Next Article in Journal
Anti-Freezing Nanocomposite Organohydrogels with High Strength and Toughness
Next Article in Special Issue
Physically and Chemically Stable Anion Exchange Membranes with Hydrogen-Bond Induced Ion Conducting Channels
Previous Article in Journal
Synergistic Effect between Piperazine Pyrophosphate and Melamine Polyphosphate in Flame Retardant Coatings for Structural Steel
Previous Article in Special Issue
Multisensory Systems Based on Perfluorosulfonic Acid Membranes Modified with Polyaniline and PEDOT for Multicomponent Analysis of Sulfacetamide Pharmaceuticals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Ultrasonication on the Morphology, Mechanical Property, Ionic Conductivity, and Flame Retardancy of PEO-LiCF3SO3-Halloysite Nanotube Composites for Use as Solid Polymer Electrolyte

by
Pattranuch Pongsuk
1 and
Jantrawan Pumchusak
1,2,*
1
Department of Industrial Chemistry, Faculty of Science, Chiang Mai University, 239 Huay Kaew, Su Thep, Mueang, Chiang Mai 50200, Thailand
2
Materials Science Research Center, Faculty of Science, Chiang Mai University, 239 Huay Kaew, Su Thep, Mueang, Chiang Mai 50200, Thailand
*
Author to whom correspondence should be addressed.
Polymers 2022, 14(18), 3710; https://doi.org/10.3390/polym14183710
Submission received: 17 July 2022 / Revised: 24 August 2022 / Accepted: 30 August 2022 / Published: 6 September 2022
(This article belongs to the Special Issue Polymers for Electrochemical Applications)

Abstract

:
PEO-LiCF3SO3-halloysite nanotube (HNT) composites were fabricated by solution casting together with hot compression to form a solid polymer electrolyte (SPE) membrane. Different ultrasonic exposure times were used to disperse HNT nanoparticles in the PEO-20%LiCF3SO3-HNT composite solutions prior to casting. An exposure time of 15 min gave the highest ionic conductivity in the SPE membrane, the ionic conductivity significantly increased by two orders of magnitude from 6.6 × 10−6 to 1.1 × 10−4 S/cm. TEM, FE-SEM, and EDS-mapping were used to study the dispersion of HNTs in the SPE membrane. ATR-FTIR revealed that the bonding of PEO-LiCF3SO3 and PEO-HNT was created. XRD and DSC showed a reduction in the crystallinity of PEO due to HNT addition. The ultrasonication for an optimal period gave uniform dispersion of HNT, reduced the polymer crystallinity and strengthened the tensile property of SPE membrane. Moreover, the electrochemical stability, flame retardance and dimensional stability were improved by the addition of HNT and by ultrasonication.

1. Introduction

Currently, lithium-ion batteries (LIBs) typically contain liquid electrolytes (LEs), which limit operation to ≤4.8 V [1]. LEs are organic liquids or gels the use of which present safety issues due to risks such as generation of volatile gaseous species, leakage, thermal runaway and flammability [2]. Safer high-performance non-flammable electrolytes and solid electrolytes are needed to overcome such risks. Solid electrolytes, consisting of complex Li inorganic compounds or polymer composites, can also inhibit the growth of Li dendrites, can operate at higher voltages and be charged at higher rates, and their use may enable reductions in battery size. Solid polymer electrolytes (SPEs) have attracted increasing attention as potential solid electrolyte materials because of their flexibility in fabrication, chemical stability (non-reactive to Li metal), and wide electrochemical stability [3,4]. Considerable research efforts have been aimed at the technical and commercial development of SPE for solid-state LIBs. Electric vehicles and mobile phones are expected to benefit from solid-state LIBs because of their higher safety during practical use. However, there is a still a need for further SPE development and improvement since current SPE remain limited by poor ionic conductivity and poor mechanical properties. An ionic conductivity higher than 10−4 S/cm is the target for SPE to be considered for practical use in LIBs [5]. Poly (ethylene oxide) (PEO) is usually chosen as a matrix for SPE because lithium salts are well soluble in PEO [6], thus facilitating generation of charge carriers for conducting ions. However, the PEO matrix shows lower ionic conductivity due to the presence of crystalline domains in its structure at room temperature [7]. The mixture of PEO and lithium salt provides ionic conductivity lower than 10−6 S/cm. It has been reported that poor ionic conductivity of SPEs is usually found when high crystallinity polymer matrices are used. Low ion conductivity is also caused if ion aggregation or anion pair formation occur in the system [8,9]. Therefore, to increase ionic conductivity in an SPE system, polymer crystallization must be inhibited. In addition, ion aggregation and anion pair formation should be avoided. Plasticizer has been incorporated into SPE to reduce polymer crystallinity in an effort to increase ionic conductivity; however, this usually reduces mechanical properties and lowers the electrochemical potential of SPE [7]. An interesting method to boost the ionic conductivity of SPE is the addition of inorganic fillers, which also offers the possibility to improve the mechanical properties of SPE [8,9]. However, composite-based SPE that contains ceramic fillers suffer from the problem of particle aggregation at high content which reduces the number of dissociated Li+ ions and decreases ionic conductivity [10]. Ultrasonication has been widely used to disperse nanoparticles in polymer matrices [11,12]. However, excessive ultrasonic exposure times may cause a re-agglomeration of nanoparticles [13] or degrade the properties of nanocomposite due to an increased temperature [12]. Jokar et al. [14] showed that ultrasonication provided ease of dispersion of clay in a polymer matrix, which led to an increase of d-spacing of the clay structure. Parveen et al. [12] found that microcrystalline cellulose (MCM) had slightly decreased agglomeration when ultrasonication times of up to 30 min were applied; however, agglomeration increased at longer ultrasonication durations.
This work has selected halloysite nanotubes (HNTs) for use as nanofiller because of their large surface area, high porosity, and tunable surface chemistry [15]. The chemical structure of HNT consists of silicon dioxide tetrahedrons at the outer surface of the tube and aluminum oxide octahedrons at the inner surface of the tube with a lumen at the center. These characteristics should facilitate ion transfer for ion conductive materials [16,17]. Ultrasonic stirring was applied to facilitate the dispersion of HNTs in a PEO-20%LiCF3SO3-HNT composite. The effects of different ultrasonic exposure times on ionic conductivity, crystallinity, thermal properties, electrochemical stability, flame retardance and dimensional stability and mechanical properties of SPE were investigated.

