Next Article in Journal
Modeling the Solubility of Phenolic Acids in Aqueous Media at 37 °C
Next Article in Special Issue
A Novel Cathode Material Synthesis and Thermal Characterization of (1-x-y) LiCo1/3Ti1/3Fe1/3PO4, xLi2MnPO4, yLiFePO4 Composites for Lithium-Ion Batteries (LIBs)
Previous Article in Journal
Antioxidant and Antimicrobial Evaluation and Chemical Investigation of Rosa gallica var. aegyptiaca Leaf Extracts
Previous Article in Special Issue
A Hierarchically Ordered Mesoporous-Carbon-Supported Iron Sulfide Anode for High-Rate Na-Ion Storage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Development on Solid Polymer Electrolytes for Electrochemical Devices

Centre for Ionics University of Malaya, Physics Department, Faculty of Science, University of Malaya, Kuala Lumpur 50603, Malaysia
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(21), 6499; https://doi.org/10.3390/molecules26216499
Submission received: 10 September 2021 / Revised: 22 October 2021 / Accepted: 25 October 2021 / Published: 28 October 2021
(This article belongs to the Special Issue Materials for Emerging Electrochemical Devices)

Abstract

:
Electrochemical devices, especially energy storage, have been around for many decades. Liquid electrolytes (LEs), which are known for their volatility and flammability, are mostly used in the fabrication of the devices. Dye-sensitized solar cells (DSSCs) and quantum dot sensitized solar cells (QDSSCs) are also using electrochemical reaction to operate. Following the demand for green and safer energy sources to replace fossil energy, this has raised the research interest in solid-state electrochemical devices. Solid polymer electrolytes (SPEs) are among the candidates to replace the LEs. Hence, understanding the mechanism of ions’ transport in SPEs is crucial to achieve similar, if not better, performance to that of LEs. In this paper, the development of SPE from basic construction to electrolyte optimization, which includes polymer blending and adding various types of additives, such as plasticizers and fillers, is discussed.

1. Introduction

In any electrochemical devices, the electrolyte is viewed as the core constituent separating the cathode and anode, and it serves as the medium for charge transport. The electrolytes in electrochemical cells exist in three forms, viz liquid, gel and solid. Liquid electrolytes (LEs) are usually found to be highly conducting; thus, they are favored in all kinds of electrochemical devices, namely lithium-ion batteries, dye-sensitized solar cells (DSSCs), quantum dot sensitized solar cells (QDSSCs), fuel cells, electrochemical double layer capacitors (EDLCs), electrochromic devices (ECDs) and so on. However, LEs are sometimes volatile and vulnerable to evaporation, leakage and corrosion in the long-term. This will lead to stability and safety issues of the electrochemical devices. Moreover, LEs can also cause the formation of solid electrolyte interphase (SEI) at the anode surface not only in Li-ion batteries but also for all alkali-metal-ion batteries. This can bring adverse effects on their performance. On the other hand, gel-type electrolytes that inherit both characteristics from electrolytes in liquid and solid forms can reduce complications due to safety, but they suffer from structural instability and low mechanical property. Hence, electrolytes in the solid state are the better and more trustworthy choice to resolve the safety problems without compromising the structural and mechanical traits.
Solid polymer electrolyte (SPE), which is fundamentally made up from dissolved conducting salt within the polymer host, was first discovered by three polymer chemists from Sheffield, UK, i.e., Fenton, Parker and Wright, about 48 years ago [1]. Then, in 1978, Armand and co-workers [2] envisioned the promising implementation of SPE in solid-state batteries. Since then, research on SPE has improved tremendously. In order for SPE to be utilized in electrochemical devices, it must have reasonably good ionic conductivity, good chemical, and thermal and mechanical stabilities, as well as good interfacial contact with electrodes. SPE is considered safe, due to the absence of volatile and flammable solvents, as well as flexible, since it can accommodate volume changes in electrodes during charging/discharging reactions of electrochemical cells (namely in Li-ion batteries). Furthermore, SPE can be prepared in various shapes and desired designs to meet the needs of electrochemical devices. It can be prepared as an ultra-thin membrane which is lightweight, thus offering promising high energy and power density. SPE is unique compared to liquid electrolyte in the sense that the polymer matrix is in charge for the solvation of ions, while the ions are located at sites which are flexible and can travel within the polymer segments. In order to be functional in electrochemical devices, the SPE must fulfil some prerequisites [3]:
  • Good ambient ionic conductivity of the order between 10−4 and 10−2 S cm−1.
  • Wide electrochemical stability window.
  • Mechanically stable in order to be safe and durable for devices operation
  • Thermally stable when in contact with electrode components.
  • Chemical and electrochemical compatibility with electrodes.
  • Obtainability—the raw materials should be abundantly found and low-priced.
A basic SPE contains a polymer and a salt. The salt is dispersed in the polymer matrix, and the ions can conduct through the polymer chains. In order for the polymer to function as host in polymer electrolytes, the macromolecules must possess polar groups which have lone-pair electrons for dative bonding with the cation from the salt. Both synthetic and natural polymers can serve as polymer matrices in SPEs as long as they have functional group(s) to fulfil the complexation criterion. Various polymer electrolytes have been developed from polymers, such as poly(ethylene oxide) (PEO) [1,4,5], cellulose [6], polyvinylidene fluoride (PVDF) [7], poly(vinyl alcohol) (PVA) [8], polyacrylonitrile (PAN) [9], chitosan [10], poly(methyl methacrylate) (PMMA) [11], starch [12] and poly(vinylidene fluoride-hexafluoropropylene) (PVDF-HFP) [13], to name a few. This review presents some achievements on both the synthetic polymers and biopolymers as polymer host in SPEs used, along with working and counter electrodes in various electrochemical devices namely lithium-ion batteries, sodium-ion batteries, DSSCs, QDSSCs, fuel cells, supercapacitors, EDLCs, ECDs, proton-conducting batteries and proton-exchange membrane fuel cells (PEMFCs). Here, we pay special attention to the conductivity aspect of SPEs in these electrochemical devices. Different types of approaches taken by researchers to improve the ionic conductivity of SPEs are also discussed.

