetics of LSBs. Through Lewis acid–base interaction, The Li3PO4 with high conductivity was formed by RP and LiPSs, which can improve the conductivity of lithium ions and promote the redox reaction of sulfur. However, the formation of Li3PO4 consumes some active sites for adsorbing LiPSs in the separator, which may lead to the decrease of the capacity of the LSBs under long-term cycling. The LSBs with RP-modified separator exhibited a capacity of 729.6 mAh g−1 after 500 cycles at 1 C with a capacity retention rate of 82%.

3. Electrostatic Effect

The sulfur atoms at the end of LiPSs possess a large number of negative charges and free radicals. The introduction of negatively charged species on the separator can inhibit the shuttle of polysulfide ions through electrostatic repulsion and simultaneously facilitate Li+ migration. This could decrease the resistance between the positive electrode and the separator, preventing the energy generated by the battery from dissipating in the form of thermal energy [39,55].

3.1. Negatively Charged Group

Song et al. [56] grafted polyacrylic acid (PAA) on the surface of PP separator (PP-g-PAA). A large number of carboxyl functional groups in PAA can not only repel LiPSs anions through electrostatic interaction, but also enhance the hydrophilicity of the separator and reduce the internal resistance of LSBs. Chiu et al. [57] prepared a gel polymer electrolyte separator coating by mixing polymethyl methacrylate with rich -COO- and LiTFSI. When the sulfur loading of the cathode was increased from 4 to 10 mg cm−2, the lithium-ion diffusion coefficient of LSBs slightly decreased from 5.27 × 10−9–3.1 × 10−8 cm2 s−1 to 2.09 × 10−9–1.27 × 10−8 cm2 s−1, which indicates that the coating enhances the lithium-ion transfer efficiency and suppresses the diffusion of LiPSs. Under the high sulfur loading of 8 mg cm−2, LSBs with this separator exhibit a high area capacity of 7.1 mWh cm−2. Li et al. [58] modified PP with Fe3O4 nanoparticles and carboxylated carbon nanotubes (CTs-COOH). The Fe3O4 layer acts as an “upper current collector” and facilitates electron transport, while -COOH can inhibit the transport of LiPSs through electrostatic effects. The synergistic effect of the two factors enables LSBs to have a capacity decay rate as low as 0.032% per cycle when cycled for 2000 times at 1 C.
Nafion with rich sulfonic acid groups (-SO3H) is also widely used in the modification of the separators of LSBs. But due to the poor conductivity, an overly thick coating of Nafion will reduce the lithium-ion transport capacity of the separator [59]. Zhuang et al. [60] designed a ternary-layer separator composed of PP, graphene oxide (GO) and Nafion for LSBs (Figure 8), in which the ultrathin GO layer (0.0032 mg cm−2, about 40 layers) can cover the macropores of the PP matrix and provides support for the Nafion layer. The Nafion layer with the areal loading of 0.05 mg cm−2 also retains the conductivity of the separator while repelling LiPSs. The specific capacity of LSBs using the ternary layer separator can reach a discharge capacity of 1057 mAh g−1 at 0.1 C.
In addition, the confinement of LiPSs can also be achieved by using the charge environment between the two-dimensional material layers. Xu et al. [61] used two-dimensional vermiculite nanosheets to prepare the separators of LSBs, the structure of which is shown in Figure 9. The vermiculite nanosheet is negatively charged, so the positively charged lithium ions can be transferred between the layers, while the negatively charged polysulfide ions cannot diffuse due to electrostatic interaction. This effectively suppressed the transmembrane diffusion of LiPSs.

3.2. Polar Particles

Besides negatively charged groups, some polar particles can also repel LiPSs through electrostatic effects. BaTiO3 (BTO) with high dielectric constant can generate a large number of positive and negative charges on the surface of BTO through spontaneous polarization, which can electrostatically interact with lithium ions and polar molecules LiPSs [62]. Yim et al. [63] embedded BTO particles into PE separators. The polar BTO forms permanent dipoles under the action of electric field (Figure 10), which can block LiPSs through electrostatic repulsion, thereby improving the cycling performance of LSBs. The polarized BTO coating can also enhance the mechanical stability of the PE separator and improve the capacity and cycle life of the LSBs. At a current density of 0.5 C, the LSBs with this separator applied achieves an initial discharge capacity of 1122.1 mAh g−1.

4. Steric Hindrance Effect

The pore size of commonly used polyolefin separators (about 100 nm) is much larger than the size of LiPSs, which cannot block the shuttle of LiPSs between the cathode and anode of LSBs. Thus, reducing the pore size of the diaphragm can hinder the diffusion of LiPSs to the anode [64]. The modification methods of LSBs separators based on this steric hindrance effect can be broadly classified into two categories: (1) accumulation of particles on the surface of the separator to form a densely modified layer covering the original macropores; (2) construction of a new microporous separator using microporous materials.

