Zhiheng
Ren‡
,
Jixiao
Li‡
,
Minghui
Cai
,
Ruonan
Yin
,
Jianneng
Liang
*,
Qianling
Zhang
,
Chuanxin
He
,
Xiantao
Jiang
and
Xiangzhong
Ren
*
College of Chemistry and Environmental Engineering, Shenzhen University, Shenzhen, Guangdong 518060, China. E-mail: jlian46@uwo.ca; renxz@szu.edu.cn; Fax: +86-755-26558134; Tel: +86-755-26558134
First published on 13th December 2022
Polymer electrolytes (PEs) synthesized from an in situ polymerization strategy are considered as promising candidates to improve interfacial compatibility. However, most in situ formed PEs still suffer from problems such as low lithium-ion transference number (tLi+) and insufficient ionic conductivity at room temperature (RT). Herein, a series of copolymer electrolytes (CPEs) consisting of 1,3-dioxolane (DOL) and 1,3,5-trioxane (TXE) monomers are synthesized. Due to the additive of succinonitrile (SN), and the promoted dissociation of lithium salt, the obtained polymer electrolyte (SN-CPE) exhibits an excellent ionic conductivity (4.06 × 10−4 S cm−1) and a high lithium-ion transference number (tLi+ = 0.881) at RT, as well as a high voltage oxidation window of 5.1 V. Thanks to the in situ formed intimate contacted electrode/electrolyte interface, the symmetric Li/Li cell can achieve long-term cycling stability over 1500 h without a short circuit. A LiF-rich organic–inorganic composite layer is formed on both the solid electrolyte interphase (SEI) and cathode electrolyte interphase (CEI) as demonstrated by XPS, TEM, and SEM analyses. Thereafter, the LiFePO4/Li cell has a high capacity retention of 84.1% after 900 cycles at RT, and it can still work effectively at high temperatures (e.g. 80 °C). Furthermore, SN-CPE exhibits excellent electrochemical performance and safety in the LiCoO2/Li cell and LiFePO4/graphite pouch cell, which again supports the promising application of SN-CPE.
The promising SSEs for SSLMBs can be generally classified into inorganic solid electrolytes (ISEs) and polymer electrolytes (PEs).11–13 PEs have good flexibility and provide relatively good interfacial contact. However, the performance of PE-based LMBs is still unsatisfactory, which should be attributed to some inherent drawbacks of PEs, including the low ionic conductivity, the narrow electrochemical window and the low lithium-ion transference number.14,15 In addition, the contact between the PE and the electrode only remains at the surface. As a result, no SSE can penetrate the porosities of the thick electrode, and the electrochemical performances of solid-state batteries (SSBs) are bad. Therefore, further investigations on how to build an intimate contact interface between SSEs and electrodes are still critical.16
Recently, in situ polymerization strategies for synthesizing PEs and building SSBs have received plenty of attention. This is because such an approach has great potential for achieving highly ionic conductive PEs, low interfacial resistance and a simple fabrication process for commercial applications.17 The in situ polymerization strategy generally involves pre-injection of a low-viscosity precursor into the cell followed by the initiation of a free-radical polymerization reaction or ring-opening polymerization reaction of the monomer with the help of a thermal or chemical initiator. Therefore, PEs obtained from the in situ process can fill the gaps/voids of the electrode, thus achieving a tight interfacial contact between the PEs and electrodes.18 Up to now, PEs synthesized through an in situ polymerization method have been widely reported, including poly(vinylene carbonate) (PVC),19 poly(methyl methacrylate) (PMMA),20 poly-tetrahydrofuran (PTHF),21 poly(triethylene glycol diacrylate) (PTEGDA),22 poly(1,3-dioxolane) (PDOL)23 and poly(trimethylpropane triacrylate) (ETPTA)24-based PEs. To further improve the interfacial compatibility, mechanical properties and the overall electrochemical performances of in situ polymerized PEs and SSBs, the strategies including integrations of inorganic fillers,24,25 making crosslinking polymers and copolymers,26–28 and introducing functional groups into the polymer chains19,29,30 have also been reported.