2. Ion Transport Properties and Conduction Mechanism in SPEs

Ionic conductivity is one vital characteristic in determining whether the SPE is practical and efficient for device applications. The ionic conductivity of an electrolyte is calculated by using the bulk resistance from the Nyquist impedance plot. The electrolyte is subjected to a small potential typically 10 mV to ensure a linear current–voltage relationship for impedance measurement. According to the Nernst–Einstein relationship, the ionic conductivity, σ, is directly proportional to the diffusion coefficient of ions, D, as shown in the following equation:
σ = q 2 n k b T D
where q is elementary charge, n is the number of ions, kb is Boltzmann constant and T is temperature. Since the electrolyte consists of salts that can ionize into cations and anions, Equation (1) can be written as follows:
σ = σ + + σ = q 2 k b T   ( n + D + + n D )
where σ + and σ are the ionic conductivity of cations and anions respectively, n + is the number of cations, n is the number of anions, D + is the cationic diffusion coefficient and D is the anionic diffusion coefficient. The movement of ions are varied in the electrolyte systems depending on the solvation environment (liquid, solid and gel).
As aforementioned, a polymer electrolyte generally consists of inorganic salts solvated in a polymer matrix that contains functional groups, such as -OH, -C=O, -NH2 and -COC-. The functional groups help to solvate salts into ions via electrostatic interaction. It has been acknowledged for the past three decades that the motion of ions in polymer matrix is coupled with the segmental movement of the polymer chains, and this is described as the Brownian motion by many researchers [14]. Therefore, it is believed that the ion movements can only occur in amorphous polymers where the polymer chains are freely moveable to transport the ions. It is well-known that the motion and flexibility of polymer chains are higher in polymer matrix that possess low glass transition temperature, Tg. Hence, the selection of polymer is crucial in the development of polymer electrolyte. The relation between conductivity and Tg can be explained by the Vogel-Tammann-Fulcher (VTF) equation:
σ = σ o T 1 / 2 e x p ( B T ( T g 50 K ) )
where σ o is the prefactor and B is related to the activation energy. Tg can also be determined from differential scanning calorimetry (DSC). Based on Equations (2) and (3), the conductivity of a polymer electrolyte can be amplified by increasing the number density of mobile ions and lowering the Tg. Another approach is by decoupling ion diffusion with the polymer chains. A simplest way to fulfil the above criteria is by adding high dielectric constant plasticizers into the polymer electrolyte. A high dielectric constant can weaken the attraction forces between the ions in the salt and thus enhances the salt dissociation and the increase in number density of mobile ions [3]. Plasticizers also facilitate the pathway of ions transportation by increasing the amorphousness of the polymer electrolyte film and providing the additional coordination sites for ions to conduct.
Whu and Wick [4] have investigated the role of propylene carbonate (PC) as plasticizer in PEO-based electrolytes having lithium bis(trifluoromethanesulfonyl)imide (LiTFSI) via Monte Carlo and molecular dynamics simulations. They found that the conductivity showed significant improvement with the addition of PC due to the increase in mobility of TFSI- ions [4]. The lithium ions’ diffusion remained unchanged and showed stronger interaction with the ethylene oxide oxygen as compared to the carbonate oxygen. Radial distribution function of lithium also showed weak binding between lithium and oxygen atom of TFSI- in the presence of PC [4]. Therefore, the probability of lithium ions’ transportation through the oxygen atom of PEO was higher than that of carbonate and TFSI- oxygen [4].
Webb and co-workers [5] calculated the diffusion of lithium ions in different types of polymers, namely PEO, poly(methylene oxide) (PMO), poly(propylene oxide) (PPO), poly(trimethylene oxide) (PTMO), poly(ethylene oxide-alt-methylene oxide) (P(EO-MO)) and poly(ethylene oxide-alt-trimethylene oxide) (P(EO-TMO), using molecular dynamics and dynamic bond percolation models. They have assigned short molecular dynamics trajectories to identify the distribution of the solvation sites and used it in dynamic bond percolation model to calculate the ionic hopping rate. According to them, the largest number of site density belongs to the polymer with the highest oxygen atom content which is PMO (oxygen to carbon atomic ratio is 1:1). This is followed by P(EO-MO) (2:3), PEO (1:2), P(EO-TMO) (2:5), PPO (1:3) and PTMO (1:3). Even though the ratio of oxygen to carbon atoms for PPO and PTMO are the same, the site density is slightly higher in PPO which may be due to the arrangement of atoms in the PPO [5]. The rate of hopping sites, k 0 , with distance, r, can be calculated by using the following equation:
k 0 ( r ) = τ 1 e E d i s ( r ) / k b T   e E λ ( r ) / k b T
where τ 1 is frequency, and E d i s and E λ are the dissociation and reorganization energies, respectively. Based on the calculation, the authors found that PMO exhibited the smallest value of k 0 , whereas that of PPO was the largest [5]. They have used these values to calculate lithium-ion diffusion, using the kinetic Monte Carlo simulation. Among the polyethers studied, the diffusion of lithium ions in P(EO-TMO) was the highest [5].
Diffusion coefficient of ions, D, can also be determined from the following equation:
D = d 2 4 τ 2 δ 2
where d is the thickness of the electrolyte, τ 2 = 1 / ω and δ = d k 1 / A ε ε 0 (where ω is the angular frequency taken at the minimum value of impedance in Bode plot; k 1 is double layer capacitance; ε is dielectric constant of the electrolyte; and ε 0 is the permittivity in free space). Based on Equation (5), D in PVA-based polymer electrolyte containing potassium iodide salt has been calculated [6]. The k 1 value was obtained by fitting the Nyquist plot of the electrolyte. Since the complex impedance plot shows only a tilted spike, the plot has been fitted based on the series connection of bulk resistance, Rb, and constant phase element, CPE.
The total impedance (real, Z and imaginary, Z ) can be described as follows:
Z = R b + c o s   ( π p / 2 ) / k 1 ω p
Z = s i n   ( π p / 2 ) / k 1 ω p
where p is a fraction of a right angle. The mobility, μ, and number density, n, of ions can be obtained by using the following equations:
μ = q D / k b T
n = σ / q μ
It is evident that Equation (1) is the combination of Equations (8) and (9). The calculated transport properties (D, n and μ) have been used to explain the conductivity behavior in PVA-based electrolyte containing different salt concentrations [8]. Relationship between σ, n and μ can be expressed as follows:
σ = q n μ
In our earlier studies [8], it was found that the ionic mobility has contributed more in the highest conducting electrolyte, since n decreased due to the establishment of ion pairs and ionic aggregates. Noor has calculated n for lithium triflate in gellan-gum-based SPE by using the method from Fourier transform infrared (FTIR) spectroscopy [9,15]. Areas of free ions (Af), ion pairs (Ap) and ion aggregates (Aa) have been identified via FTIR deconvolution in wavenumber region between 1030 and 1060 cm−1, and their percentage can be calculated based on the following equations:
Free   ions   ( % ) = A f A f + A p + A a × 100
Ion   pairs   ( % ) = A p A f + A p + A a × 100
Ion   aggregates   ( % ) = A a A f + A p + A a × 100
From Equation (11), n can be calculated from the following:
n = M × N A V × 2 × p e r c e n t a g e   o f   f r e e   i o n s
where NA is the Avogadro’s number, M is number of moles of lithium triflate and V is the electrolyte’s volume.
The n value can also be calculated from the famous Rice and Roth model which was originally developed to describe the conductivity behavior for superionic conductors [16]. The equation is as follows:
σ = 2 3 [ ( Z q ) 2 m k b T ] n E A τ   e x p [ E A k b T ]
where EA is the activation energy; q is the electron charge; m is the mass of ions; Z is the vacancy of conducting species; and τ = r / ν , where r is the distance between two coordinating sites and ν is the velocity of ions, which is given as follows:
ν = 2 E A m
EA is the minimum energy required for an ion to hop in between sites, and it can be obtained from the Arrhenius equation, as shown below:
σ = σ 0 e x p ( E A k b T )
where σ 0 is the pre-exponential factor. A low EA value is required for high conductivity. In general, the mechanism of ion transport in polymer electrolytes is via hopping and coupled with the segmental motion of the polymer [3]. However, Grotthuss motion which is another type of hopping mechanism has been proposed to describe proton conduction in chitosan-based SPE containing ammonium iodide (NH4I) [10]. Chitosan structure is rigid and possesses high Tg, thus resulting in minimal chain motion. This led to the proposal of the Grotthuss model as the proton-conduction mechanism in the SPEs. Grotthuss mechanism is the proton conduction process where the proton hops between the coordination sites decoupled from the polymer chain motion. This mechanism is also known as proton hopping. Besides these factors, other parameters that can influence the feasibility of polymer electrolytes in applications for electrochemical cells include ion (cationic and anionic) transference number (tion), degree of crystallinity, etc.