4.1. Particle Buildup

Maletti et al. [65] prepared a lithium vanadium oxide (LiV3O8)-coated separator. The dense coating formed by LiV3O8 nanoparticles can cover the original macropore structure of the PP separator, forming a physical barrier to inhibit the diffusion of LiPSs. LiV3O8 can disproportionate with LiPSs to promote the conversion of LiPSs to Li2S2/Li2S (Figure 11). In addition, the LiV3O8 coating significantly reduces the resistance of the cell by facilitating charge transfer at the cathode/diaphragm coating interface. Compared with pure PP, the charge transfer resistance of LSBs with LiV3O8 separator was significantly reduced by 83%, with a high initial discharge capacity of 1254 mAh g−1 at a current density of 0.2 C and a capacity decay rate of 0.063% after 500 cycles at 0.5 C.
Besides the pore size reduction that is possible by loading zero-dimensional nanoparticles on the surface of the separator, the layered structure formed by the stacking of two-dimensional nanosheets can also form a physical barrier to intercept LiPSs. Ghazi et al. [66] deposited multilayer MoS2 nanosheets with a thickness of about 350 nm on the surface of polyolefin separators. The formed dense structure can effectively block LiPSs, and at the same time, the voids between the MoS2 nanosheets can prevent solid LiS from forming an insulating layer on the separator and provide a transfer channel for lithium ions, which realizes the effective sieving of LiPSs and lithium ions.
Chang et al. [67] designed a two-dimensional MOF sheet molecular sieve (CuBDCs) coating with one-dimensional vertical channels. The structure is shown in Figure 12. The one-dimensional channels (about 5.2 Å) in the CuBDCs nanosheets are much smaller than the size of polysulfides, which can act as molecular sieves and hinder the diffusion of LiPSs to the anode. Secondly, the CuBDC sheet is approximately 10 nm thick, which combined with its internal one-dimensional channels shortens the ion transport path, improving the mobility of Li ions. The LSBs using this separator can still maintain a specific capacity of 830 mAh g−1 after 1000 long cycles at 1 C. This indicates that the 2D CuBDC sheets have excellent sieving ability for LiPSs.

4.2. New Microporous Separator

A commonly used way to reduce the pore size of polyolefin separators is to coat their surfaces with small-pore-size coatings, and carbon coatings with nanopores have been shown to have good performance in inhibiting the “shuttle effect” of LiPSs under high sulfur-loading conditions [68]. However, the surface energy of the polyolefin separator is low, and it is difficult to adhere a uniform coating material on the surface of the separator. Therefore, the phenomenon of the coating peeling off the separator during the cycle occurs occasionally. Using novel functional separators as substrates is one of the fundamental solutions to this problem.
Compared with the traditional polyolefin separator, the electrospun nanofiber separator has higher porosity and chemical stability. The strong affinity of electrospun nanofibres for the electrolyte can significantly improve the mobility of lithium ions in LSBs, thereby effectively inhibiting the “shuttle effect” of LSBs [69]. Pei et al. [70] prepared a structurally asymmetric aramid nanofiber Janus separator (ANF-JS). The separator structure is shown in Figure 13. The side of the separator facing the positive electrode is a nanoporous layer with ion selectivity, which can hinder the shuttling of LiPSs between electrodes; the side of the separator facing the negative electrode is a microporous layer with a three-dimensional porous structure, which can provide channel for electrolyte absorption and fast lithium-ion transport. After 1000 cycles at 1 C with this separator, the capacity decay rate of the LSBs is reduced to 0.0195% per cycle.
Doping functional particles in the novel separator can combine the steric hindrance effect with other effects, enabling the separator to further inhibit the shuttle of LiPSs [71]. Guo et al. [72] prepared a flexible Ti4O7/C nanofiber barrier layer (TCNFs) by electrospinning. Figure 14a shows the adsorption–desorption and BET curves, and it can be seen that the pore size of the separator was between 2 and 16 nm and the specific surface area was 40 m2 g−1, which can effectively isolate LiPSs form the positive side of LSBs. The TCNFs separator has dual effects of physical shielding and chemical adsorption on LiPSs, which make the capacity fade of the LSBs only 0.03% per cycle after 2500 cycles, even at a current density of 3 C, as shown in Figure 14b.
Wang et al. [73] prepared separators (ZIF-7@PCF) with cross-linked structures by grewing two-dimensional MOF sheets (ZIF-7) on porous carbon nanofibers (PCFs), as shown in Figure 15a. The pores of the separator are mostly mesopores and micropores with diameters of less than 10 nm, which are beneficial to hinder the transport of LiPSs. PCF porous carbonaceous framework can facilitate electron and ion transport in batteries, while ZIF-7 MOF provides an abundant active site for the capture and conversion of LiPS. Under high sulfur-loading conditions of 6 mg cm−2, the LSB with ZIF-7@PCF exhibits an initial discharge specific capacity of 1221 mAh g−1 at a current density of 0.2 C and a discharge specific capacity of 661.3 mAh g−1 at 5 C, showing good rate performance, as shown in Figure 15b.

5. Conclusions

This paper mainly introduces the methods of modifying the lithium–sulfur battery separators by the adsorption effect, electrostatic effect, and steric hindrance effect using polyolefin separator as the matrix. Compared with the traditional polyolefin separator, the modified separators have good function for inhibiting the “shuttle effect” of LiPSs. Especially, the modified separator with catalytic effect can maximize the utilization of the active species in LSBs, showing excellent cycling and rate performance under long-cycle conditions and high sulfur loading. The current separator modification materials for high sulfur loading and low electrolyte LSBs are mainly carbon-based composites with hollow or layered structures, and the robust internal porous network structure can provide buffers for sulfur volume changes and the shuttle of LiPSs. However, overly thick carbon coating will reduce the actual energy density of the battery, so lightweight materials such as graphene can be selected as the base layer to reduce the thickness of the coating and the load quality. To realize the high energy density and rate capability of LSBs, it is possible to consider using nanoscale catalytic materials with large specific surface area and short ion transport distance or gel electrolyte materials with high ionic conductivity to modify the separator. In addition, the functional modification of novel nanofiber separators with high chemical stability and electrolyte affinity also opens up new avenues for developing high-energy-density LSB separators.