1,3-Dioxolane (DOL) is one of the most popular monomers for synthesizing in situ polymerized PEs for SSBs because of the facile initiation of the ring-opening polymerization reaction by some Lewis acid salts such as Al(OTf)3,23 LiFSI,31 LiPF6 (ref. 32) and LiDFOB33 under mild conditions. For example, Archer et al.23 reported the cationic ring-opening polymerization of DOL (PDOL) initiated by a low concentration of Al(OTf)3 salt. Thanks to the good wettability between the liquid precursor and electrodes, the gel polymer electrolyte (GPE) after in situ solidification effectively overcame the high interfacial resistance problem. Guo et al.32 investigated the conversion of a typical DOL/DME-based liquid electrolyte to a quasi-solid-solid GPE through the catalyst of LiPF6. The in situ formed polymer networks changed the solvation structure, disrupting the high voltage limitations of conventional ether-based electrolytes. Though there are many investigations on PDOL-based PEs, they still suffer from the problems of low ionic conductivity, low lithium-ion transference number, poor compatibility towards the lithium metal anode, and poor electrochemical oxidation tolerance properties. As a result, SSLMBs with pristine PDOL PEs show inferior electrochemical performance.
To enhance the overall electrochemical performance of PDOL-based PEs in SSLMBs, herein, we developed a copolymerization strategy to synthesize copolymer electrolytes (CPEs), where two monomers DOL and 1,3,5-trioxane (trioxymethylene (TXE)) were used for the copolymerization reaction. Besides DOL, TXE is also a common industrial chemical and its ring-opening polymerization products (polyformaldehyde (POM)) have been reported for applications in SSLMBs.34 Therefore, in this study, TXE was used to randomly copolymerize with DOL, forming a copolymer P(DOL-TXE), which disrupts the regularity of the PDOL chain and significantly reduces the crystallinity of PDOL-based PEs. Meanwhile, the energy level of the highest occupied molecular orbital (HOMO) of the copolymer is lower compared to other counterparts, which confers a better anti-oxidation capacity. The SN plasticizer is added for increasing the RT ionic conductivity of CPE and labelled as SN-CPE. An ionic conductivity of 4.06 × 10−4 S cm−1 at 25 °C is thus achieved. The strong polar cyano could effectively restrict the movement of anions and loosen the strong coordination effect of Li–O, promoting the transportation of Li+ along the polymer chain. Therefore, SN-CPE has a Li+ transference number of 0.881, which is almost the highest value among the reported PEs. The electrochemical stability window of SN-CPE is 5.1 V. Thereafter, stable lithium plating/stripping cycling over 1500 h is achieved in the Li|SN-CPE|Li symmetric cell. Excellent cycling stability of the LFP|SN-CPE|Li SSB with a capacity retention of 84.1% after 900 cycles at 25 °C is illustrated. Besides, SN-CPE also exhibits good performance in a LiCoO2/Li battery and good stability under high temperature working conditions. This study provides new insights into the design of high-performance PEs for wide working temperature ranges and high energy density SSLMBs.
The electrochemical impedance spectroscopy (EIS) study was performed via a CHI760E electrochemical working station, where the frequency ranged from 0.1 M Hz to 0.1 Hz and the voltage amplitude was 5 mV. The ionic conductivities (σ) of PEs were measured in a symmetric stainless steel cell and their values were calculated according to the following eqn (1):
![]() | (1) |
The activation energy (Ea) of PE was derived from the slope of the Arrhenius plot.
The lithium-ion transference numbers (tLi+) of PEs were calculated according to the chronoamperometry study and EIS measurements of Li|PE|Li symmetric cells. Eqn (2) can be applied for calculating the value of tLi+
![]() | (2) |
To evaluate the electrochemical stability window, linear sweep voltammetry (LSV) was performed on SS|PE|Li cells in the voltage range of 2.0–6.0 V under the voltage scan rate of 1 mV s−1, where PE was used as the separator. The electrochemical floating test was conducted in an NCM811/Li cell at stepwise voltages from 4.0 V to 4.6 V, with 10 h holding time at each voltage, to examine the oxidative stability of PE. The impedances of LFP/Li and Li/Li cells were evaluated by EIS in the frequency range of 1 M Hz–0.1 Hz, under a voltage amplitude of 10 mV.