3. Basic Solid Polymer Electrolytes

A simple SPE consists of a polymer and a salt where both are dissolved in a solvent before the solvent is allowed to evaporate, while the ions continue to be free and mobile inside the polymer matrix [3]. The salt can be either inorganic or organic salt but when used in DSSCs application, the SPE must contain additional materials, i.e., iodine (I2) crystals to act as redox mediator, while the salt must be iodide-based salts, since iodide/triiodide (I/I3) redox couple is commonly used [17]. Likewise, for application in QDSSCs, the basic SPE should have sulfide/polysulfide (S2−/Sx2−) redox mediator with sulfide-based salts. Similarly, the salt should be lithium-based, sodium-based, magnesium-based and proton-based if it is going to be employed in cells of Li-ion, Na-ion and Mg-ion and proton batteries as well as PEMFCs. Besides this, the salt must meet the criterion of having low lattice energy for easy solvation in polymer matrix [3]. Table 1 and Table 2 list some of the most popular polymers and salts used in polymer electrolytes, along with their special traits that enable them to be favored among others by the researchers worldwide. The Tg of polymer will affect the crystallinity of the electrolytes, and usually low Tg is preferred. Meanwhile, the lattice energy of salt will determine how easy it can dissociate in the polymer matrix. The lattice energies of the salts in Table 2 are calculated in this work, using the Kapustinskii equation [18]. All materials have both their own benefits and drawbacks. For instance, PEO, being the first polymer host used in polymer electrolytes and still remaining one of the most popular polymers until now, is favored for its talent to solvate many types of salts and exhibit good mechanical strength and thermal stability. It is a semicrystalline macromolecule, and thus, its crystalline part restricts the increment of conductivity. As for the salts, lithium perchlorate (LiClO4), as an example, shows good ionic conductivity, but it is a strong oxidizer, hence making it explosive, and so it must be handled with caution. As for the solvent choice, certain features, including dielectric constant, melting point, boiling point and viscosity, have to be taken into consideration for salt dissociation and solvation of polymer. Some frequently employed solvents are summarized in Table 3, along with their physical attributes.

3.1. Preparation Methods of SPEs

SPEs can be prepared via several methods, including solution casting [31], phase inversion [32], photopolymerization [33] and electrospinning [31] techniques, to name a few. Solution cast or solvent casting method is the most conventional and easy way to prepare electrolyte film. In this procedure, the polymer and salt are dissolved in an appropriate solvent before casting onto suitable substrates (e.g., glass plates). The solvent is then evaporated after being left to dry either at ambient condition or elevated temperature inside oven. Solution casting route can be performed at room temperature atmosphere or inside glove box, depending on its needs. After complete drying, the SPE is a free standing membrane with thickness usually in millimeter range. Solangi and co-authors [31] have compared PVA-based SPE containing potassium iodide (KI), potassium chloride (KCl) and sodium chloride (NaCl) salts prepared by using electrospinning and solution-casting techniques. Regardless of which salt was used, the electrolyte fabricated via the former method demonstrated superior ionic conductivity (5.95 × 10−6 S cm−1 for NaCl-containing SPE) than that prepared by the latter (1.87 × 10−6 S cm−1 for SPE having NaCl) [31]. The electrospinning approach can produce film of fibers morphology with a large surface-area-to-volume ratio. Moreover, the electrospun membranes were more thermally stable than the solution-cast prepared films. Among the three salts, electrolytes comprising NaCl exhibited the highest conductivity irrespective of the preparation method [31]. Nonetheless, there are many parameters that need to be controlled and optimized in electrospinning process such as polymer molecular weight and solution viscosity, conductivity and surface tension. Furthermore, the electrospinning setup conditions, e.g., flow rate, applied voltage and distance from tip of syringe containing solution to collector, are tedious and time-consuming. In contrast to electrospinning, the solution-casting route is a more straightforward and convenient approach for membrane preparation.
The phase inversion technique can produce porous structure which is beneficial for ionic transport. Using this method, SPE membrane consisting of PDVF-HFP, LiClO4 and cerium oxide (CeO2) exhibited sponge-like morphology with tiny pores and ionic conductivity of 2.50 × 10−3 S cm−1 at 25 °C [32]. In this procedure, a solution containing PVDF-HFP, CeO2 and NMP solvent was stirred for 24 h before being coated onto a glass plate via the doctor-blade method. This was followed by immersion in deionized water (about 3 to 5 h) for solvent extraction and phase inversion, which took place before drying under vacuum-pressure condition. The last step was to soak the membrane in LiClO4 liquid electrolyte for 6 h [32]. Liu et al. [34] have combined phase-inversion and chemical-reaction routes to obtain PVDF-PAN-SiO2-LiPF6 SPE membrane that gave ambient conductivity of 3.32 × 10−3 S cm−1. For comparison, the same composition membrane has been prepared by phase-inversion method, and it was found that the ambient conductivity was 2.83 × 10−3 S cm−1 [34]. The higher conductivity obtained by the membrane prepared by the combined method was attributed to larger porosity and better homogeneous pores distribution as compared to that developed from the phase-inversion technique only [34]. The former also showed better properties in terms of heat resistance, mechanical strength and electrochemical stability than the latter [34]. An unconventional way to prepare SPE films is through dry-mixing without the use of solvent. Dry-mixing is combined with hot-pressing. Eriksson et al. [35] have obtained thin films of SPEs containing poly(3,3-dimethylpentane-2,4-dione) as the polymer matrix and LiTFSI salt after pressing the dried homogenous mixture between two polytetrafluoroethylene (PTFE) plates, under high pressure (2 MPa) and temperature (100 °C) conditions, inside an argon-filled glove box for 1 h. The polymer was desiccated in vacuum state before uniformly mixing and pulverizing the polymer and salt. Although the ambient conductivity value was low, of the order of 10−8 S cm−1, it gave a rather high Li+ ion transference number (tLi+ = 0.70) at 80 °C [35]. For the sake of comparison, the LiTFSI-containing SPE but with PEO as polymer host has been prepared by the same method and it showed a higher ambient conductivity (10−5 S cm−1) with low Li+ transference number (tLi+ = ~0.15) [35]. High cationic transference number is useful for application in Li-ion batteries. Table 4 shows some selected works of simple SPEs systems available in the literature.

3.2. Strategies to Improve Ionic Conductivity of SPEs

In order to boost the ionic conductivity of SPEs, there are many approaches which have been taken by researchers, as stated in the following subsections.