Author Contributions

Conceptualization, C.C., Y.P. and B.-M.X.; writing—original draft preparation, B.-W.Z.; writing—review and editing, B.-W.Z.; article revision, B.S.; literature review, P.F.; illustrations and tables, F.L. and C.Z.; visualization, C.C.; supervision, Y.P. and B.-M.X.; funding acquisition, C.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Doctoral Scientific Research Foundation of Hubei University of Technology (Grant No. BSQD2019030).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Acknowledgments

The authors are thankful to School of Materials and Chemical Engineering, Hubei University of Technology, China.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Islam, M.S.; Nazar, L.F. Advanced materials for lithium batteries. J. Mater. Chem. 2011, 21, 9810. [Google Scholar] [CrossRef]
  2. Ma, Y.; Xie, X.B.; Yang, W.Y.; Yu, Z.P.; Sun, X.Q.; Zhang, Y.P.; Yang, X.Y.; Kimura, H.; Hou, C.X.; Guo, Z.H.; et al. Recent advances in transition metal oxides with different dimensions as electrodes for high-performance supercapacitors. Adv. Compos. Hybrid Mater. 2021, 4, 906–924. [Google Scholar] [CrossRef]
  3. Zhai, I.Y.; Yang, W.; Xie, X.; Yu, Z.P.; Sun, X.Q.; Zhang, Y.P.; Yang, X.Y.; Kimura, H.; Hou, C.X.; Guo, Z.H.; et al. Co3O4 nanoparticle-dotted hierarchical-assembled carbon nanosheet framework catalysts with the formation/decomposition mechanisms of Li2O2 for smart lithium–oxygen batteries. Inorg. Chem. Front. 2022, 9, 1115–1124. [Google Scholar] [CrossRef]
  4. Ould, E.T.; Kamzabek, D.; Chakraborty, D.; Doherty, M.F. Lithium–sulfur batteries: State of the art and future directions. ACS Appl. Energy Mater. 2018, 1, 1783–1814. [Google Scholar] [CrossRef]
  5. Chen, H.; Wang, C.; Dong, W.; Lu, W.; Du, Z.L.; Chen, L.W. Monodispersed sulfur nanoparticles for lithium–sulfur batteries with theoretical performance. Nano Lett. 2015, 15, 798–802. [Google Scholar] [CrossRef] [PubMed]
  6. Zhao, M.; Li, B.Q.; Zhang, X.Q.; Huang, J.Q.; Zhang, Q. A perspective toward practical lithium–sulfur batteries. ACS Cent. Sci. 2020, 6, 1095–1104. [Google Scholar] [CrossRef] [PubMed]
  7. Li, G.; Wang, S.; Zhang, Y.; Li, M.; Chen, Z.W.; Lu, J. Revisiting the role of polysulfides in lithium–sulfur batteries. Adv. Mater. 2018, 30, 1705590. [Google Scholar] [CrossRef] [PubMed]
  8. Abbas, S.A.; Ding, J.; Wu, S.H.; Fang, J.; Boopathi, K.M.; Mohapatra, A.; Lee, L.W.; Wang, P.C.; Chang, C.C.; Chu, C.W. Modified separator performing dual physical/chemical roles to inhibit polysulfide shuttle resulting in ultrastable Li–S batteries. ACS Nano 2017, 11, 12436–12445. [Google Scholar] [CrossRef]
  9. He, Y.; Qiao, Y.; Chang, Z.; Cao, X.; Jia, M.; He, P.; Zhou, H. Developing A “Polysulfide-Phobic” Strategy to Restrain Shuttle Effect in Lithium-Sulfur Batteries. Angew. Chem. 2019, 58, 11774–11778. [Google Scholar] [CrossRef]
  10. Yan, J.; Liu, X.; Li, B. Capacity Fade Analysis of Sulfur Cathodes in Lithium-Sulfur Batteries. Adv. Sci. 2016, 3, 1600101–1600111. [Google Scholar] [CrossRef]
  11. Li, T.; Bai, X.; Gulzar, U.; Bai, Y.J.; Capiglia, C.; Deng, W.; Zhou, X.F.; Liu, Z.P.; Feng, Z.F.; Zaccaria, R.P. A Comprehensive Understanding of Lithium–Sulfur Battery Technology. Adv. Funct. Mater. 2019, 29, 1901730. [Google Scholar] [CrossRef]
  12. Tao, X.; Wang, J.; Liu, C.; Wang, H.; Yao, H.; Zheng, G.Y.; Seh, Z.W.; Cai, Q.X.; Li, W.Y.; Zhou, G.; et al. Balancing surface adsorption and diffusion of lithium-polysulfides on nonconductive oxides for lithium-sulfur battery design. Nat. Commun. 2016, 7, 11203. [Google Scholar] [CrossRef]
  13. Manthiram, A.; Fu, Y.; Su, Y.S. Challenges and prospects of lithium–sulfur batteries. Acc. Chem. Res. 2013, 46, 1125–1134. [Google Scholar] [CrossRef]
  14. Manthiram, A.; Chung, S.H.; Zu, C. Lithium–sulfur batteries: Progress and prospects. Adv. Mater. 2015, 27, 1980–2006. [Google Scholar] [CrossRef]
  15. Dent, M.; Jakubczyk, E.; Zhang, T.; Lekakou, C. Kinetics of sulphur dissolution in lithium–sulphur batteries. J. Phys. Energy 2022, 4, 024001. [Google Scholar] [CrossRef]