2. Materials and Methods

2.1. Materials

PEO (M.W. 6 × 106), lithium trifluoromethanesulfonate (LiCF3SO3) and halloysite nanotube (HNTs; Al2Si2O5(OH)4.2H2O) were purchased from Sigma-Aldrich. Tetrahydrofuran (THF) was purchased from RCI Labscan.

2.2. Fabrication of Solid Polymer Electrolyte (SPEs) Nanocomposites

PEO and LiCF3SO3 were used to prepare SPEs as the base composites. HNTs were used as an inorganic nanofiller. The optimum concentration of LiCF3SO3 in PEO was investigated in previous work [16] which found that 20 wt.% of the salt gave the highest ionic conductivity. Therefore, this concentration was used to study the effect of HNT addition. Different amounts of HNT were added into PEO-20%LiCF3SO3 solution to give HNT contents of 5, 10 and 15 wt%. THF was used as the solvent for all composite solutions. A conventional mixing method using a magnetic stirrer was applied to stir each solution at room temperature for 24 h prior to casting to produce membrane. The membrane was dried in a vacuum oven at 60 °C for 2 h and then subjected to compression molding at 1500 psi for 10 min to form a thin SPE membrane. These samples were denoted as Non-US. The solution after magnetic stirring was sonicated using a high-intensity ultrasonic probe at 50% amplitude for 10 s at pulse mode (Sonics & Materials, Inc, Newtown, CT, USA, VCX750). The effect of ultrasonication was explored with various ultrasonic exposure times of 5, 10, 15, 30, and 60 min. SPE membranes were fabricated using the same route as for the Non-US material as outlined above. The ultrasonication SPE samples are denoted as US. The two fabrication routes are shown in Figure 1.

2.3. Measurements

The ionic conductivity was obtained by impedance spectra measurements using Autolab PGSTAT (302N, Metrohm Autolab Inc., Utrecht, Netherland) in the frequency ranges of 1 Hz–100 kHz at room temperature. The SPE membrane was inserted between two stainless steel (SS) blocking electrodes. An electric field was applied to the cell to cause movement of Li+ ions. The bulk resistance (Rb) of the SPE was calculated from the intercept of the x-axis on a Nyquist plot at high frequency. The ionic conductivity of the SPE was calculated using the Equation (1),
σ = t/Rb·A
where t is the thickness, Rb is the bulk resistance, and A is the area of the electrolyte membrane.
The electrochemical stability windows of SPE were determined by linear sweep voltammetry (LSV) that was performed on SS/SPE/SS cells at room temperature. The LSV was measured from 0 to 5 V at a scan rate of 5 mV s−1. DSC measurements were performed by a DSC1 calorimeter (Mettler Toledo). The sample was subjected to a temperature scanning range from −70 °C to 110 °C at a heating rate of 10 °C/min. Values for glass transition temperature (Tg), melting point (Tm), and % crystallinity were obtained from this experiment. The X-ray diffraction spectra of the SPE membranes were determined using a Rigaku rotating anode goniometer (Rigaku Denki., Ltd., Tokyo, Japan). XRD scans were obtained using Cu Kα radiation (1.541 °A) operating at 40 kV and 100 mA in the 2θ range of 10° to 60° using a step width of 0.02°. FTIR spectra were obtained using an ATR- FTIR Perkin Elmer (Frontier L128-0010) within a range of 450–4000 cm−1 and using a resolution of 4 cm−1. A field-emission scanning electron microscope (FE-SEM, JSM 6335 F) equipped with an energy dispersive X-ray (EDX) analyzer operated at 15 kV was utilized to monitor the form of the HNT dispersion in SPE. Transmission electron microscopy (TEM, TALOS F200X, Thermo Fisher Scientific, Bleiswijk, Netherlands) was used to observe the distribution and the length of HNT particles in the SPE membranes. The tensile properties of the SPE were determined using a Lloyd universal testing machine (model LRX). The samples were tested using a crosshead speed of 50 mm/min at room temperature. The size of each test piece size was 100 mm (length) × 5mm (width) × 0.1–0.2 mm (thickness), with a gauge length of 50 mm. Flame retardant testing was performed on the sample membranes (2 × 3 cm2). The distance between sample membrane and the top of the flame was 2 cm. The time required to start ignition and the time for self-extinguishing after removal of the ignition source were recorded. The dimensional stability of the sample membrane (2 × 3 cm2) was evaluated by observing the dimensional changes after the membrane was held for 5 and 10 h at 60 °C in an electric oven.

3. Results

3.1. Effect of Ultrasonication on the Dispersion of HNTs in SPE

Morphology Study

The incorporation of nanoparticles (NPs) in a polymer matrix is usually limited by the problem of aggregation of the NPs, thus reducing their reinforcing capability. In this work ultrasonication was applied to PEO-20%LiCF3SO3-HNT mixtures to improve the dispersion of the HNTs. The effects of ultrasonic exposure times were investigated at 5, 10, 15, 30, and 60 min and the resultant HNT dispersions compared to those from the conventional magnetic stirrer method. SEM and TEM were used to investigate the morphology of the SPE membranes and the dispersion of HNT in the PEO matrix. The HNTs exhibited a typical nanotube structure (Figure 2a). The nanotubes had an external diameter of 50–100 nm, an inner diameter of 6–30 nm, and are 60–480 nm in length [18]. The tubular nature of the HNT was confirmed by XRD spectra, which show a high-intensity peak at 12.03° (Figure 2b). SEM-EDS was used to inspect the distribution of HNTs in the PEO matrix. Figure 2c shows the distributions of Al, Si, and O atoms which correspond with the alumina and silica in the HNT. For the Non-US treated mixes, the SEM image in Figure 3 shows that the composite with 10 wt.% HNT is more prone to HNT aggregation formation compared with the composite with 5 wt.%. The SEM-EDS mapping of PEO-20%LiCF3SO3-5%HNT shows the relatively uniform distribution of Al (red) and Si (green), indicating that the HNT was dispersed homogenously in the SPE.
Ultrasonication was applied to the PEO-20%LiCF3SO3-5%HNT and the PEO-20%LiCF3SO3-10%HNT samples for various ultrasonic exposure times. The PEO-20%LiCF3SO3-5%HNT sample was studied in more detail to determine the effect of ultrasonication on the membrane properties. Figure 4 presents SEM images and EDS mapping for Si and Al to reveal the distribution of HNT in the PEO matrix for the 5% HNT addition. It was found that after 15 min of ultrasonication, the HNT appears to be well dispersed in the PEO matrix, as confirmed by the TEM image in Figure 5. Ultrasonication for 15 min. prevented formation of aggregates giving a uniform dispersion of nanoparticles in the polymer matrix. However, as illustrated by Figure 5c,d, when the ultrasonic exposure times were increased to 30 and 60 min, the lengths of the HNT tubes were reduced by fragmentation and re-agglomeration of the HNTs was observed in the PEO matrix. Measurements from TEM images revealed that the average length of HNT decreased from 482 ± 215 nm to 275 ± 108, 241 ± 56, and 216 ± 81 nm for the ultrasonic exposure times of 0, 15, 30, and 60 min, respectively. The TEM images (Figure 5) also provided information on the arrangement of the HNTs in the membrane matrix. Without ultrasonication, the HNT tubes were aligned in the matrix as indicated by the red arrows in Figure 5a. Using ultrasonication gave rise to random orientation of HNTs in the membrane as shown in Figure 5b–d. The physical and mechanical properties of polymer nanocomposites strongly depend on the orientation of nanoparticles [19]. It is well known that the reinforcement effect is greater when the fillers are aligned in the direction of strain (one-direction) but is much less in other directions. However, for application as SPE, random orientation is preferred. Mechanical properties in terms of the tensile behavior of the composites produced in this study and the effects of ultrasonication are discussed later in Section 3.7.