3.2.1. Incorporation of More Than One Salt

One viable approach to boost the ionic conductivity of SPEs is to add another salt. The combination of two or more salts in the right proportions can give synergistic outcomes. In most cases, the sum concentration of salts is the same as the highest conducting electrolyte that contained only on one type of salt. This means that the total weight of salts remained unchanged, but an appropriate amount for salt one has been substituted by salt two. Since the amount of each salt in a mixed salt electrolytes is lower than the concentration of salt in single-salt electrolyte systems, the probability of ion recombination to form ion pairs will be reduced in former case, and thus, there are more free ions for conduction [48,49]. In addition, decrement in Tg and degree of crystallinity (%χc) of the electrolyte are observed in mixed salt systems [48]. Tao and Fujinami [50] reported that highest conductivity was achieved at 5 × 10−5 S cm−1 (30 °C) in PEO-based SPE containing lithium borate and lithium aluminate salts as compared to PEO electrolyte with one salt only regardless which one. Only minimal capacity fading was observed in LiNi0.8Co0.2O2/Li cell with double salt-containing SPE after 30th cycle, but no results were given for that of single salt [50]. The type of cation and anion determine the electrochemical performance of devices. For Li-ion, Na-ion and Mg-ion cells, it is obvious that dissimilar anions influence the performance of the devices. For DSSCs, the variation of cations since the anion is fixed following the redox mediator used. Studies of mixed salt systems containing different cations [51,52] or dissimilar anions [49,53,54] have shown superior ionic conductivity than single salt systems. For instance, higher conductivity was obtained for an electrolyte based on PEO-CaBr2-CaI2 (30:1:1) than for a PEO-based electrolyte with only one salt, i.e., either calcium iodide (CaI2) or calcium bromide (CaBr2) [52]. Similarly, enhanced conductivity has been detected for PEO electrolyte having dual salts of zinc bromide (ZnBr2) and lithium bromide (LiBr) [51] and PEO-based electrolyte comprising the same compositions of binary salts of copper trifluoromethanesulfonate (Cu(CF3SO3)2) and zinc trifluoromethanesulfonate (Zn(CF3SO3)2) [55].
The incorporation of double salts does not always warrant higher conductivity than single salt but can still result in better performance of electrochemical devices, since conductivity is not the sole influencing factor. This has been showcased in the work of Dissanayake et al. [56], who have prepared PEO-based SPE with KI and TPAI salts via the solution-cast method. SPE containing sole salt of KI gave higher conductivity than electrolyte having double salts. Conductivity was lowest for electrolyte with lone TPAI salt since its cation is larger than K+ and thus less mobile. Moreover, the degree of crystallinity for PEO-TPAI electrolyte was higher than that of PEO-TPAI-KI SPE [56]. DSSC employing hybrid salts displayed the best performance, with an efficiency (η) of 4.22% and short circuit current density (Jsc) of 8.00 mA cm−2, followed by cells having single TPAI (Jsc = 6.38 mA cm−2; η = 3.54%) and KI (Jsc = 6.52 mA cm−2; η = 3.09%) under illumination of 100 mW cm−2. Their open-circuit voltage (OCV) values were 0.69, 0.77 and 0.75 V for KI, TPAI and KI+TPAI electrolytes, respectively [56]. A large-sized cation boosted OCV value, while it reduced Jsc, as can be seen when comparing the two single salt-containing electrolytes. The addition of a second salt decreased the cation conduction and enhanced Jsc, leading to a higher η [56].
Zhao and co-workers [57] have studied the influence of primary, secondary and ternary salts on PEO-based SPEs incorporated with halloysite nanoclay (HNC) for application in Li-ion batteries. For PEO-LiTFSI-HNC SPEs, the ambient conductivity was ~5.62 × 10−5 S cm−1, and there was a slight increment when a secondary salt, i.e., lithium bis(oxalate)borate (LiBOB), was added, but then there was a decrement in conductivity (~1.99 × 10−5 S cm−1) was observed with the addition of a third salt, lithium nitrate (LiNO3) [57]. Nonetheless, Tg values were enhanced from −47.02 to −40.48 °C for single to ternary salt-containing SPEs [57]. Mechanical strength also improved as proven from dynamic mechanical analysis (DMA). The authors have concluded that each respective salts have their own role with LiTFSI being the main contributor to ionic conduction, whereas the presence of LiBOB and LiNO3 salts was to ensure the emergence of cathode electrolyte interphase and solid electrolyte interphase SEI on cathode and anode, respectively [57]. Each of these salts have complemented each other and yielded Li-ion-cell performance with better electrochemical stability. Cells employed one and two salts suffered failure within the first five cycles. In contrast, the lithium nickel cobalt manganese oxide (NMC)/Li cell with SPE containing trio salts exhibited the discharge capacity of ~116 mAh g−1 at 60th cycle [57]