  16. Deng, N.; Wang, Y.; Yan, J.; Ju, J.; Li, Z.J.; Fan, L.L.; Zhao, H.J.; Kang, W.M.; Cheng, B.W. A F-doped tree-like nanofiber structural poly-m-phenyleneisophthalamide separator for high-performance lithium-sulfur batteries. Power Sources 2017, 362, 243–249. [Google Scholar] [CrossRef]
  17. Dong, Q.; Shen, R.; Li, C.; Gan, R.Y.; Ma, X.T.; Wang, J.C.; Li, J.; Wei, Z.D. Construction of Soft Base Tongs on Separator to Grasp Polysulfides from Shuttling in Lithium-Sulfur Batteries. Small 2018, 14, 1804277. [Google Scholar] [CrossRef]
  18. Liu, Y.T.; Liu, S.; Li, G.R.; Guo, X.P. Strategy of enhancing the volumetric energy density for lithium–sulfur batteries. Adv. Mater. 2021, 33, 2003955. [Google Scholar] [CrossRef]
  19. Wang, P.; Xi, B.; Huang, M.; Chen, W.H.; Feng, J.K.; Xiong, S.L. Emerging catalysts to promote kinetics of lithium–sulfur batteries. Adv. Energy Mater. 2021, 11, 2002893. [Google Scholar] [CrossRef]
  20. Fang, R.; Zhao, S.; Sun, Z.; Wang, D.W.; Chen, H.M.; Li, F. More reliable lithium-sulfur batteries: Status, solutions and prospects. Adv. Mater. 2017, 29, 1606823. [Google Scholar] [CrossRef]
  21. Chung, S.H.; Chang, C.H.; Manthiram, A. Hierarchical sulfur electrodes as a testing platform for understanding the high-loading capability of Li-S batteries. J. Power Sources 2016, 334, 179–190. [Google Scholar] [CrossRef]
  22. Dang, C.; Mu, Q.; Xie, X.; Sun, X.; Yang, X.; Zhang, Y.P.; Maganti, S.; Huang, M.; Jiang, Q.L.; Seok, I.; et al. Recent progress in cathode catalyst for nonaqueous lithium oxygen batteries: A review. Adv. Compos. Hybrid Mater. 2022, 5, 606–626. [Google Scholar] [CrossRef]
  23. Zhang, K.; Chen, Z.X.; Ning, R.Q.; Xi, S.B.; Du, Y.H.; Liu, C.B.; Ren, Z.Y.; Chi, X.; Bai, M.H.; Shen, C.; et al. Single-Atom Coated Separator for Robust Lithium-Sulfur Batteries. ACS Appl. Mater. Interfaces 2019, 11, 25147–25154. [Google Scholar] [CrossRef]
  24. Song, C.W.; Peng, C.X.; Bian, Z.H.; Dong, F.; Xu, H.Y.; Zheng, S.Y. Stable and Fast Lithium–Sulfur Battery Achieved by Rational Design of Multifunctional Separator. Energy Environ. Mater. 2019, 2, 216–224. [Google Scholar] [CrossRef]
  25. Liao, H.Y.; Zhang, H.Y.; Hong, H.Q.; Li, Z.H.; Lin, Y.X. Novel flower-like hierarchical carbon sphere with multi-scale pores coated on PP separator for high-performance lithium-sulfur batteries. Electrochim. Acta 2017, 257, 210–216. [Google Scholar] [CrossRef]
  26. Chung, S.H.; Manthiram, A. Bifunctional Separator with a Light-Weight Carbon-Coating for Dynamically and Statically Stable Lithium-Sulfur Batteries. Adv. Funct. Mater. 2014, 24, 5299–5306. [Google Scholar] [CrossRef]
  27. Chung, S.H.; Manthiram, A. High-Performance Li-S Batteries with an Ultra-lightweight MWCNT-Coated Separator. J. Phys. Chem. Lett. 2014, 5, 1978–1983. [Google Scholar] [CrossRef]
  28. Zhai, P.Y.; Peng, H.J.; Cheng, X.B.; Zhu, L.; Huang, J.Q.; Zhu, W.C.; Zhang, Q. Scaled-up fabrication of porous-graphene-modified separators for high-capacity lithium–sulfur batteries. Energy Storage Mater. 2017, 7, 56–63. [Google Scholar] [CrossRef]
  29. Tan, L.; Li, X.H.; Wang, Z.X.; Guo, H.J.; Wang, J.X.; An, L. Multifunctional Separator with Porous Carbon/Multi-Walled Carbon Nanotube Coating for Advanced Lithium−Sulfur Batteries. Chemelectrochem 2018, 5, 71–77. [Google Scholar] [CrossRef]
  30. Lee, D.K.; Ahn, C.W.; Jeon, H.J. Web-structured graphitic carbon fiber felt as an interlayer for rechargeable lithium—sulfur batteries with highly improved cycling performance. J. Power Sources 2017, 360, 559–568. [Google Scholar] [CrossRef]
  31. Zhang, B.; Yu, L.; Zhao, Y. Ultralight carbon flakes modified separator as an effective polysulfide barrier for lithium-sulfur batteries. Electrochim. Acta 2019, 295, 910–917. [Google Scholar] [CrossRef]
  32. Zhang, K.M.; Dai, L.Q.; Xie, L.J.; Kong, Q.Q.; Su, F.Y.; Shi, J.; Liu, Y.Z. Graphene/Carbon Black Co-modified Separator as Polysulfides Trapper for Li-S Batteries. Chemistryselect 2019, 4, 6026–6034. [Google Scholar] [CrossRef]
  33. Guo, Y.F.; Xiao, J.R.; Hou, Y.X.; Wang, Z.Y.; Jiang, A.H. Carbon nanotube doped active carbon coated separator for enhanced electrochemical performance of lithium–sulfur batteries. J. Mater. Sci. Mater. Electron. 2017, 28, 17453–17460. [Google Scholar] [CrossRef]