3.2. Effect of Ultrasonication on Crystallinity and Thermal Properties of SPE

3.2.1. XRD Analysis

The effect of ultrasonic exposure time (0, 15, 30, and 60 min) on the crystallinity change of PEO is shown by the XRD patterns in Figure 6. The XRD pattern of neat PEO contained two main diffraction peaks at 2θ = 19.14° (1 2 0) and 23.3° (0 3 2), (1 1 2) [20,21,22,23] corresponding to the crystalline phase of the PEO polymer matrix. The addition of the LiCF3SO3 salt reduced the crystallinity of the PEO, as indicated by the significant decrease in the intensity of the prominent peaks at 19.33° and 23.3°. The decrease in the crystallinity of the PEO matrix is due to the complexation between PEO and Li+ ions that inhibits crystallization of the PEO [24]. It was also found that the intensity of the main peaks of PEO decreased upon the incorporation of 5 wt.% HNT. The explanation for this crystallinity decrease is that the HNT may impede the diffusion of the polymer chain to the crystal growth front [25]. When 15 min of ultrasonication was applied to the composite solutions, the diffraction peaks of PEO in the composite became broader and lower in intensity (Figure 6 pattern (d)). The main peak, located at 19.00°, was displaced to a lower angle indicating an increase of d-spacing [26]. However, the second main peak became divided into multi-peaks. When the ultrasonic exposure times increased to 30 and 60 min, the intensity of the peaks tended to increase, which corresponded to increasing crystallinity. It was also found that the prominent peaks were displaced to higher angles, corresponding to decreases in d−spacing.

3.2.2. DSC Analysis

It is well known that Li+ can move in two ways: (a) along the molecular chains of a polymer, and (b) in the amorphous phase of a polymer electrolyte [27]. Therefore, the ionic conductivity of SPE should relate to decreases of Tg and the proportion of crystalline phase of the SPE. Tg and % crystallinity of the SPE were evaluated by DSC. The resultant DSC chromatogram of SPE is shown in Figure 7. The endothermic peak of PEO−20%LiCF3SO3 at 58.2 °C is the melting peak of SPE. The addition of 5%HNT combined with 15 min ultrasonic exposure reduced the area of the melting peak and lowered Tg of the PEO, indicating that the fraction of the crystalline phase in the SPE was decreased. The crystallinity of PEO is calculated by Equation (2)
Xc = ΔHm/(Wi × ΔH°m)
where Xc is crystallinity (%), Wi is the weight fraction of the polymer in the composite, ΔHm is the melting enthalpy of a sample, and ΔH°m is the melting enthalpy of reference for 100% crystallinity of the polymer matrix (PEO = 213.7 J/g) [28]. The Tg of PEO was reduced from −60.7 °C to −67.9 °C by the addition of 5%HNT. Decreases in Tg in the SPE mean that more flexible polymer chains are obtained [29]. The lowest degree of crystallinity of the SPE was found in PEO-20%LiCF3SO3−5%HNT−15 min. This result agreed well with XRD results. The crystalline fraction (Xc) from the XRD pattern can be calculated by Equation (3)
Xc = Ac/(Ac + Aa)
where Ac and Aa are the crystallized and amorphous areas. Decreases of Tg and crystallinity lead to increased ion transportation within the SPE [29]. Higher ionic conductivity is related to the lowest crystallinity, which in this work was found in the PEO−20%LiCF3SO3−5%HNT−15 min composite. When ultrasonication was applied for 30 min and 60 min, the crystallinity tended to increase since the longer time led to segmentation of the polymer chains. The resultant shorter chain lengths were then able to reorient easily to give the higher crystallinity [30]. This consequently restricted the mobility of the polymer chains, thus giving a decrease in ionic conductivity. The values of Tg, Tm, ΔHm, and XC of the various samples are listed in Table 1.

3.3. Effect of Ultrasonication on PEO−LiCF3SO3−HNT Complex

The FT-IR technique has been used to investigate the complexation between Li+ and polymer matrices [31,32]. In SPE, the Li+ would transport upon the application of current into the electrolyte membrane by an ion hopping mechanism [33]. The ether oxygen group (C-O-C) of PEO is strongly electronegative, thus it can readily undergo complexation with Li+ ions [34]. It is noted that the characteristic peak of C-O-C showed a triplet stretching vibration, located at 1147, 1097, 1060 cm−1 as shown in box 4 of Figure 8a. The shift of the C-O-C peaks suggests coordination between Li+ ions and C-O-C bonds in SPE [30]. According to literature, the bonds located at 1032 cm−1, 1042 cm−1, and 1056 cm−1 are assigned to free ions, ion pairs, and ion clusters, respectively [35,36]. FT−IR spectra of PEO−20%LiCF3SO3 displayed peaks from both ion pairs and free ions, which were located at 1041, 1029 cm−1, respectively. The spectra shows that the addition of HNT and the use of ultrasonic treatment caused the disappearance of the peak related to ion pairs, which corresponds with the increase in the free ions peak located at 1041 cm−1. This observation indicates that the addition of HNTs and the use of ultrasonication can dissociate LiCF3SO3 in SPE by reducing the formation of ion pairs. On the other hand, an anion of a lithium salt (CF3SO3, ionic radius is 0.270 nm) may react with aluminal (Al-OH) on the inner surface of HNT (diameter of the lumen is 20–40 nm) thus causing reduced ionic bonding strength between Li+ and CF3SO3- [37,38]. This effect provided more dissociation of lithium salt in the polymer matrix, enhancing the ionic conductivity of the SPE. The schematic illustration of the Li+ conduction mechanism in PEO–LiCF3SO3−HNT is shown in Figure 8b. The shift of the vibration bond is listed in Table 2.