3.2.2. Modification on Polymer Host

The partial crystallinity nature of polymer host is usually the main culprit behind the low conductivity of SPEs. This issue can be resolved by making some alteration on the polymer itself. For example, biopolymers, i.e., cellulose and chitin, are highly crystalline with rigid polymer chain backbone, even though they can still serve as host in polymer electrolyte. Cellulose can be chemically modified into methyl cellulose (MC), carboxymethyl cellulose (CMC), hydroxypropyl cellulose (HPC) and sodium carboxymethyl cellulose (NaCMC), to name a few. Similarly, chitosan structure can be altered into carboxymethyl chitosan, hexanoyl chitosan (HCh), N-methylene phosphonic chitosan, chitosan–N-propyl sulfonic acid, etc. Besides reducing crystallinity, the modified polymer will have greater solubility to dissolve in various solvents, as compared to its predecessor. N-phthaloyl chitosan (PhCh) is the product obtained after the chemical adjustment of chitosan in a solution mixture of phthalic anhydride and DMF under a nitrogen gas environment. Chitosan has been successfully transformed into PhCh with the presence of peaks belonging to the pthalimido functional group at wavenumbers 1713 and 1733 cm−1, as proven by FTIR analysis [58]. Dilute acids of formic, acetic, butyric, lactic, maleic and malic are some solvating agents for chitosan [59,60], but the water content may deteriorate the performance of some electrochemical devices, such as Li-ion batteries and DSSC. PhCh can dissolve in pyridine, DMAc, DMF, DMSO and m-cresol [61,62]. Aziz et al. [58] reported that PhCh is more amorphous and thermally stable than the antecedent, chitosan. The former requires a higher decomposition temperature (356 °C) than the latter (290 °C). PhCh-based SPE consisting of NH4SCN prepared via solution-casting method displayed conductivity of 2.42 × 10−5 S cm−1 at 25 °C and tion = 0.91 [58]. The PhCh film had the ambient conductivity of ~1.93 × 10−10 S cm−1 [58]. O-nitrochitosan is another derivative of chitosan after reacting with sodium hydroxide and fumed nitric acid, as synthesized by Rahman et al. [63]. The appearance of new bands at wavenumbers 1355 and 1646 cm−1 were attributed to nitro (-O-N=O) functional groups, as evidenced by the FTIR spectrum of o-nitrochitosan [63]. The ambient conductivity was greatly enhanced for the modified chitosan film (5.22 × 10−6 S cm−1) at room temperature, while the unmodified chitosan film exhibited low conductivity of 8.88 × 10−10 S cm−1 [63]. The large increment was attributed to the high electronegativity of nitro group [63]. Although no dopant salt was incorporated, the authors believed that o-nitrochitosan can be a promising polymer electrolyte membrane in PEMFC [63].
PEO is excellent to solvate Li+ ion, which is especially useful for application in Li-ion cells, even though its crystalline portion hampers the conductivity enhancement. Usually, the conductivity of PEO electrolyte containing lithium salts is no greater than 10−4 S cm−1 under ambient condition with a low value of lithium-ion transference number (tLi+ < 0.40) [64]. Luckily, its chemical structure can be modified and incorporated with various types of functional groups or moieties. Another strategy to improve the conduction of PEO-based electrolyte is to modify its linear chain into hyper-branched structure. Jing et al. [64] have prepared LiTFSI-containing SPE based on PEO having a hyper-branched arrangement and grafting with another PEO of linear chains to improve the ionic conductivity without sacrificing its mechanical strength. An increment by one order of magnitude was observed when compared to SPE having PEO linear chain, which was attributed to reduction in crystallinity [64].
Besides polymer modification, the polymer can also undergo copolymerization, grafting, crosslinking and blending with a second polymer in order to lower the crystallization. Imperiyka and co-authors [65] have photo-copolymerized polyglycidyl methacrylate (P(GMA)) with PMMA as the polymer host. SPE with configuration P(GMA-co-MMA)-LiClO4 displayed the ambient conductivity of 8.70 × 10−6 S cm−1 [65]. The bonding of one polymer chain with that of another polymer is known as a crosslinking reaction. This process can take place physically and chemically (photo-induced polymerization, curing, etc.). In 2016, Lehmann et al. [66] prepared SPE containing PEO (Mw 2000 g mol−1) crosslinked with poly(ethyleneimine) (PEI) (Mw 600 g mol−1) and LiTFSI salt and obtained the conductivity of 3.90 × 10−5 S cm−1 at 40 °C. The values of Tg, percentage of crystallinity, tensile strength and cationic transference number were −33 °C, 20%, 0.47 MPa and 0.76 (40 °C), respectively. Meanwhile, SPE, which comprises 600 Mw PEI, with the same salt, attained the conductivity of 1.50 × 10−6 S cm−1 at 40 °C and Tg of −65 °C [66]. In 2016, Youcef et al. [67] systematically studied the concentration effect of crosslinking agent (divinylbenzene (DVB)) on SPE based on PEO crosslinked poly(ethylene glycol) diacrylate (PEGDA) with LiTFSI salt that had been prepared via polymerization under ultraviolet (UV) irradiation, using a photo-initiator, -hydroxy-2-methylpropiophenone and acetonitrile solvent. At optimum concentration of 10%mol DVB, conductivity of 1.40 × 10−4 S cm−1, lithium-ion transference number (tLi+) of 0.21 and high electrochemical window of 4.30 V were obtained at 70 °C [67]. At the same temperature and C/10 rate, LiFePO4/Li battery employing this electrolyte delivered the first discharge capacity of 138 mAh g−1 which was sustainable even after the 20th cycle (138 mAh g−1) [67]. The SPE has also been tested in Li-sulfur battery under similar conditions and acquired the capacity of 175 mAh g−1 after the 50th cycle, with an initial capacity of 375 mAh g−1 [67]. After cycling, there was no lithium dendrite formation, as evidenced through SEM micrograph. In addition to inhibit crystallinity percentage, the crosslinked electrolyte is said to have the ability to hamper lithium dendrite growth when being used in Li-ion cells and to improve dimensional stability at elevated temperatures [66,67,68,69].
In a separate report, PEO underwent crosslinking reaction with tetraethylene glycol dimethyl ether (TEGDME) employing 4-methyl benzophenone (MBP) as crosslinker agent, LiTFSI salt under UV radiation without solvent usage [68]. It is said that MBP can remove H+ from its methylene group via hydrogen abstraction to form free-radical chains [68]. Likewise, TEGDME that contained the methylene group produced free radicals. These free-radical chains have interlinked with -EO- unit of PEO chain under UV exposure to form SPE film that displayed features, such as being highly amorphous and having a uniform morphology, high mechanical strength, good flexibility, low Tg of −34 °C and compatibility with Li metal electrode. At 25 °C, the SPE showed a conductivity of 0.11 × 10−3 S cm−1, tLi+ of 0.55 and diffusion coefficient of Li+ ion of 5.6 × 10−6 cm2 s−1 [68]. The TiO2/Li cell having such an electrolyte had good cyclability over 100 cycles, with a capacity of ~138 mAh g−1 and almost 90% capacity retention [68]. Lim et al. [70] employed an in situ polymerization procedure to prepare SPE based on crosslinking poly(ethylene glycol) diglycidyl ether (PEGDE) and bisphenol A diglycidyl ether (BADGE) with 1-ethyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide (EMIMTFSI), using PEI as crosslinker without solvent. The SPE was prepared directly on activated carbon electrode for supercapacitor application before the curing process. Such an SPE film showed a conductivity of 2.41 × 10−3 S cm−1 at 25 °C and capacitance value of 19.8 F g−1 at 2 mV s−1 scan rate with decent cyclability upon 10,000 cycles [70].
Chitosan in 1% acetic acid solution was subjected to a crosslinking process with methanesulfonic acid (MSA) and dodecylbenzene sulfonic acid sodium salt (DBSA-Na), using sulfuric acid as an agent to form chitosan-MSA and chitosan-DBSA membranes, respectively [71]. Such an action improved the proton conductivity and the thermal and mechanical stabilities of the chitosan-based membranes, which can be applied as solid electrolyte membranes in PEMFCs. At the elevated temperature of 90 °C, the proton conductivity was 3.09 × 10−4 S cm−1 for chitosan-DBSA film, whereas 2.18 × 10−4 S cm−1 was the conductivity exhibited by the chitosan-MSA membrane [71]. The non-crosslinked chitosan membrane was also prepared for comparison and showed lowest conductivity of 1.04 × 10−4 S cm−1 at same temperature [71]. The grafted copolymers consisting of natural rubber and methyl methacrylate, namely 49% PMMA-grafted NR (MG49) [48,72,73] and 30% PMMA-grafted NR (MG30) [74], have been used as polymer hosts to produce polymer electrolytes. Su’ait et al. [72] reported a conductivity of 1.00 × 10–12 S cm–1 for pure MG49 at 30 °C, and this value increased to 2.30 × 10–7 S cm–1 when MG49 was incorporated with 20 wt.% LiBF4 salt. Such an electrolyte of MG49-LiBF4 was then added to another salt (LiI and LiCF3SO3) by the same group of researchers for the mixed-salt effect [48]. It was found that the ambient conductivity increased with different salts and combination of salts as follows: LiCF3SO3 (2.15 × 10−11 S cm−1) < LiI (4.73 × 10−9 S cm−1) < LiCF3SO3+LiBF4 (7.71 × 10−9 S cm−1) < LiBF4 (1.42 × 10−8 S cm−1) < LiBF4+LiI (1.89 × 10−6 S cm−1) [48]. This clearly demonstrates that the ionic conductivity was enhanced significantly for the MG49-based SPE having double salts of LiBF4 and LiI at a ratio of 30:70, while the addition of LiCF3SO3 into the MG49-LiBF4 electrolyte had an adverse effect on the conductivity. The transference number of ions for the five samples was observed to increase in the same order of conductivity [48]. It can be understood that the degree of crystallinity increased in the reverse order [48]. Modified natural rubber has superior elastic features for better interfacial contact between electrolyte and electrode when applied in electrochemical devices. Another strategy to improve the properties of electrolyte is to bind the anion directly to the polymer chain to form a single-ion-conducting polymer electrolyte. Such an electrolyte consisting of poly[lithium 1-[3-(methacryloyloxy) propylsulfonyl]-1-(trifluoromethylsulfonyl) imide] (PLiMTFSI) attached to PEG-crosslinked poly(2-hydroxyethylacrylate) (PHEA) showed improved mechanical strength as compared to an electrolyte of crosslinked PEG-PHEA doped with LiTFSI, even though its conductivity is lower than the latter [75]. Nonetheless, the single-ion-conducting polymer electrolyte could minimize lithium dendrite formation and gave better results in Li-ion cells.