  34. Zhang, Z.A.; Wang, G.C.; Lai, Y.Q.; Li, J. A freestanding hollow carbon nanofiber/reduced graphene oxide interlayer for high-performance lithium–sulfur batteries. J. Alloys Compd. 2016, 663, 501–506. [Google Scholar] [CrossRef]
  35. Balach, J.; Jaumann, T.; Klose, M.; Oswald, S.; Eckert, J.; Giebeler, L. Functional Mesoporous Carbon-Coated Separator for Long-Life, High-Energy Lithium-Sulfur Batteries. Adv. Funct. Mater. 2015, 25, 5285–5291. [Google Scholar] [CrossRef]
  36. Chiu, L.L.; Chung, S.H. A Poly(ethylene oxide)/Lithium bis(trifluoromethanesulfonyl)imide-Coated Polypropylene Membrane for a High-Loading Lithium-Sulfur Battery. Polymers 2021, 13, 535. [Google Scholar] [CrossRef]
  37. Fan, Y.; Niu, Z.; Zhang, F.; Zhang, R.; Zhao, Y.; Lu, G. Suppressing the Shuttle Effect in Lithium-Sulfur Batteries by a UiO-66-Modified Polypropylene Separator. ACS Omega 2019, 4, 10328–10335. [Google Scholar] [CrossRef]
  38. Li, N.; Xie, Y.; Peng, S.T.; Xiong, X.; Han, K. Ultra-lightweight Ti3C2T MXene modified separator for Li–S batteries: Thickness regulation enabled polysulfide inhibition and lithium ion transportation. J. Energy Chem. 2020, 42, 116–125. [Google Scholar] [CrossRef]
  39. Zuo, X.; Zhen, M.; Wang, C. Ni@N-doped graphene nanosheets and CNTs hybrids modified separator as efficient polysulfide barrier for high-performance lithium sulfur batteries. Nano Res. 2019, 12, 829–836. [Google Scholar] [CrossRef]
  40. Li, P.Y.; Lv, H.W.; Li, Z.L.; Meng, H.P.; Lin, Z.; Wang, R.H.; Li, X.J. The Electrostatic Attraction and Catalytic Effect Enabled by Ionic-Covalent Organic Nanosheets on MXene for Separator Modification of Lithium-Sulfur Batteries. Adv. Mater. 2021, 33, 2007803. [Google Scholar] [CrossRef]
  41. Hou, T.Z.; Chen, X.; Peng, H.J.; Huang, J.Q.; Li, B.Q.; Zhang, Q.; Li, B. Design Principles for Heteroatom-Doped Nanocarbon to Achieve Strong Anchoring of Polysulfides for Lithium-Sulfur Batteries. Small 2016, 12, 3283–3291. [Google Scholar] [CrossRef]
  42. Zeng, P.; Huang, L.W.; Zhang, X.L.; Zhang, R.X.; Wu, L.; Chen, Y.G. Long-life and high-areal-capacity lithium-sulfur batteries realized by a honeycomb-like N, P dual-doped carbon modified separator. Chem. Eng. J. 2018, 349, 327–337. [Google Scholar] [CrossRef]
  43. Pei, C.B.; Li, J.D.; Lv, Z.Z.; Wang, H.M.; Dong, W.; Yao, Y.Y. Inhibiting polysulfides with PDA/PEI-functionalized separators for stable lithium-sulfur batteries. Int. J. Energy Res. 2021, 46, 10099–10110. [Google Scholar] [CrossRef]
  44. Zhang, H.; Lin, C.; Hu, X.H.; Zhu, B.K.; Yu, D.S. Effective Dual Polysulfide Rejection by a Tannic Acid/Fe(III) Complex-Coated Separator in Lithium-Sulfur Batteries. ACS Appl. Mater. Interfaces 2018, 10, 12708–12715. [Google Scholar] [CrossRef]
  45. Liu, N.; Wang, L.; Tan, T.Z.; Zhao, Y.; Zhang, Y.G. TiO2/GO-coated functional separator to suppress polysulfide migration in lithium-sulfur batteries. Beilstein J. Nanotechnol. 2019, 10, 1726–1736. [Google Scholar] [CrossRef]
  46. Xu, G.Y.; Yan, Q.B.; Wang, S.T.; Kushima, A.; Bai, P.; Liu, K.; Zhang, X.G.; Tang, Z.L.; Li, J. A thin multifunctional coating on a separator improves the cyclability and safety of lithium sulfur batteries. Chem. Sci. 2017, 8, 6619–6625. [Google Scholar] [CrossRef]
  47. Bhargav, A.; He, J.; Gupta, A.; Manthiram, A. Lithium-Sulfur Batteries: Attaining the Critical Metrics. Joule 2020, 4, 285–291. [Google Scholar] [CrossRef]
  48. Cheng, P.; Guo, P.Q.; Liu, D.Q.; Wang, Y.R.; Sun, K.; Zhao, Y.G.; He, D.Y. Fe3O4/RGO modified separators to suppress the shuttle effect for advanced lithium-sulfur batteries. J. Alloys Compd. 2019, 784, 149–156. [Google Scholar] [CrossRef]
  49. Lv, X.; Lei, T.; Wang, B.; Chen, W.; Jiao, Y.; Hu, Y.; Yan, Y.; Huang, J.; Chu, J.; Yan, C.; et al. An Efficient Separator with Low Li-Ion Diffusion Energy Barrier Resolving Feeble Conductivity for Practical Lithium–Sulfur Batteries. Adv. Energy Mater. 2019, 9, 1091800. [Google Scholar] [CrossRef]
  50. Zhao, M.; Peng, H.J.; Zhang, Z.W.; Li, B.Q.; Chen, X.; Xie, J.; Chen, X.; Wei, J.Y.; Zhang, Q.; Huang, J.Q. Activating Inert Metallic Compounds for High-Rate Lithium-Sulfur Batteries through In Situ Etching of Extrinsic Metal. Angew. Chem. 2019, 58, 3779–3783. [Google Scholar] [CrossRef]