Eb = E(PDOL-Li) − E(Li+) − E(PDOL) | (3) |
The calculating methods of binding energies of CPE-Li+ and SN-Li+ are the same as eqn (3).
The GF membrane is used as the 3D porous skeleton to sponge the PE precursor, and as the separator for batteries. After filling with PE, the white GF membrane transformed into a translucent membrane, and a uniform interpenetrating network was consequently formed, which ensured continuous and fast Li+ transport (Fig. S1†). It is interesting that the GF membrane with PEs filling exhibited higher elasticity (Fig. S2†), which was beneficial for improving the electrode/electrolyte interface contact and blocking lithium dendrites. To illustrate the characteristics and advantages of in situ polymerization of precursors more clearly inside the cell, a Li (in situ)/SN-CPE/Li (ex situ) cell was constructed, and its interface is studied. As shown in Fig. S3,† the Li (ex situ)/SN-CPE interface prepared by ex situ method is not tight, which will lead to high interfacial impedance and uneven lithium flux. In contrast, the Li (in situ)/SN-CPE interface is intimately contacted. Therefore, the in situ polymerization method can endow the lithium battery with a low interfacial impedance and uniform lithium flux.
Subsequently, 1H NMR was performed to study the chemical structures of samples and prove the ring-open polymerization of monomers,40 and the results are shown in Fig. 1d. First, 1H NMR peaks at 4.78 ppm and 3.78 ppm corresponding to the H (a) and H (b) in DOL molecule were shifted to 4.64 ppm (d) and 3.61 ppm (e), respectively, in PDOL and CPE, SN-CPE samples, representing the formation of –(O–CH2–O–CH2)– and –(O–CH2CH2–O)– units. Secondly, the 5.12 ppm peak comes from H (c) in the O–CH2–O structure of the TXE monomer (Fig. S4†). The NMR peaks from DOL or TXE monomers can be observed in the 1H NMR spectra of PDOL, CPE and SN-CPE samples, which indicates that DOL and TXE monomers are not completely converted into copolymers and the presence of trace amounts of the monomers would be favourable to the increase of ionic conductivity due to the plasticizer effect. In addition, the 4.72 ppm peak should have originated from the H (f) in the –(OCH2)3– structure from the ring-opening polymerization of the TXE monomer. The presence of multiple peaks at the ranges between 3.5 and 3.8 ppm and between 4.6 and 4.9 ppm in CPE and SN-CPE samples indicates the random sequence of DOL and TXE units during the polymerization reaction. The peak at 2.91 ppm comes from the proton of SN. FTIR was used to further demonstrate the structural and molecular interactions of PEs (Fig. 1e). For DOL monomer, a distinct C–H out-of-plane vibrational peak located at 910 cm−1, C–O–C vibrational peak located at 1000–1150 cm−1, and the saturated C–H stretching vibration peak located at 2850–3000 cm−1 were observed. The former belongs to the typical features of the ring-ether structure.42,43 Compared to the DOL monomer, a long-chain –(CH2)n– vibrational mode located at 840 cm−1 appeared in these PEs samples, demonstrating the transition from cyclic monomers to long-chain polymers. In addition, the C–O–C vibrational modes of CPE and SN-CPE were shifted to 1110 cm−1 and 1130 cm−1, respectively, demonstrating interactions among DOL, TXE and SN. Moreover, compared to PDOL and CPE, the CO stretching vibration double peaks belong to LiDFOB in SN-CPE shifted from 1753 cm−1 and 1791 cm−1 to 1764 cm−1 and 1801 cm−1, respectively, which implies more dissociation of LiDFOB in SN-CPE, possibly attributed to the strong interaction between LiDFOB and SN. The gel permeation chromatography (GPC) study again supported the polymerization reaction of monomers. As shown in Table S1,† the high average molecular weights (Mw) of PDOL, CPE and SN-CPE are 6975, 5570, and 4479, respectively. It seems that the copolymerization strategy tends to generate smaller molecular weight polymers under the same initiation conditions, which sacrifices some mechanical properties, but facilitates the rapid transport of Li+ ions. In addition, the polydispersity index (Mn/Mw) of SN-CPE is 1.67, which is slightly lower than those of PDOL and CPE, indicating that SN-CPE has better homogeneity.