3.2.3. Treatment of SPEs

The polymer electrolytes can also be subjected to radiation in order to tune their properties in terms of structural, thermal, electrical and mechanical aspects. Various radiation sources, including gamma rays, X-rays, ultraviolet rays, electron, neutron and ion beams, can be employed. Solution-cast-prepared SPE based on PEO with lithium sulfate (Li2SO4) salt has been irradiated with low-energy electron beams of 8 MeV by Raghu et al. [76]. At 30 °C, the conductivity was enhanced to the order of 10−4 S cm−1 for PEO-Li2SO4 electrolyte after exposure to 30 kGy dose of electron beam from 10−5 S cm−1 for non-irradiated electrolyte. Analyses from DSC and X-ray diffraction (XRD) demonstrated that the former have a lower melting temperature (67.6 °C) and are more amorphous in nature, with reduced crystallinity, than the latter (69.4 °C melting temperature and more crystalline), while FTIR results indicated that the polymer chain was disrupted and degraded after irradiation [76]. Irradiating of Li3+ ion at high energy of 60 MeV on two different electrolytes of PEO-NaI and PVP-NaI showed increment in conductivity as compared to unirradiated electrolyte films [77]. Besides conductivity, higher values for number of mobile ions, ionic mobility and dielectric constant have been reported for the solution-cast-prepared SPE after ion-beam irradiation [77]. PAN-based SPE having LiBOB salt also exhibited enhancement in conductivity and transport properties (n, μ and D) after being radiated with a gamma-ray dose of 15 kGy [78]. Ambient conductivity was increased by 58.6 times from 1.74 × 10−6 S cm−1 to 1.02 × 10−4 S cm−1, while improvement of 12.2, 4.8 and ~4.7 times was obtained in n, μ and D, respectively [78]. Likewise, PVDF-LiBOB SPE has attained higher values of conductivity and ionic transport upon gamma-ray radiation as compared to non-irradiated electrolyte [79]. Hema et al. [80] also reported similar observation in PVA-PVDF-LiCF3SO3-SiO2 SPEs, where conductivity was boosted to 9.40 × 10−4 S cm−1 at 30 °C after gamma-ray exposure (15 Gy dose). The conductivity of such an electrolyte without radiation treatment was 1.70 × 10−4 S cm−1. Ionic transference number for the irradiated SPE also showed increment (tion = 0.99) as compared to that of non-gamma ray treated SPE (tion = 0.98) [80]. On the other hand, lower values for the percentage of crystallinity, Tg and melting point were detected in the former than in the latter [80]. In general, it is to be noted that conductivity enhancement has been observed at optimized dosage of radiation irrespective of the sources due to polymer chain scission, whereas too much radiation dose will be detrimental to the properties of SPE, and this is attributed to crosslinking and increase in crystallinity.
In another work of Raghu et al. [81], a comparison was made between gamma-ray (cobalt-60 source) and electron-beam (energy of 8 MeV) treatments on PEO-based SPEs of cadmium chloride (CdCl2) salt. Non-irradiated electrolyte exhibited the lowest conductivity value at 8.63 × 10−6 S cm−1 at 30 °C, as compared to gamma-ray-treated (0.115 S cm−1) and electron-beam-irradiated (0.175 S cm−1) electrolytes. Micrographs from scanning electron microscopy (SEM) showed less PEO spherulites in both irradiated SPEs as compared to that of non-irradiated SPE, indicating that a more amorphous region existed in the former case. In addition, the SEM image for electrolyte after electron-beam exposure displayed a homogenous morphology, while the formation of aggregates was seen in the micrograph of SPE irradiated with gamma radiation [81]. The SPEs were irradiated with electron beam and gamma ray separately at two different doses (50 and 150 kGy), and it was found that, at the 150 kGy dosage, both types of radiation gave a higher conductivity (of the order of 10−4 S cm−1) than that irradiated at the 50 kGy dose (around 10−5 S cm−1) [81]. The authors have concluded that the electron-beam treatment is an effective approach as compared to gamma-ray radiation in making the SPE more conducting due to reduction in crystallinity. Recent work illustrated contradictory results, wherein no effect was observed on the conductivity upon neutron irradiation at varied dosages (15 to 47 kGy) on PVDF–LiBOB SPEs prepared via solution-casting route [82]. Ambient conductivity was the highest at 1.27 × 10−5 S cm−1 for the non-irradiated SPE as compared to similar electrolyte after exposed to neutron radiation with lowest value at 4.30 × 10−7 S cm−1 for SPE subjected to 47 kGy dose [82]. This was attributed to decrement in μ and D, as well as higher degree of crystallinity. The authors have deduced that an improvement in mechanical integrity in the electrolyte has occurred after being irradiated with a neutron [82]. Sinha et al. [83] have observed that there was a slight decrement in conductivity at ambient temperature for the solution-cast-prepared SPE based on PEO with ammonium perchlorate (NH4ClO4) salt after being irradiated with gamma rays, followed by minor conductivity increment at 50 °C. In the following sections, additional ingredients such as secondary polymer, plasticizers, fillers and other additives that have been integrated into various SPEs are discussed.