  51. Zhang, H.Y.; Dai, R.Q.; Zhu, S.; Zhou, L.Z.; Xu, Q.J.; Min, Y.L. Bimetallic nitride modified separator constructs internal electric field for high-performance lithium-sulfur battery. Chem. Eng. J. 2022, 429, 132454. [Google Scholar] [CrossRef]
  52. Hong, X.J.; Song, C.L.; Yang, Y.; Tan, H.C.; Li, G.H.; Cai, Y.P.; Wang, H.X. Cerium Based Metal-Organic Frameworks as an Efficient Separator Coating Catalyzing the Conversion of Polysulfides for High Performance Lithium-Sulfur Batteries. ACS Nano 2019, 13, 1923–1931. [Google Scholar] [CrossRef] [PubMed]
  53. Peng, L.; Yu, Z.J.; Zhang, M.K.; Zhen, S.Y.; Shen, J.H.; Chang, Y.; Wang, Y.; Deng, Y.F.; Li, A.J. A novel battery separator coated by a europium oxide/carbon nanocomposite enhances the performance of lithium sulfur batteries. Nanoscale 2021, 13, 16696–16704. [Google Scholar] [CrossRef] [PubMed]
  54. Wang, Z.; Feng, M.; Sun, H.; Li, G.R.; Fu, Q.; Li, H.B.; Liu, J.; Sun, L.Q.; Mauger, A.; Julien, A.M.; et al. Constructing metal-free and cost-effective multifunctional separator for high-performance lithium-sulfur batteries. Nano Energy 2019, 59, 390–398. [Google Scholar] [CrossRef]
  55. Lin, H.L.; Zhang, S.L.; Zhang, T.R.; Cao, S.; Ye, H.L.; Yao, Q.F.; Zheng, G.W.; Lee, J.Y. A Cathode-Integrated Sulfur-Deficient Co9S8 Catalytic Interlayer for the Reutilization of “Lost” Polysulfides in Lithium-Sulfur Batteries. ACS Nano 2019, 13, 7073–7082. [Google Scholar] [CrossRef]
  56. Song, S.L.; Shi, L.Y.; Lu, S.C.; Pang, Y.C.; Wang, Y.K.; Zhu, M.; Ding, D.W.; Ding, S.J. A new polysulfide blocker-poly(acrylic acid) modified separator for improved performance of lithium-sulfur battery. J. Membr. Sci. 2018, 563, 277–283. [Google Scholar] [CrossRef]
  57. Chiu, L.L.; Chung, S.H. Composite gel-polymer electrolyte for high-loading polysulfide cathodes. J. Mater. Chem. A 2022, 10, 13719–13726. [Google Scholar] [CrossRef]
  58. Li, D.D.; Yang, J.F.; Xu, X.; Wang, X.W.; Chen, J.T.; Xu, J.; Zhao, N. Synergistic inhibitory effect of ultralight CNTs-COOH@Fe3O4 modified separator on polysulfides shuttling for high-performance lithium–sulfur batteries. J. Membr. Sci. 2020, 611, 118300. [Google Scholar] [CrossRef]
  59. Wang, J.T.; Zhai, P.F.; Zhao, T.K.; Li, M.J.; Yang, Z.H.; Zhang, H.Q.; Huang, J.J. Laminar MXene-Nafion-modified separator with highly inhibited shuttle effect for long-life lithium–sulfur batteries. Electrochim. Acta 2019, 320, 134558. [Google Scholar] [CrossRef]
  60. Zhang, T.Z.; Huang, J.Q.; Peng, H.J.; He, L.Y.; Cheng, X.B.; Chen, C.M.; Zhang, Q. Rational Integration of Polypropylene/Graphene Oxide/Nafion as Ternary-Layered Separator to Retard the Shuttle of Polysulfides for Lithium-Sulfur Batteries. Small 2016, 12, 381–389. [Google Scholar] [CrossRef]
  61. Xu, R.; Sun, Y.Z.; Wang, Y.F.; Huang, J.Q.; Zhang, Q. Two-dimensional vermiculite separator for lithium sulfur batteries. Chin. Chem. Lett. 2017, 28, 2235–2238. [Google Scholar] [CrossRef]
  62. Saroha, R.; Heo, J.; Li, X.Y.; Angulakshmi, N.; Lee, Y.; Ahn, H.J.; Ahn, J.H.; Kim, J.H. Asymmetric separator integrated with ferroelectric-BaTiO3 and mesoporous-CNT for the reutilization of soluble polysulfide in lithium-sulfur batteries. J. Alloys Compd. 2022, 893, 162272. [Google Scholar] [CrossRef]
  63. Yim, T.; Han, S.H.; Park, N.H.; Park, M.S.; Lee, J.H.; Shin, J.; Choi, J.W.; Jung, Y.J.; Jo, Y.N.; Yu, J.S.; et al. Effective Polysulfide Rejection by Dipole-Aligned BaTiO3Coated Separator in Lithium-Sulfur Batteries. Adv. Funct. Mater. 2016, 26, 7817–7823. [Google Scholar] [CrossRef]
  64. Babu, D.B.; Giribabu, K.; Ramesha, K. Permselective SPEEK/Nafion Composite-Coated Separator as a Potential Polysulfide Crossover Barrier Layer for Li-S Batteries. ACS Appl. Mater. Interfaces 2018, 10, 19721–19729. [Google Scholar] [CrossRef]
  65. Maletti, S.; Podetti, S.; Oswald, S.; Giebeler, L.; Barbero, C.A.; Balach, J. LiV3O8-Based Functional Separator Coating as Effective Polysulfide Mediator for Lithium–Sulfur Batteries. ACS Appl. Energy Mater. 2020, 3, 2893–2899. [Google Scholar] [CrossRef]
  66. Ghazi, Z.A.; He, X.; Khattak, A.M.; Khan, N.A.; Liang, L.B.; Iqbal, A.; Wang, J.X.; Sin, H.; Li, L.S.; Tang, Z.Y. MoS2 /Celgard Separator as Efficient Polysulfide Barrier for Long-Life Lithium-Sulfur Batteries. Adv. Mater. 2017, 29, 1606817. [Google Scholar] [CrossRef]