The crystallinity of PE is an important parameter for evaluating the Li+ transport capacity. The more the amorphous polymer phase, the higher the ionic conductivity. The crystallinity of PEs was visually characterized by X-ray powder diffraction (XRD), as shown in Fig. S5.† PDOL showed a clear crystalline peak at 2θ = 22.6°, while no diffraction peak is observed in the other two copolymer systems, indicating that the introduction of TXE would cause the copolymers to be less aligned, thus reducing the crystallinity of PEs.44 The thermal behaviours of PEs were analysed by applying a DSC study. As shown in Fig. S6,† SN-CPE exhibited a lower glass transition temperature (Tg) and a lower melting temperature (Tm), indicating that SN-CPE is almost amorphous at the working temperature.27 Fig. S7† shows the TGA results of these PEs. It was reported that 25–80 °C is the working temperature range of the vast majority of LMBs.45 So, the stability of PEs at this temperature range was studied. TGA results suggest that SN-CPE and CPE had almost no weight loss in this temperature range, indicating their high safety performance.
In general, the ideal electrolyte should have a low HOMO and a high LUMO to obtain a wide voltage stability window thermodynamically.46,47 The HOMO and LUMO of the DOL monomer, PDOL, TXE monomer, P(DOL-TXE) and SN were calculated separately using DFT. The results in Fig. 2a show that TXE and P(DOL-TXE) have a lower HOMO value compared to DOL and PDOL, suggesting that the introduction of the TXE unit will increase the oxidation tolerance of PEs. For SN, the highly polar cyano (–CN) has a strong electron-absorbing effect, which confers a very low HOMO to SN. Therefore, SN-CPE should have a stronger capacity against oxidation. The LUMO value represents the tendency of the components to be reduced. These results suggest that ether-based DOL and TXE monomers have a higher LUMO compared to PDOL and P(DOL-TXE), respectively, indicating that the polymerization can decrease the capacity against reduction. The relatively lower LUMO for P(DOL-TXE) will not significantly sacrifice its stability toward the Li anode since a LiF and Li3N rich, and stale SEI layer was created at the SN-CPE/Li anode interface, as discussed below. Fig. 2b shows the molecular surface electrostatic potential (ESP) maps of PDOL, P(DOL-TXE), and SN with and without Li+ ion coordination, respectively.37,38 Li+ ions migrate through the coordination with ether oxygen (EO) and hop from one site to another. Fig. 2b shows that the more well-arranged polymer chains can provide a denser negative potential distribution, which can increase the Li+ binding sites.48,49 The binding energies between different polymer chains and Li, which are calculated based on the optimized model, were −1.697 eV for PDOL-Li, and −1.281 eV for P(DOL-TXE)-Li, which means Li+ ions can move faster in the main body of the copolymer (Table S2† displays the calculation details). In addition, the binding energy of SN-Li is −1.786 eV, which is more negative than those between both polymer bodies and Li+, indicating that SN can promote the dissociation of lithium salts. The 7Li NMR results are consistent with DFT calculations. As shown in Fig. S8,† compared to PDOL, CPE had an up-field shift of the Li NMR peak, which confirmed that the copolymer structure formed by DOL and TXE weakened the coordination between Li+ and EO of the polymer chain. With the introduction of SN, a low-field shift of Li peak was exhibited in SN-CPE, which suggested that SN was involved in the coordination with Li+, and the tendency of the coordination was stronger than that between Li+ and EO. Therefore, SN acts synergistically with the copolymer to ensure the solvation of salts and the rapid migration of Li+ ions.50
The ionic conductivities of PEs with different mass ratios of DOL, TXE and SN, were evaluated by EIS, and the results are illustrated in Fig. 3a and S9.† It is found that with the mass ratio of 8:
2
:
2 (DOL
:
TXE
:
SN), SN-CPE can achieve the highest ionic conductivity of 4.06 × 10−4 S cm−1 at 25 °C. The Arrhenius curves of PDOL, CPE and SN-CPE are shown in Fig. 3b, depicting the relationship between the temperature and ionic conductivity at 25–80 °C. All these PEs samples fit the Vogel–Tamman–Fulcher mode well,51 and the calculated activation energies (Ea) of PDOL, CPE, and SN-CPE are 0.61 eV, 0.55 eV and 0.35 eV, respectively.