3.4. Effect of Ultrasonication on Ionic Conductivity

The Nyquist plots of all samples show only the spike (Figure 9a), which can be attributed to ionic diffusion. This signifies good ionic conductivity of the SPE membranes. The bulk resistance (Rb) can be obtained from where the plot of the results intercepts the x-axis. The small value of Rb proves that the current easily transfers through the electrode–electrolyte interface due to the high mobility of Li+ [39]. The ionic conductivity was found to increase with the addition of HNTs. The highest ionic conductivity was found at 5 wt.% of HNT. This can be ascribed to increasing lithium salt dissociation due to the interaction between the oxygen atoms of PEO and Li+ ions on the external surface of HNT (−Si-O-Si- group) [6] and the formation of hydrogen bonds between anions (CF3SO3) and -OH groups on the internal surfaces of HNT [40]. However, at 10 and 15 wt.% loadings of HNT into PEO−20%LiCF3SO3 the ionic conductivity decreased. This decrease in ionic conductivity may be caused by a blocking effect where the HNT particles are too close to one another [41]. The effect of HNT content on the ionic conductivity of SPE is presented in Figure 9b. The results also show that the application of ultrasonication gave higher ionic conductivity than by using only the conventional magnetic stirring method. The highest ionic conductivity, which corresponded to the lowest Rb value, was found in the PEO−20%LiCF3SO3−5%HNT sample that was obtained using 15 min of ultrasonication (Figure 9c). This indicates that ultrasonication improved the dispersion of HNTs in the PEO matrix. As a consequence, the Lewis acid–base interaction could readily take place thus enhancing the mobility of ions and increasing the ionic conductivity. When ultrasonic stirring was applied for 30 and 60 min, the ionic conductivity was reduced by about one order of magnitude (Table 3). This reduction was caused by HNT re-agglomeration that hindered the motion of the mobile ions. The same pattern of ionic conductivity results was found when ultrasonication was also applied to the PEO−20%LiCF3SO3−10%HNT system. (see Table 3).

3.5. Effect of Ultrasonication on Electrochemical Stability

An important factor for SPE is the electrochemical stability window (ESW) which refers to safety aspects during application of voltage to a battery [42]. The ESW was determined by linear sweep voltammogram (LVS), using stainless steel (SS) as the working and reference electrodes. The decomposition of SPE was investigated in the potential range of 0–5 V at room temperature, as shown in Figure 9d. The decomposition of the composites without HNT, PEO-20%LiCF3SO3-5%HNT, and as PEO−20%LiCF3SO3−5%HNT−15 min was observed at the 2.96, 3.11, and 3.20 Volts, respectively, indicating that the electrochemical stability of SPE was improved when HNT was present and when ultrasonication was also applied.

3.6. Flame Retardance Testing and Dimensional Stability

The results of flame retardance testing in air of PEO−20%LiCF3SO3−5%HNT and PEO−20%LiCF3SO3−5%HNT−15 US in air are shown in Figure 10. The tests revealed that the PEO−20%LiCF3SO3−5%HNT was ignited at 4s after flame exposure as shown in Figure 10a. The specimen was burned completely at 13 s. Figure 10b shows the flame retardance characteristic of the PEO−20%LiCF3SO3−5%HNT−15 US membrane, which melted at 10 s and then ignited after 30s followed by self-extinguishing at 40 s. These observations indicate that suitably uniform dispersion of HNT in the composite is crucial for flame retardance. PEO−20%LiCF3SO3−5%HNT−15 US showed excellent flame retardancy which was better than that of the PEO−20%LiCF3SO3−5%HNT. Figure 10c shows dimensional changes of the SPE at 60 °C for 0, 5 and 10 h. It was found that the PEO−20%LiCF3SO3−5%HNT showed clear changes in the membrane size, while the PEO−20%LiCF3SO3−5%HNT−15 US membrane showed improved dimensional stability with less change than that of the non-ultrasonically treated membrane.

3.7. Effect of Ultrasonication on Tensile Properties

The tensile stress–strain curves of the SPEs are shown in Figure 11. The tensile strength of PEO−20%LiCF3SO3 increased by about 65% with the addition of 5%HNT. This indicated that the nanotube acted as a reinforcing agent such that the applied load could transfer from the matrix to the nanotube [43]. The tensile strength of PEO−20%LiCF3SO3−5%HNT−US 15 min increased by about 116% when compared to that of PEO−20%LiCF3SO3. Ultrasonication provided improved nanoparticle distribution; hence, load transfer was more effective providing better mechanical properties of the polymer composites [44,45,46]. The onset of plastic deformation was investigated at a 2% offset strain, which is a significant property measure when using materials for engineering applications. The 2% offset of a strain of PEO−20%LiCF3SO3−5%HNT−15 min showed the highest value of all the prepared composites in this study as 1.53 MPa. The tensile properties of SPE are listed in Table 4.

4. Conclusions

PEO−20%LiCF3SO3−HNT composite membranes were prepared by solution casting, ultrasonication and compression molding as SPE for potential application in lithium-ion batteries. It was found that ultrasonication could disperse the HNT nanoparticles in the sample solutions effectively thus reducing the crystallinity of the PEO. A treatment time of 15 min provided membrane with high ionic conductivity, good electrochemical stability, a 2% higher offset strain stress and higher tensile strength. Ultrasonication exposure time affected the HNT length and the properties of the SPE. PEO−20%LiCF3SO3−5%HNT subjected to 15 min of ultrasonication was found to be the optimal condition for SPE preparation, providing a SPE membrane with not only the high tensile properties but also highest ionic conductivity of 1.1 ± 0.19 × 10−4 S/cm. This SPE membrane could be stable up to 3.2 Volts and also showed excellent flame retardancy and dimensional stability. The uniform dispersion of HNT during preparation was found to be a key factor in producing composite membrane with optimum properties. Composite membrane is considered to be a suitable candidate for possible use as SPE in a solid-state Li-ion battery.