  67. Chang, Z.; Qiao, Y.; Wang, J.; Deng, H.; Zhou, H.S. Two-dimensional metal–organic framework with perpendicular one-dimensional nano-channel as precise polysulfide sieves for highly efficient lithium–sulfur batteries. J. Mater. Chem. A 2021, 9, 4870–4879. [Google Scholar] [CrossRef]
  68. Huang, Y.C.; Yen, Y.J.; Tseng, Y.H.; Chung, S.H. Module-Designed Carbon-Coated Separators for High-Loading, High-Sulfur-Utilization Cathodes in Lithium-Sulfur Batteries. Molecules 2021, 27, 228. [Google Scholar] [CrossRef]
  69. Wu, S.L.; Shi, J.Y.; Nie, X.L.; Yao, Y.M.; Jiang, F.; Wei, Q.F.; Huang, F.L. Microporous Cyclodextrin Film with Funnel-type Channel Polymerized on Electrospun Cellulose Acetate Membrane as Separators for Strong Trapping Polysulfides and Boosting Charging in Lithium–Sulfur Batteries. Energy Environ. Mater. 2022. [Google Scholar] [CrossRef]
  70. Pei, H.J.; Yang, C.Y.; Wu, Q.Y.; Zhou, X.P.; Xie, X.L.; Hwang, B.; Ye, Y.S. Ion-selective aramid nanofiber-based Janus separators fabricated by a dry-wet phase inversion approach for lithium–sulfur batteries. J. Mater. Chem. A 2022, 10, 5317–5327. [Google Scholar] [CrossRef]
  71. Sun, J.; Mu, Q.; Kimura, H.; Murugadoss, V.; He, M.; Du, W.; Hou, C.X. Oxidative degradation of phenols and substituted phenols in the water and atmosphere: A review. Adv. Compos. Hybrid Mater. 2022, 5, 627–640. [Google Scholar] [CrossRef]
  72. Guo, Y.; Li, J.; Pittcheri, R.; Zhu, J.H.; Wen, P.; Qiu, Y.J. Electrospun Ti4O7/C conductive nanofibers as interlayer for lithium-sulfur batteries with ultra long cycle life and high-rate capability. Chem. Eng. J. 2019, 355, 390–398. [Google Scholar] [CrossRef]
  73. Wang, X.B.; Luo, Y.H.; Wang, H.Y.; Wu, C.C.; Zhang, Z.S.; Lia, J.D. Tow-dimensional metal organic framework decorated porous carbon fiber as efficient interlayer for lithium-sulfur battery. J. Electroanal. Chem. 2021, 897, 115564. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of PEO/LiTFSI coating hindering LiPSs [36].
Figure 1. Schematic diagram of PEO/LiTFSI coating hindering LiPSs [36].
Membranes 12 00790 g001
Figure 2. The relationship between the binding energy (Eb) of the doping element−Li2S4 and the electronegativity of the element [41].
Figure 2. The relationship between the binding energy (Eb) of the doping element−Li2S4 and the electronegativity of the element [41].
Membranes 12 00790 g002
Figure 3. Cycle performance of LSBs using different separators at 0.2 C [42].
Figure 3. Cycle performance of LSBs using different separators at 0.2 C [42].
Membranes 12 00790 g003
Figure 4. (a) Schematic diagram of LBSs with TiO2/GO coated functional separator. (b) Cyclic performance and coulombic efficiency of the LSBs batteries with different coated separators at a rate of 0.2 C [45].
Figure 4. (a) Schematic diagram of LBSs with TiO2/GO coated functional separator. (b) Cyclic performance and coulombic efficiency of the LSBs batteries with different coated separators at a rate of 0.2 C [45].
Membranes 12 00790 g004
Figure 5. (a) Mechanism of in-situ etching of LiPSs by bimetallic nitride. (b) Direction of the electric field inside the LSBs [51].
Figure 5. (a) Mechanism of in-situ etching of LiPSs by bimetallic nitride. (b) Direction of the electric field inside the LSBs [51].
Membranes 12 00790 g005
Figure 6. (a) Schematic representation of the catalytic conversion of LiPSs on the surface of Ce-MOFs/CNT separators. (b) Cyclic performance of cells with different separator at 1 C for 800 cycles (2.5 mg/cm2 sulfur loading) [52].
Figure 6. (a) Schematic representation of the catalytic conversion of LiPSs on the surface of Ce-MOFs/CNT separators. (b) Cyclic performance of cells with different separator at 1 C for 800 cycles (2.5 mg/cm2 sulfur loading) [52].
Membranes 12 00790 g006
Figure 7. Adsorption energies of different LiPSs on the Eu2O3 (222) plane [53].
Figure 7. Adsorption energies of different LiPSs on the Eu2O3 (222) plane [53].
Membranes 12 00790 g007
Figure 8. Schematic diagram of ternary PP/GO/Nafion separator [60].
Figure 8. Schematic diagram of ternary PP/GO/Nafion separator [60].