![]() | ||
Fig. 3 (a) Ionic conductivity of PEs with various DOL, TXE and SN mass ratios at 25 °C on the ternary contour diagram. (b) Arrhenius plots and (c) linear sweep voltammetry (LSV) of PDOL, CPE and SN-CPE (inset is the enlarged profile at 3.5 to 5.5 V). Chronoamperometry curves with a step voltage of 10 mV (the insets display the EIS plot before and after polarization) of (d) CPE and (e) SN-CPE. (f) The plot of lithium-ion transference number against ionic conductivity for SN-CPE and other reported PEs.18,54–63 The colour scale of the symbol represents the high voltage stability window of the PEs. |
The lower activation energy implies a lower energy barrier for Li+ migration, which makes PE easier to achieve a rapid Li+ migration.49 Therefore, the higher ionic conductivity and lower Li+ migration energy barrier of SN-CPE mean it is superior than other PEs.
The electrochemical stability window of PE is crucial when coupling with high-voltage cathodes. It has been well known that the electrochemical stability window of ether-based electrolytes is usually below 4.0 V vs. Li/Li+.52 In this study, we found that the copolymerization strategy can enhance the oxidation tolerance of PE by changing the molecular structure of copolymer chains. The electrochemical oxidation windows of CPE and SN-CPE illustrated a clear increase compared to that of PDOL PE, as determined by LSV studies (Fig. 3c). It presents that PDOL PE exhibited an oxidation voltage of 4.5 V, which increased to 4.7 V and 5.1 V for CPE and SN-CPE, respectively. This result was consistent well with the HOMO result in Fig. 2. Electrochemical float tests were used to further evaluate the oxidation tolerance of PEs in SSBs. As shown in Fig. S10,† the leakage currents of NCM811|SN-CPE|Li cell under high voltage (4.2–4.6 V) are always smaller than those with PDOL and CPE as the electrolytes, and the leakage currents can decay faster, indicating that SN-CPE had better oxidation stability, indicating that it can be matched with high-voltage cathode materials for SSBs.53
The lithium-ion transference number (tLi+) is another important parameter for evaluating the performance of PEs, which reflects how much percentage of the ionic conductivity is contributed by the movement of Li+ ions in the electrolyte. A higher tLi+ can alleviate the concentration polarization and inhibit the formation of lithium dendrites.54 The tLi+ of these PEs was determined using the constant potential polarization method.55 Compared to PDOL PE (tLi+ = 0.478, Fig. S11†) and CPE (tLi+ = 0.514, Fig. 3d), SN-CPE with a tLi+ as high as 0.881 (Fig. 3e) exhibited a closed single ion conductor behaviour. The reasons for such a high tLi+ should be (1) the weak complexation between EO and Li+ ions resulting in fast Li+ transportation; (2) the anions are confined in the copolymer framework, making them less likely to transfer over long distances. In addition, the high dielectric constant of SN (∼55) provides superb lithium salt dissociation, resulting in more free Li+ ions and fast ionic conduction.56 The above computational and experimental results indicate that the as-prepared SN-CPE had a fast Li+ migration ability by suppressing the movement of anions, which is favourable for the electrochemical performance of lithium batteries. Compared to other reported PEs, SN-CPE exhibits excellent properties in terms of tLi+, ionic conductivity, and electrochemical stability window (Fig. 3f).21,57–66
The lithium plating/stripping behaviour of symmetric Li/Li cells with different PEs was studied to evaluate the interface stabilities between the Li metal anode and PEs. As shown in Fig. 4a, at a current density of 0.1 mA cm−2, the symmetric cell using PDOL PE had a relatively high overpotential of up to 120 mV, and an increase in polarization voltage upon cycling. It experienced short circuiting after around 270 h charge/discharge cycling, possibly due to the continuous generation of dead Li and Li dendrites.