3.2.4. Blended Solid Polymer Electrolytes

Polymer blending can be regarded as the most convenient and simplest way to prepare SPEs that are highly conducting, flexible and thermally stable and that have good mechanical strength without additional employment of materials such as initiator, linker agent, radiation sources, etc., thus indirectly making this approach more cost-effective as compared to other techniques aforementioned. In this technique, two or more polymers which are miscible in the appropriate solvent form a homogenous solution before addition of the conducting salt. The blending of polymer approach is able to (i) reduce the glass transition temperature, (ii) decrease the degree of crystallinity, (iii) increase the ionic conductivity, (iv) improve the thermal stability, (v) enhance the mechanical strength and (vi) boost the performance of electrochemical devices. There are many studies on blended SPEs available in the literature. Table 5 lists some of the systems under this category.
Sengwa et al. [100] have investigated the effect of SPE preparation method on conductivity. The SPE comprising PEO blended with PMMA at a ratio of 1:1 and LiCF3SO3 salt has been prepared in four ways: (i) conventional solution cast coupled with hot melt-pressing; (ii) ultrasonication followed by solution casting and hot melt press; (iii) irradiation of microwave before solution-cast and hot-melt-press techniques; and (iv) ultrasonication followed by microwave exposure, solution casting and hot melt press. The authors have observed that conductivity decrement of the similar SPE in the following order: σ(route iii) (1.99 × 10−6 S cm−1) > σ(route iv) (1.56 × 10−6 S cm−1) > σ(route i) (1.36 × 10−6 S cm−1) > σ(route ii) (0.11 × 10−6 S cm−1) [100]. They have speculated that the electrolyte made from microwave-assisted solution-cast/hot-melt-press procedure has produced more of a pathway for ionic conduction as compared to other methods [100]. Another unrelated work has compared the usefulness of the blending approach with grafting technique on LiClO4-containing SPE system with sulfonated polyether ether ketone (SPEEK) and polyethylene glycol (PEG) as polymer hosts [101]. The SPEEK grafted with PEG electrolyte gave higher conductivity (around 10−5 S cm−1) when compared to SPEEK–PEG blended electrolyte (10−6 S cm−1) at 30 °C, thus implying the grafting method can suppress PEG crystallinity more than the blending technique [101].
Liu and co-workers [102] have combined both chemical (crosslinking) and physical (blending) routes together, i.e., crosslinked (3-glycidyloxypropyl)trimethoxysilane (GLYMO) with polyetheramine and also crosslinked poly(ethylene glycol) diglycidyl ether (PEGDGE) with polyetherdiamine before blending their products at 70:30 ratio in THF solvent and added with LiClO4 salt. The [EO]/[Li] ratio was optimized at 16:1. Such a unique SPE exhibited the conductivities of 1.20 × 10−4 S cm−1 and 8.30 × 10−4 S cm−1 at 30 and 80 °C, respectively [102]. The Tg value was −43.6 °C, and the transference number of Li+ ions was 0.28 at 70 °C. This electrolyte was thermally stable up to 260 °C and gave the capacities of 110 mAh g−1 (The first cycle) and 61 mAh g−1 (100th cycle) when applied in LiFePO4/Li battery at 0.2 C rate [102]. As a summary, blended SPEs show superior properties than SPEs containing single polymer matrix.

3.2.5. Plasticized Solid Polymer Electrolytes

The main idea of plasticized polymer electrolyte is to improve the elasticity by lowering the Tg. As stated in the previous section and Equation (3), the segmental movement of the polymer backbone contributes to the conductivity of SPEs. Increasing the segmental motion will improve the ionic conductivity and can be easily achieved by adding plasticizers into the polymer matrix. There are many types of plasticizers such as EC, PC, PEG, DMSO, diethyl carbonate (DEC), dimethyl carbonate (DMC), glycerol, gBL and DMF to name a few that affect the ionic conductivity of SPEs in different ways. Note that EC, PC, DMSO, gBL and DMF can also serve as a solvent, as listed in Table 3, while PEG of low molecular weight (Mw) is usually used as a plasticizing agent. Plasticizers with high a dielectric constant, e.g., EC, as mentioned in the previous section, helps in salt dissociation, lowering the formation of ion pairs and ion aggregates leading to conductivity enhancement. Besides high dielectric constant, a good plasticizer should possess low viscosity to increase the mobility of ions. Furthermore, the incorporation of plasticizer in an electrolyte offers additional passage for ion transport. Some plasticized polymer electrolytes in the solid form that have been studied are tabulated in Table 6.
It is interesting to highlight that ionic liquids can play three roles, viz as plasticizer (to enhance the flexibility of polymer backbone), conducting salt (provides cation and anion) and solvent. They are non-volatile, safe (due to non-flammability), chemically and thermally stable, and are not prone to evaporation and crystallization [130]. In addition, they exhibit low melting temperature and outstanding ionic conductivity at room temperature condition, and they also have superior solvating ability. In the work of Widstrom et al. [114], increment in conductivity has been observed after incorporating ionic liquid, S2TFSI into PEO-LiTFSI SPE without affecting Li+ transference number. From Table 6, it is to be noted that N1222FSI is not an ionic liquid but an ionic plastic crystal that shows similar features to that of ionic liquids [115]. In contrast to conventional plasticizers, the mechanical integrity can still be retained upon addition of N1222FSI into the electrolyte besides reducing the polymer crystallinity and dissociating the salt. Moreover, the interfacial contact between electrolyte and electrode can be improved [115].
Another organic plastic crystal, i.e., succinonitrile (SN) is favored for its outstanding solvating capability in dissociating many salts including Li-based salts in battery application. As a result, conductivity is enhanced as can be observed in refs [105,112] from Table 6. Unfortunately, the plastic nature of SN deteriorates the mechanical strength of the electrolyte. Nonetheless, this can be easily avoided by preparing the electrolyte membrane in a different way. He et al. [131] prepared free-standing film containing PEGDA (Mw 6000 g mol−1), LiTFSI and SN via photopolymerization procedure using bis(2,4,6-trimethylbenzoyl)-phenylphosphine oxide (also known as Irgacure 819) as photo-initiator but without solvent. Such electrolyte which was highly amorphous with crosslinking network exhibited good ambient conductivity of 1.10 × 10−3 S cm−1, electrochemical stability window of 4.80 V with improved mechanical strength (0.24 MPa tensile strength, 84% elongation at break). Li4Ti5O12/Li cell employing this electrolyte gave promising performance (The first discharge capacity of 140 mAh g−1 at 0.2 C) with 93% capacity retention for 25 cycles [131]. On the other hand, it is evident from Table 6 that PEG of low molecular weight acted as plasticizer rather than secondary polymer in PEO-KI electrolyte which showed better conductivity [113] than PEG-free SPE even though PEG might have crosslinked with PEO. Instead, conductivity improvement is attributed to extra passageways for ion migration through PEG segmental movement.