Author Contributions

Conceptualization, methodology, investigation, writing—original draft, P.P.; conceptualization, writing—review & editing, supervision, project administration, J.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Program Management Unit for Human Resources & Institutional Development, Research and Innovation, Office of National Higher Education Science Research and Innovation Policy Council (NXPO) [Grant Number B16F640001], Fundamental Fund 2023, Chiang Mai University, and Chiang Mai University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available upon request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kim, H.W.; Manikandan, P.; Lim, Y.J.; Kim, J.H.; Nam, S.; Kim, Y. Hybrid solid electrolyte with the combination of Li7La3Zr2O12 ceramic and ionic liquid for high voltage pseudo-solid-state Li-ion batteries. J. Mater. Chem. A 2016, 4, 17025–17032. [Google Scholar] [CrossRef]
  2. Zhang, Y.; Lu, W.; Cong, L.; Liu, J.; Sun, L.; Mauger, A.; Julien, C.M.; Xie, H.; Liu, J. Cross-linking network based on Poly(ethylene oxide): Solid polymer electrolyte for room temperature lithium battery. J. Power Sources 2019, 420, 63–72. [Google Scholar] [CrossRef]
  3. Choi, J.-H.; Lee, C.-H.; Yu, J.-H.; Doh, C.-H.; Lee, S.-M. Enhancement of ionic conductivity of composite membranes for all-solid-state lithium rechargeable batteries incorporating tetragonal Li7La3Zr2O12 into a polyethylene oxide matrix. J. Power Sources 2015, 274, 458–463. [Google Scholar] [CrossRef]
  4. Liu, Y.-H.; Chen, K.T.; Liu, B.H. Fabrication of composite probe electrode used for localized impedance analysis of solid-state electrolyte LATP. Solid State Ion. 2019, 336, 11–18. [Google Scholar] [CrossRef]
  5. Zhao, Y.; Wang, L.; Zhou, Y.; Liang, Z.; Tavajohi, N.; Li, B.; Li, T. Solid Polymer Electrolytes with High Conductivity and Transference Number of Li Ions for Li-Based Rechargeable Batteries. Adv. Sci. 2021, 8, 2003675. [Google Scholar] [CrossRef] [PubMed]
  6. Fu, X.; Yu, D.; Zhou, J.; Li, S.; Gao, X.; Han, Y.; Qi, P.; Feng, X.; Wang, B. Inorganic and organic hybrid solid electrolytes for lithium-ion batteries. CrystEngComm 2016, 18, 4236–4258. [Google Scholar] [CrossRef]
  7. Khurana, R.; Schaefer, J.L.; Archer, L.A.; Coates, G.W. Suppression of Lithium Dendrite Growth Using Cross-Linked Polyethylene/Poly(ethylene oxide) Electrolytes: A New Approach for Practical Lithium-Metal Polymer Batteries. J. Am. Chem. Soc. 2014, 136, 7395–7402. [Google Scholar] [CrossRef]
  8. Dam, T.; Karan, N.K.; Thomas, R.; Pradhan, D.K.; Katiyar, R.S. Observation of ionic transport and ion-coordinated segmental motions in composite (polymer-salt-clay) solid polymer electrolyte. Ionics 2015, 21, 401–410. [Google Scholar] [CrossRef]
  9. Mohanta, J.; Singh, U.P.; Panda, S.K.; Si, S. Enhancement of Li+ ion conductivity in solid polymer electrolytes using surface tailored porous silica nanofillers. Adv. Nat. Sci. Nanosci. Nanotechnol. 2016, 7, 035011. [Google Scholar] [CrossRef]
  10. Jia, W.; Li, Z.; Wu, Z.; Wang, L.; Wu, B.; Wang, Y.; Cao, Y.; Li, J. Graphene oxide as a filler to improve the performance of PAN-LiClO4 flexible solid polymer electrolyte. Solid State Ion. 2018, 315, 7–13. [Google Scholar] [CrossRef]
  11. Dean, K.; Krstina, J.; Tian, W.; Varley, R.J. Effect of Ultrasonic Dispersion Methods on Thermal and Mechanical Properties of Organoclay Epoxy Nanocomposites. Macro. Mol. Mater. Eng. 2007, 292, 415–427. [Google Scholar] [CrossRef]
  12. Parveen, S.; Rana, S.; Ferreira, S.; Filho, A.; Fangueiro, R. Ultrasonic dispersion of micro crystalline cellulose for developing cementitious composites with excellent strength and stiffness. Ind. Crops Prod. 2018, 122, 156–165. [Google Scholar] [CrossRef]
  13. Yuhana, N.Y.; Ahmad, S.; Bahri, A.R.S. The Effect of Ultrasonic Treatment on Thermal Stability of the Cured Epoxy/Layered Silicate Nano composite. Adv. Mater. Sci. Eng. 2012, 2012, 789815. [Google Scholar] [CrossRef]
  14. Jokar, A.; Azizi, M.H.; Esfehani, Z.H.; Abbasi, S. Effects of ultrasound time on the properties of methylcellulose-montmorillonite films. Int. Nano Lett. 2017, 7, 59–68. [Google Scholar] [CrossRef]
  15. Lun, H.; Ouyang, J.; Yang, H. Enhancing dispersion of halloysite nanotubes via chemical modification. Phys. Chem. Miner. 2014, 41, 281–288. [Google Scholar] [CrossRef]
  16. Pongsuk, P.; Pumchusak, J. Effects of Natural Clay on Ionic Conductivity, Crystallinity and Thermal Properties of PEO-LiCF3SO3-Natural Clay as Solid Polymer Electrolyte Nanocomposites. Key Eng. Mater. 2020, 803, 98–103. [Google Scholar] [CrossRef]
  17. Zhou, X.; Yin, Y.; Wang, Z.; Zhou, J.; Huang, H.; Mansour, A.N.; Zaykoski, J.A.; Fedderly, J.J.; Balizer, E. Effect of hot pressing on the ionic conductivity of the PEO/LiCF3SO3 based electrolyte membranes. Solid State Ion. 2011, 196, 18–24. [Google Scholar] [CrossRef]