The Li/Li symmetric cell with CPE was able to operate stably for 800 h. However, its polarization voltage increased progressively during cycling, meaning the poor interface stability between the CPE and Li metal anode. Encouragingly, the Li/Li symmetric cell using SN-CPE exhibited a small polarization potential (52 mV), and it was stable even after 1500 h cycles, indicating that SN-CPE was stable towards the Li metal anode. On further increasing the current density and the capacity (0.2 mA cm−2, 0.2 mA h cm−2, and 1 mA h cm−2), the Li/Li symmetric cell with SN-CPE can still deliver great cycling stability, indicating the superior electrochemical performance of SN-CPE (Figs. 4c and S12†). However, PDOL and CPE exhibited rapid increases in overpotentials, indicating severe side reactions and poor reaction kinetics at a higher current density.
The interfacial stability between PEs and Li was further evaluated by EIS. As shown in Fig. 4b, the interfacial impedance (Ri) of the Li|SN-CPE|Li cell was pretty low, which was about 80 Ω. It increased to 140 Ω after 100 cycles, which was much lower than those of the Li|PDOL|Li cell (370 Ω) and Li|CPE|Li cell (295 Ω). The low interfacial impedance also suggests good interfacial contact and high compatibility between the SN-CPE and Li metal anode. To verify the effect of SN-CPE in inhibiting Li dendrite growth, the surface morphologies of Li metal anodes after 100 cycles in Li/Li symmetric cells with different PEs were studied by FE-SEM. As shown in Fig. 4g and the schematic illustrated in Fig. 4d, the surface of the Li metal anode from the Li|PDOL|Li cell after cycling showed an uneven and mossy structure. However, the Li metal anode from the Li|CPE|Li cell was flattened but it also presented some irregular protrusions (Fig. 4e and h). Notably, the Li metal anode from the Li|SN-CPE|Li cell had a very smooth and condense surface morphology, indicating uniform Li plating/stripping performance (Fig. 4f and i). The FE-SEM study results again illustrated the stable interface between the SN-CPE and Li metal anode.
The composition of the SEI layer is critical to the stability of the Li metal cathode. An appropriate SEI layer contributes to fast Li+ transport, uniform Li deposition and circumvents unfavourable side reactions between the electrolyte and Li anode.7 To study the chemical components of the SEI layer between the SN-CPE and Li metal anode, and to reveal the mechanism for the stable Li plating/stripping performance, XPS with ion etching for an in-depth analysis was performed.67 As shown in Fig. 5a, in the C 1s XPS spectrum, C–C–O (286.6 eV) and O–C–O (288.0 eV) of the ether-based compounds came from the ring-opening polymerized DOL and TXE. The carbonyl carbon (O–CO, 289.5 eV) was originally from the decomposed products of CPE.33 In the F 1s XPS spectrum (Fig. 5b), there were distinct signals at 684.8 eV and 688.4 eV, which could be attributed to C–F and LiF, respectively.68 LiF has been widely regarded as a crucial SEI component for resisting lithium dendrite growth and inhibiting interface side reactions.69 The O 1s spectra in Fig. 5c can be divided into three segments: lithium oxide (ROLi, 530.4 eV), carbonate (Li2CO3/O–C
O, 531.5 eV) and ether oxygen (C–O, 533.0 eV).29 N 1s XPS spectra in Fig. 5d display a peak at 399.5 eV, which can be attributed to TFSI− from LiTFSI, and C
N from SN.70 Another peak at 398.4 eV belongs to Li3N, which should be the decomposition product from SN.71 Li3N is also considered as a favourable component of SEI because of its high ionic conductivity, which is important in reducing the interfacial resistance and inhibiting Li dendrites.72
![]() | ||
Fig. 5 In-depth XPS analyses of the cycled Li metal after 100 cycles at 0.2 mA cm−2 and 0.2 mA h cm−2 with SN-CPE. (a) C 1s, (b) F 1s, (c) O 1s and (d) N 1s. |
As the etching proceeded, chemical information of the inner surface was exposed, and the peak intensities of C–C–O, O–C–O and C–F became weakened and weakened. This means the surfaces of samples were rich with poly-organic materials, while the inside lacks poly-organic materials. On the other hand, the intensities of carbonate, LiF, Li3N and ROLi peaks, which represent inorganic components of the SEI layer, increase as the etching time increase, indicating that the inner layer of the SEI contains more inorganic materials. The organic/inorganic composite SEI layers were beneficial for the stability of the SEI and uniform lithium plating/stripping behaviour.69,73
The application of PEs in SSBs was further investigated. Fig. 6a shows the long-term cycling performance of LFP SSBs with different PEs. Compared to LFP SSBs with CPE and PDOL, which had significant capacity decay after tens of cycles due to low ionic conductivity and poor interfacial stability at RT, the LFP|SN-CPE|Li cell can be stably cycled 900 times at a current density of 0.5C, with a coulomb efficiency close to 100% and a capacity retention of 84.1%.