Membranes 12 00790 g008
Figure 9. (a) Schematic diagram of LSBs sheet vermiculite separator, (b) vermiculite sheet structure [61].
Figure 9. (a) Schematic diagram of LSBs sheet vermiculite separator, (b) vermiculite sheet structure [61].
Membranes 12 00790 g009
Figure 10. Schematic diagram of the polarization process of BTO/PE and the repulsive effect of polarized BTO on LiPSs [63].
Figure 10. Schematic diagram of the polarization process of BTO/PE and the repulsive effect of polarized BTO on LiPSs [63].
Membranes 12 00790 g010
Figure 11. Schematic representation of the adsorption and redox transformation of LiPSs at the LiV3O8 diaphragm interface [65].
Figure 11. Schematic representation of the adsorption and redox transformation of LiPSs at the LiV3O8 diaphragm interface [65].
Membranes 12 00790 g011
Figure 12. (a) Schematic illustration of granular and 2D sheet MOFs for inhibiting LiPSs; (b) schematic illustration of 2D CuBDC sheets with crystal structures of 1D nanochannels and 2D CuBDC sheets [66].
Figure 12. (a) Schematic illustration of granular and 2D sheet MOFs for inhibiting LiPSs; (b) schematic illustration of 2D CuBDC sheets with crystal structures of 1D nanochannels and 2D CuBDC sheets [66].
Membranes 12 00790 g012
Figure 13. Schematic illustration of ANF-JS applied to LiPSs and Li-ion transport in LSBs [70].
Figure 13. Schematic illustration of ANF-JS applied to LiPSs and Li-ion transport in LSBs [70].
Membranes 12 00790 g013
Figure 14. (a) Pore size distribution and N2 adsorption−desorption isotherms of TCNFs; (b) cycling performance of CMK3−S/TCNFs at high current density 3 C [72].
Figure 14. (a) Pore size distribution and N2 adsorption−desorption isotherms of TCNFs; (b) cycling performance of CMK3−S/TCNFs at high current density 3 C [72].
Membranes 12 00790 g014
Figure 15. (a) Schematic diagram of preparation process of ZIF-7@PCF interlayer; (b) the rate capability of cells with ZIF-7@PCF, ZnO@PCF interlayer, and without interlayer [73].
Figure 15. (a) Schematic diagram of preparation process of ZIF-7@PCF interlayer; (b) the rate capability of cells with ZIF-7@PCF, ZnO@PCF interlayer, and without interlayer [73].
Membranes 12 00790 g015
Table 1. The properties of PP separators modified with different carbon materials in LSBs.
Table 1. The properties of PP separators modified with different carbon materials in LSBs.
MaterialsTotal Pore Volume/cm3 g−1Surface Area/m2 g−1Coating Binder ContentArea Density /mg cm−2Cathode Sulfur Content/wt%S Loading/mg cm−2Capacity/(mAh g−1) (Rate)Cycling Performance/(mAh g−1) (Cycles, Rate)Capacity Decay/%Ref.
Super P--None0.2551.1–1.31289 (0.5 C)828 (200/0.2 C)0.19[26]
MWCNTs2.76410.42None0.1755-1107 (0.5 C)881 (150/0.2 C)0.14[27]
PG3.361144320 wt% PVP0.54631.8–2.01165 (0.5 C)877 (150/0.5 C)-[28]
PC/MWCNT0.1783.410 wt% PVDF0.51701.6–1.7911 (0.5 C)659 (200/0.5 C)0.138[29]
GCFF--None-600.71280.14 (0.2 C)1004.62 (100/0.2 C)-[30]
CFs--None0.166011280.14 (0.2 C)683 (500/0.5 C)0.071[31]
rGO/CB2.334861.12None---1014.5 (0.2 C)850.9 (100/0.2 C)0.17[32]
CNT/AC0.19131220 wt% PVDF-70-1495.6 (0.2 C)742 (200/0.2 C)0.25[33]
HCNF/rGO--None1.2601.41318.4 (0.2 C)779.1(100/1 C)0.13[34]
mesoC2.98435 wt% super P and
10 wt% PVDF-HFP
0.5703.51378 (0.2 C)1021 (100/0.5 C)0.081[35]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, B.-W.; Sun, B.; Fu, P.; Liu, F.; Zhu, C.; Xu, B.-M.; Pan, Y.; Chen, C. A Review of the Application of Modified Separators in Inhibiting the “shuttle effect” of Lithium–Sulfur Batteries. Membranes 2022, 12, 790. https://doi.org/10.3390/membranes12080790

AMA Style

Zhang B-W, Sun B, Fu P, Liu F, Zhu C, Xu B-M, Pan Y, Chen C. A Review of the Application of Modified Separators in Inhibiting the “shuttle effect” of Lithium–Sulfur Batteries. Membranes. 2022; 12(8):790. https://doi.org/10.3390/membranes12080790

Chicago/Turabian Style

Zhang, Bo-Wen, Bo Sun, Pei Fu, Feng Liu, Chen Zhu, Bao-Ming Xu, Yong Pan, and Chi Chen. 2022. "A Review of the Application of Modified Separators in Inhibiting the “shuttle effect” of Lithium–Sulfur Batteries" Membranes 12, no. 8: 790. https://doi.org/10.3390/membranes12080790

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