  18. Shao, L.; Wang, X.; Yang, B.; Wang, Q.; Tian, Q.; Ji, Z.; Zhang, J. A Highly Sensitive Ascorbic Acid Sensor Based on Hierarchical Polyaniline Coated Halloysite Nanotubes Prepared by Electrophoretic Deposition. Electrochim. Acta 2017, 255, 286–297. [Google Scholar] [CrossRef]
  19. Weon, J.-I.; Xia, Z.Y.; Sue, H.J. Morphological characterization of nylon-6 nanocomposite following a large-scale simple shear process. J. Polym. Sci. Part B Polym. Phys. 2005, 43, 3555–3566. [Google Scholar] [CrossRef]
  20. Chrissopoulou, K.; Andrikopoulos, K.S.; Fotiadou, S.; Bollas, S.; Karageorgaki, C.; Christofilos, D.; Voyiatzis, G.A.; Anastasiadis, S.H. Crystallinity and Chain Conformation in PEO/Layered Silicate Nanocomposites. Macromolecules 2011, 44, 9710–9722. [Google Scholar] [CrossRef]
  21. Elashmawi, I.S.; Elsayed, N.H.; Altalhi, F.A. The changes of spectroscopic, thermal and electrical properties of PVDF/PEO containing lithium nanoparticles. J. Alloys Compd. 2014, 617, 877–883. [Google Scholar] [CrossRef]
  22. Polu, A.R.; Rhee, H.-W. Nanocomposite solid polymer electrolytes based on poly(ethylene oxide)/POSS-PEG (n = 13.3) hybrid nanoparticles for lithium ion batteries. J. Ind. Eng. Chem. 2015, 31, 323–329. [Google Scholar] [CrossRef]
  23. Talaat, H.; Turky, G.; Abdelghany, A.M.; Fathia, M. Structure-dynamic properties relationships in poly(ethylene oxide)/silicon dioxide nanocomposites:dielectric relaxation study. Polym. Bull. 2021, 78, 5205–5223. [Google Scholar]
  24. Garcia-Garcia, D.; Ferri, J.M.; Ripoll, L.; Hidalgo, M.; Lopez-Martinez, J.; Balart, R. Characterization of selectively etched halloysite nanotubes by acid treatment. Appl. Surf. Sci. 2017, 422, 616–625. [Google Scholar] [CrossRef]
  25. Abreha, M.; Subrahmanyam, A.R.; Kumar, J.S. Ionic conductivity and transport properties of poly(vinylidene fluoride-co-hexafluoropropylene)-based solid polymer electrolytes. Chem. Phys. Lett. 2016, 658, 240–247. [Google Scholar] [CrossRef]
  26. Kiliaris, P.; Papaspyrides, C.D. Polymer/layered silicate (clay) nanocomposites: An overview of flame retardancy. Prog. Polym. Sci. 2010, 35, 902–958. [Google Scholar] [CrossRef]
  27. Angulakshmi, N.; Thomas, S.; Nahm, K.S.; Stephan, A.M.; Elizabeth, R.N. Electrochemical and mechanical properties of nanochitin-incorporated PVDF-HFP-based polymer electrolytes for lithium batteries. Ionics 2011, 17, 407–414. [Google Scholar] [CrossRef]
  28. Klongkan, S.; Pumchusak, J. Effects of Nano Alumina and Plasticizers on Morphology, Ionic Conductivity, Thermal and Mechanical Properties of PEO-LiCF3SO3 Solid Polymer Electrolyte. Electrochim. Acta 2015, 161, 171–176. [Google Scholar] [CrossRef]
  29. Jang, H.S.; Raj, C.J.; Lee, W.-G.; Kim, B.C.; Yu, K.H. Enhanced supercapacitive performances of functionalized activated carbon in novel gel polymer electrolytes with ionic liquid redox-mediated poly(vinyl alcohol)/phosphoric acid. RSC Adv. 2016, 6, 75376–75383. [Google Scholar] [CrossRef]
  30. Pu, Y.; Zou, Q.; Hou, D.; Zhang, Y.; Chen, S. Molecular weight kinetics and chain scission models for dextran polymers during ultrasonic degradation. Carbohydr. Polym. 2017, 156, 71–76. [Google Scholar] [CrossRef]
  31. Sim, L.H.; Gan, S.N.; Chan, C.H.; Yahya, R. ATR-FTIR studies on ion interaction of lithium perchlorate in polyacrylate/poly(ethylene oxide) blends, Spectrochim. Acta A Mol. Biomol. Spectrosc. 2010, 76, 287–292. [Google Scholar] [CrossRef] [PubMed]
  32. Selvasekarapandian, S.; Baskaran, R.; Kamishima, O.; Kawamura, J.; Hattori, T. Laser Raman and FTIR studies on Li+ interaction in PVAc–LiClO4 polymer electrolytes. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2006, 65, 1234–1240. [Google Scholar] [CrossRef] [PubMed]
  33. Isa, K.B.M.; Osman, Z.; Arof, A.K.; Othman, L.; Zainol, N.H.; Samin, S.M.; Chong, W.G.; Kamarulzaman, N. Lithium ion conduction and ion–polymer interaction in PVdF-HFP based gel polymer electrolytes. Solid State Ion. 2014, 268, 288–293. [Google Scholar] [CrossRef]
  34. Frech, R.; Chintapalli, S.; Bruce, P.G.; Vincent, C.A. Crystalline and Amorphous Phases in the Poly(ethylene oxide)−LiCF3SO3 System. Macromolecules 1999, 32, 808–813. [Google Scholar] [CrossRef]
  35. Winie, T.; Muhammad, F.; Hazwani, N. Effect of Anion Size on the Conductivity Behaviour of Hexanoyl Chitosan-Based Polymer Electrolytes. Adv. Mat. Res. 2012, 545, 317–320. [Google Scholar] [CrossRef]
  36. Vignarooban, K.; Dissanayake, M.A.K.L.; Albinsson, I.; Mellander, B.-E. Effect of TiO2 nano-filler and EC plasticizer on electrical and thermal properties of poly(ethylene oxide) (PEO) based solid polymer electrolytes. Solid State Ion. 2014, 266, 25–28. [Google Scholar] [CrossRef]
  37. Suait, M.S.; Ahmad, A.; Hamzah, H.; Rahman, M.Y.A. Effect of lithium salt concentrations on blended 49% poly(methyl methacrylate) grafted natural rubber and poly(methyl methacrylate) based solid polymer electrolyte. Electrochim. Acta 2011, 57, 123–131. [Google Scholar] [CrossRef]