Besides, the capacity retention and the rate performance of LFP battery with SN-CPE also far exceeded those of the LFP battery with ether-based liquid electrolytes (Fig. 6b). At the current density of 1C, the LFP|SN-CPE|Li cell had an initial capacity of 149.8 mA h g−1, which remains 88.3% after 350 cycles (Fig. 6c). Charge/discharge voltage profiles in Fig. 6d also suggest stable and reversible charge/discharge plateaus from 1 to 300 cycles with only a small increase in polarization voltages (0.17 V to 0.31 V), which was attributed to the stable electrode/electrolyte interface and the fast internal and interfacial Li+ transport. As exhibited in Fig. S13,† the impedance of the LFP|SN-CPE|Li cell was relatively small and it slightly increased in subsequent cycles (below 450 Ω), which was beneficial to the electrochemical performance of SSBs. Unfortunately, the LFP|LE|Li cell exhibited a continuous increase in the polarization potential and a fast capacity decay (Fig. S14†).
The performances and safety of batteries are significantly affected by the working temperature during practical operation. Ether-based electrolytes will decompose and generate gases and other side-reaction products at elevated temperatures, resulting in a significant compromise in safety and suitability of batteries. Here, the LFP|SN-CPE|Li cell was found to be able to operate stably at 80 °C, with an initial capacity of ∼165 mA h g−1 at 1C. It experienced almost no decay within 100 cycles (Fig. S15†), which showed excellent high-temperature adaptability.
To further study the compaticity between the PEs and layer-structured cathode, the LiCoO2|SN-CPE|Li cell was studied. As shown in Fig. 6e, this cell had a high initial capacity of 137.6 mA h g−1 at 0.2C and a high capacity retention (∼86.4%) after 50 cycles. Fig. S16† shows corresponding charge/discharge curves of the first 5 cycles, which were smooth and well coincided, suggesting the decent high voltage performance of SN-CPE.
Further study on LFP|SN-CPE|graphite pouch cell was done to evaluate the potential of SN-CPE in industrial applications. As presented in Fig. 6f, the pouch cell can successfully light up LED lamps. And thanks to the flexibility of SN-CPE and the good interfacial contact, the cell can still work after bending and cutting treatment, which suggested excellent safety properties induced by SN-CPE.
SEM, TEM and XPS analyses were further used to investigate the cathode electrolyte interfaces (CEI) in the LFP|LE|Li battery and LFP|SN-CPE|Li cell, respectively, to reveal the reason for the good interface stability with SN-CPE. Fig. 7a and e shows surface SEM images of the cathode after 20 cycles in LFP batteries with LE and SN-CPE, respectively. It shows that the LFP cathode after cycling with SN-CPE prepared by the in situ polymerization method was totally covered by the PE, suggesting the intimate contact interface and the formation of a favourable CEI layer. The morphology of the CEI layer was observed using ex situ TEM, and the results are shown in Fig. 7b and f. It is found that there was a thick and non-uniform CEI layer at the LFP/LE interface, compared to a uniform thin CEI layer (∼4.6 nm) at the LFP/SN-CPE interface. Such a thin and uniform CEI at the LFP/SN-CPE interface indicates less side reaction. Meanwhile, the chemical components of both CEIs were analysed using XPS (Fig. 7c, d, g and h). Compared to LFP/LE, there was more LiF compound at the LFP/SN-CPE interface, which could effectively inhibit the oxidative decomposition of PE at high potentials since LiF has high oxidation resistance.8 Therefore, SSBs with SN-CPE exhibit better electrochemical performance.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2ta07516d |
‡ Zhiheng Ren and Jixiao Li contributed equally to this work and should be considered as co-first authors. |
This journal is © The Royal Society of Chemistry 2023 |