  38. Polu, A.R.; Rhee, H.-W. Effect of TiO2 nanoparticles on structural, thermal, mechanical and ionic conductivity studies of PEO12–LiTDI solid polymer electrolyte. J. Ind. Eng. Chem. 2016, 37, 347–353. [Google Scholar] [CrossRef]
  39. Barzegar, F.; Dangbegnon, J.K.; Bello, A.; Momodu, D.Y.; Johnson, A.T.C.; Manyala, N. Effect of conductive additives to gel electrolytes on activated carbon-based supercapacitors. AIP Adv. 2015, 5, 097171. [Google Scholar] [CrossRef]
  40. Lin, Y.; Wang, X.; Liu, J.; Miller, J.D. Natural halloysite nano-clay electrolyte for advanced all-solid-state lithium-sulfur batteries. Nano Energy 2017, 31, 478–485. [Google Scholar] [CrossRef]
  41. Chew, K.W.; Tan, K.W. The Effects of Ceramic Fillers on PMMA-Based Polymer Electrolyte Salted with Lithium Triflate, LiCF3SO3. Int. J. Electrochem. Sci. 2011, 6, 5792–5801. [Google Scholar]
  42. Jinisha, B.; Anilkumar, K.M.; Manoj, M.; Pradeep, V.S.; Jayalekshmi, S. Development of a novel type of solid polymer electrolyte for solid state lithium battery applications based on lithium enriched poly (ethylene oxide) (PEO)/poly (vinyl pyrrolidone) (PVP) blend polymer. Electrochim. Acta 2017, 235, 210–222. [Google Scholar]
  43. Deng, D. Li-ion batteries: Basics, progress, and challenges. Energy Sci. Eng. 2015, 3, 385–418. [Google Scholar] [CrossRef]
  44. Pavlinakova, V.; Fohlerova, Z.; Pavlinak, D.; Khunova, V.; Vojtova, L. Effect of halloysite nanotube structure on physical, chemical, structural and biological properties of elastic polycaprolactone/gelatin nanofibers for wound healing applications. Mater. Sci. Eng. C 2018, 91, 94–102. [Google Scholar] [CrossRef] [PubMed]
  45. Pignanelli, F.; Romero, M.; Faccio, R.; Fernández-Werner, L.; Mombru, A.W. Enhancement of Lithium-Ion Transport in Poly(acrylonitrile) with Hydrogen Titanate Nanotube Fillers as Solid Polymer Electrolytes for Lithium-Ion Battery Applications. J. Phys. Chem. C 2018, 122, 1492–1499. [Google Scholar] [CrossRef]
  46. Carli, L.N.; Crespo, J.S.; Mauler, R.S. PHBV nanocomposites based on organomodified montmorillonite and halloysite: The effect of clay type on the morphology and thermal and mechanical properties. Compos. Part A Appl. Sci. Manuf. 2011, 42, 1601–1608. [Google Scholar] [CrossRef]
Figure 1. Schematic outline of the two fabrication routes for solid polymer electrolyte nanocomposite.
Figure 1. Schematic outline of the two fabrication routes for solid polymer electrolyte nanocomposite.
Polymers 14 03710 g001
Figure 2. FE−SEM image of (a) As−supplied HNT, (b) XRD pattern and crystalline structure of HNT and (c) SEM image and EDS spectrum of HNT in SPE.
Figure 2. FE−SEM image of (a) As−supplied HNT, (b) XRD pattern and crystalline structure of HNT and (c) SEM image and EDS spectrum of HNT in SPE.
Polymers 14 03710 g002
Figure 3. SEM micrographs of fracture surfaces of the Non−US prepared PEO-20%LiCF3SO3 with (a) 5%HNT, (b) 10%HNT at ×5000, ×10,000 and EDS mapping of Al (red), Si (green), respectively.
Figure 3. SEM micrographs of fracture surfaces of the Non−US prepared PEO-20%LiCF3SO3 with (a) 5%HNT, (b) 10%HNT at ×5000, ×10,000 and EDS mapping of Al (red), Si (green), respectively.
Polymers 14 03710 g003
Figure 4. FE−SEM−EDS mapping at ×5000 (scale bar = 10 μm) of silicon (Si) atoms (green) and aluminum (Al) atoms (red) to indicate HNT distribution in PEO−20%LiCF3SO3−5%HNT of (a) Non−US, (b) US 5 min, (c) US 10 min, (d) US 15 min, (e) US 30 min and (f) US 60 min.
Figure 4. FE−SEM−EDS mapping at ×5000 (scale bar = 10 μm) of silicon (Si) atoms (green) and aluminum (Al) atoms (red) to indicate HNT distribution in PEO−20%LiCF3SO3−5%HNT of (a) Non−US, (b) US 5 min, (c) US 10 min, (d) US 15 min, (e) US 30 min and (f) US 60 min.
Polymers 14 03710 g004
Figure 5. TEM images of PEO−20%LiCF3SO3−5%HNT (a) Non−US, (b) US 15 min, (c) US 30 min and (d) US 60 min.
Figure 5. TEM images of PEO−20%LiCF3SO3−5%HNT (a) Non−US, (b) US 15 min, (c) US 30 min and (d) US 60 min.
Polymers 14 03710 g005
Figure 6. XRD patterns of (a) As−received PEO, (b) PEO−20%LiCF3SO3, (c) PEO−20%LiCF3SO3−5%HNT, (d) PEO−20%LiCF3SO3−5%HNT−15 min, (e) PEO−20%LiCF3SO3−5%HNT−30 min and (f) PEO−20%LiCF3SO3−5%HNT−60 min.
Figure 6. XRD patterns of (a) As−received PEO, (b) PEO−20%LiCF3SO3, (c) PEO−20%LiCF3SO3−5%HNT, (d) PEO−20%LiCF3SO3−5%HNT−15 min, (e) PEO−20%LiCF3SO3−5%HNT−30 min and (f) PEO−20%LiCF3SO3−5%HNT−60 min.
Polymers 14 03710 g006
Figure 7. DSC curves of PEO−20%LiCF3SO3 and PEO−20%LiCF3SO3−5%HNT with and without ultrasonication.
Figure 7. DSC curves of PEO−20%LiCF3SO3 and PEO−20%LiCF3SO3−5%HNT with and without ultrasonication.
Polymers 14 03710 g007
Figure 8. (a) ATR−FTIR spectra in the region of 450–4000 cm−1 of (a) PEO−20%LiCF3SO3, (b) PEO−20%LiCF3SO3−5%HNT, (c) PEO−20%LiCF3SO3−5%HNT−15 min, (d) PEO−20%LiCF3SO3−5%HNT−30 min, and (e) PEO−20%LiCF3SO3−5%HNT−60 min. (b) Schematic illustration of Li+ conduction mechanism in PEO−LiCF3SO3−HNT (left) and the sample of solid polymer electrolyte (right).
Figure 8. (a) ATR−FTIR spectra in the region of 450–4000 cm−1 of (a) PEO−20%LiCF3SO3, (b) PEO−20%LiCF3SO3−5%HNT, (c) PEO−20%LiCF3SO3−5%HNT−15 min, (d) PEO−20%LiCF3SO3−5%HNT−30 min, and (e) PEO−20%LiCF3SO3−5%HNT−60 min. (b) Schematic illustration of Li+ conduction mechanism in PEO−LiCF3SO3−HNT (left) and the sample of solid polymer electrolyte (right).
Polymers 14 03710 g008
Figure 9. (a) Nyquist plot for Non−US of PEO−20%LiCF3SO3 at room temperature, (b) the ionic conductivity for Non−US of PEO−20%LiCF3SO3−y%HNT, (c) Nyquist plot of PEO−20%LiCF3SO3−5%HNT with various ultrasonic exposure time and (d) Linear sweep voltammogram.
Figure 9. (a) Nyquist plot for Non−US of PEO−20%LiCF3SO3 at room temperature, (b) the ionic conductivity for Non−US of PEO−20%LiCF3SO3−y%HNT, (c) Nyquist plot of PEO−20%LiCF3SO3−5%HNT with various ultrasonic exposure time and (d) Linear sweep voltammogram.
Polymers 14 03710 g009
Figure 10. Flammability test of (a) PEO−20%LiCF3SO3−5%HNT, (b) PEO−20%LiCF3SO3−5%HNT−15 US. (c) dimensional stability at 60 °C of (1), (3), (5) for PEO−20%LiCF3SO3−5%HNT at 0 h, 5 h and 10 h, and (2), (4), (6) for PEO−20%LiCF3SO3−5%HNT−15 US at 0 h, 5 h, 10h.
Figure 10. Flammability test of (a) PEO−20%LiCF3SO3−5%HNT, (b) PEO−20%LiCF3SO3−5%HNT−15 US. (c) dimensional stability at 60 °C of (1), (3), (5) for PEO−20%LiCF3SO3−5%HNT at 0 h, 5 h and 10 h, and (2), (4), (6) for PEO−20%LiCF3SO3−5%HNT−15 US at 0 h, 5 h, 10h.
Polymers 14 03710 g010
Figure 11. Stress–strain curves of solid polymer electrolytes at room temperature.
Figure 11. Stress–strain curves of solid polymer electrolytes at room temperature.
Polymers 14 03710 g011
Table 1. Tg, Tm, ΔHm, and degree of crystallinity of SPE samples.
Table 1. Tg, Tm, ΔHm, and degree of crystallinity of SPE samples.
SamplesTg end point
(°C)
Tm
(°C)
ΔHm
(J/g)
Degree of Crystallinity (%)
DSCXRD
PEO−20%LiCF3SO3−60.758.267.3231.5035.25
PEO−20%LiCF3SO3−5%HNT−67.958.454.0923.4130.26
PEO−20%LiCF3SO3−5%HNT−15 min−67.758.852.5422.7426.24
PEO−20%LiCF3SO3−5%HNT−30 min−67.858.768.3829.6035.58
PEO−20%LiCF3SO3−5%HNT−60 min−60.162.469.3430.0136.11
Table 2. The assignment of each wavenumber of FT-IR spectrum in Figure 8.
Table 2. The assignment of each wavenumber of FT-IR spectrum in Figure 8.
SamplesWavenumber (cm−1)
Box1Box2Box3Box4Box5 Box6Box7Box8Box9
CF3
Stretch
CH2 RockCH2 RockC-O-CCH2 TwistCH2 WagCH2 ScissorCH2 StretchOH
Stretch
a639841960, 9481147, 1097, 10601279, 1259, 12411359, 13411467, 145428823494
b639841961, 9461147, 1097, 10601279, 1259, 12421359, 13421467, 145528843487
c641844963, 9491148, 1098, 10621281, 1260, 12441361, 13431469, 145728853493
d638842960, 9471146, 1096, 10611278, 1255, 12421359, 13411467, 145428813487
e638842960, 9461144, 1095, 10611278, 1253, 12661359, 13421467, 145528823487
Table 3. The ionic conductivity at room temperature of solid polymer electrolytes.
Table 3. The ionic conductivity at room temperature of solid polymer electrolytes.
Solid Polymer ElectrolytesUltrasonication Time (min)Ionic Conductivity (S/cm)
PEO−20%LiCF3SO3−5%HNT06.6 ± 1.33 × 10−6
51.3 ± 0.83 × 10−5
101.5 ± 0.03 × 10−5
151.1 ± 0.19 × 10−4
301.7 ± 0.84 × 10−5
601.5 ± 1.24 × 10−5
PEO−20%LiCF3SO3−10%HNT01.4 ± 0.17 × 10−6
155.1 ± 1.15 × 10−5
303.6 ± 1.15 × 10−5
Table 4. Tensile properties of SPEs.
Table 4. Tensile properties of SPEs.
SamplesTensile Strength
(MPa)
%Strain at
Break
Stress @ 2% Offset of Strain
PEO-20%LiCF3SO31.47 ± 0.99793.54 ± 3.460.43
PEO-20%LiCF3SO3-5%HNT2.43 ± 0.44525.01 ± 44.721.04
PEO-20%LiCF3SO3-5%HNT-15 min3.18 ± 0.06433.34 ± 57.741.53
PEO-20%LiCF3SO3-5%HNT-30 min0.71 ± 0.12799.98 ± 0.210.40
PEO-20%LiCF3SO3-5%HNT-60 min1.02 ± 0.22791.87 ± 4.330.87
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pongsuk, P.; Pumchusak, J. Effect of Ultrasonication on the Morphology, Mechanical Property, Ionic Conductivity, and Flame Retardancy of PEO-LiCF3SO3-Halloysite Nanotube Composites for Use as Solid Polymer Electrolyte. Polymers 2022, 14, 3710. https://doi.org/10.3390/polym14183710

AMA Style

Pongsuk P, Pumchusak J. Effect of Ultrasonication on the Morphology, Mechanical Property, Ionic Conductivity, and Flame Retardancy of PEO-LiCF3SO3-Halloysite Nanotube Composites for Use as Solid Polymer Electrolyte. Polymers. 2022; 14(18):3710. https://doi.org/10.3390/polym14183710

Chicago/Turabian Style

Pongsuk, Pattranuch, and Jantrawan Pumchusak. 2022. "Effect of Ultrasonication on the Morphology, Mechanical Property, Ionic Conductivity, and Flame Retardancy of PEO-LiCF3SO3-Halloysite Nanotube Composites for Use as Solid Polymer Electrolyte" Polymers 14, no. 18: 3710. https://doi.org/10.3390/polym14183710

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