3.2.6. Composite Solid Polymer Electrolytes

Although addition of plasticizers in SPEs can easily improve the ionic conductivity, the mechanical strength of the SPEs is compromised. On the contrary, incorporating fillers in SPEs is known to improve the performance of SPEs not merely the ionic conductivity but also its mechanical strength since fillers can reduce the crystallinity of SPE films. SPE incorporated with an inorganic filler are termed as composite solid polymer electrolytes (CSPEs). In most CSPEs systems, the polymer is usually regarded as Lewis base if it donates its lone pair of electrons, while the filler is a Lewis acid in accepting the electrons. Fillers can be categorized into two types, i.e., active and inactive. Some examples of inactive fillers are titanium dioxide (TiO2), silicon dioxide (SiO2), alumina (Al2O3) and zirconia (ZrO2). The effects of fillers are varied depending on their particles size, surface nature and amount or concentration in the CSPEs.
Tan et al. [132] investigated the effect of SiO2 and Al2O3 on the PMMA-based SPEs containing LiCF3SO3 salt. Addition of Al2O3 has improved the ionic conductivity by one order of magnitude from 1.36 × 10−5 S cm−1 to 2.05 × 10−4 S cm−1. However, this effect cannot be seen for CSPE containing SiO2 [132]. The reason behind Al2O3 showed significant improvement in conductivity may be because of the OH surface nature in alumina that provided additional coordination sites to oxygen atom in PMMA for ionic conduction. Moreover, the presence of H+ on the alumina surface also help in salt dissociation thus increasing the number density of ions leading to ionic conductivity improvement of the CSPE [132]. In the same work, they also reported that the amount of Al2O3 influenced the conductivity of CSPE [132]. The presence of high amount of alumina grains in CSPE may cause the polymer chain to be less mobile thus affecting the transport of ions that resulted in conductivity decrement. Nonetheless, it has been reported that the mechanical strength of CSPE was enhanced at high concentrations of filler [133]. Too much or too low content of fillers would not bring positive effect on the ionic conductivity [134]. Only at optimized concentration can lead to its improvement in the CSPE [134]. The findings on acidic surface nature of alumina that enhanced the conductivity of SPE [132] are also supported by Croce et al. [135]. Following Reference [135], it can be understood that extra pathway for ion transport was formed via hydrogen bonding when the surface of Al2O3 nanofiller (particle size 5.8 nm) was acidic. The SPE system that comprised PEO, LiCF3SO3 and acidic Al2O3 exhibited ionic conductivity of 2.10 × 10−5 S cm−1 at 30 °C. Slight reduction in conductivity (1.20 × 10−5 S cm−1 at 30 °C) was observed for Al2O3 neutral surface which may be due to weaker interactions as compared to that in acidic form [135]. On the contrary, there was no interaction between filler, anion of salt and polymer in the case of basic Al2O3 surface thus subsequently causing detrimental outcome to conductivity (7.30 × 10−6 S cm−1 at 30 °C) [135].
In another work, it was also observed that the addition of SiO2 successfully boosted the ionic conductivity for PVA-PVP-PEG-NaI electrolyte [136]. An improvement in amorphous region of the electrolyte film via the interaction of functional groups of the polymer with the Si-O-Si of the filler may be the reason for the conductivity increase. Li and Lian [137] studied the effect of SiO2 particle sizes (4 μm and 7 nm) in hydroxide conducting poly(acrylamide) (PAM) based SPE for alkaline batteries. The hydroxide conductivity of the SPE was amplified with the incorporation of 7 nm SiO2 but not for SPE containing 4 μm SiO2. Based on the micrographs and Raman studies, smaller size of SiO2 significantly helped to delay polymer crystallization [137]. Similar observation has been reported for Al2O3 where the smaller particle size prevented the formation of dendrites [138]. Other than size, the morphology of filler will also give different outcome to the CSPE. Lithium dendrites formation was minimized in the case of electrolyte containing PEO, lithium (bis trifluoromethyl)sulfate, SCN and SiO2 nanofibers as compared to similar electrolyte but with SiO2 nanoparticles [139]. It was observed that SiO2 aggregates was found in the latter, while the SiO2 nanofibers were homogeneously distributed in the electrospun CSPE as evidenced from SEM images. The CSPE of SiO2 nanofibers that had a 3D structure demonstrated better mechanical strength and flexibility with improved performance in lithium battery [139]. Incorporation of filler in electrolyte also yielded better interfacial contact with lithium metal electrode [140]. At 30 °C, SiO2 nanofibers CSPE exhibited the conductivity of 9.32 × 10−5 S cm−1 which was highest when comparing to CSPE with SiO2 nanoparticles (5.28 × 10−5 S cm−1) and SPE without SiO2 (4.72 × 10−5 S cm−1) [139]. Occasionally, an increment in cationic transference number was observed upon the addition of fillers in SPEs [3].
As discussed earlier, the contribution of inactive or passive fillers in ionic conductivity of SPEs is clear, i.e., improving the mobility of charge carriers either through the mobility of polymer chain (amorphousness) or by providing the additional transportation sites. Active fillers, such as Li1+xAlxGe2−x(PO4)3 [141], Li1+xAlxTi2−x(PO4)3 [142], Li7La3Zr2O12 [143], Li0.33La0.557TiO3 [144] and Li6.4La3Zr1.4Ta0.6O12 [145] on the other hand, contribute directly to the lithium-ion conductivity apart from impeding crystallization and improving the mechanical integrity of the SPEs. These materials by their own serve as solid electrolytes [146,147,148,149,150]. Nonetheless, the fabrication processes of these ceramics are usually complicated and costly [139]. In addition, the thickness of these solid electrolytes have to be large (around 100 μm in pellet form), since they are fragile and brittle [139] thus resulting in high interfacial resistance. These drawbacks lead to introducing them in SPEs as fillers instead. Filler-added SPEs are flexible, more stable and have better interfacial compatibility with Li metal electrodes [141,151]. Furthermore, the SPEs possess better processability and higher ionic conductivity than the solid electrolytes [151,152]. Similar to inactive fillers, the size, concentration and morphology of active fillers impose different impact on the CSPEs.
Among the inorganic active fillers, Li7La3Zr2O12 (LLZO) having garnet structure is favored due to its chemical stability against lithium metal and large electrochemical window (>5 V) [143]. LLZO nanofibers prepared via the electrospinning method have been introduced to an electrolyte comprising PVDF–HFP as a polymer host, LiTFSI salt and ionic liquid plasticizer [153]. The CSPE delivered high ambient conductivity (6.50 × 10−3 S cm−1), reduced crystallinity and wide electrochemical window (5.3 V) than plasticized SPE without LLZO as filler [153]. Comparison was made with basic SPE configuration of PVDF-HFP-LiTFSI in electrochemical performance of Li-ion battery. The LLZO CSPE cell displayed good cyclability over 1000 cycles with specific capacity of 149 mAh g−1, whereas the capacity decreased sharply after 50th cycle for the Li-ion cell with basic SPE [153]. Synergistic effect of LLZO filler and poly(ethylene glycol) dimethyl ether (PEGDME) plasticizer was also studied in PEO–LiTFSI SPEs system [154]. Again, improvements in conductivity, Li-ion transference number and electrochemical performance were reported [154]. In the work of Keller et al. [155], LLZO-added CSPE containing PEO and LiTFSI did not contribute to ionic conductivity but exhibited improved electrolyte-electrode interfacial contact and higher electrochemical performance in lithium battery. It was found that the addition of LLZO in SPEs suppressed lithium dendrites growth [152,153,154,155,156]. The lithium and zirconium sites of LLZO can be doped with other elements to boost its properties. Zhang and co-authors [157] employed tantalum (Ta) as dopant in LLZO to form Li6.75La3Zr1.75Ta0.25O12 (LLZTO). The PAN-LiClO4-LLZTO CSPE showed conductivity increment by 12 times (2.20 × 10−4 S cm−1 at 40 °C) and higher Li+ transference number of 0.30 than filler-free SPE [157]. In addition, the CSPE was thermally and electrochemically stable with enhanced mechanical strength [157]. It is worthy to mention that Li+ transference number as high as 0.75 was obtained in CSPE containing LLZTO, poly(propylene carbonate) and LiTFSI with conductivity of 5.20 × 10−4 S cm−1 at 20 °C [158]. It is understood that the addition of nano-sized LLZTO promoted salt dissociation due to the interactions between polymer, cation and filler and subsequently impeded the diffusion of anions [158].
LLZO is preferred over NASICON-type Li1.5Al0.5Ge1.5(PO4)3 (LAGP) due to their inexpensive elements [155]. In contrast, germanium in the latter is costly [155]. Nevertheless, LLZO has shortcomings; for example, it has an unstable at room-temperature condition, and thus, it is reactive towards moisture and humidity [155]. Similar to LLZO, LAGP is also unreactive towards lithium metal electrodes and has been used as a filler in polymer electrolytes. PEO, LiTFSI and LAGP were mixed and ground without solvent before heating and hot-pressing to form CSPE as reported by Piana et al. [159]. The CSPE showed improvements in term of conductivity, mechanical integrity and specific capacity as compared to electrolyte without filler [159]. On the other hand, the prepared CSPE in Reference [159] showed superior interfacial contact than pelletized LAGP solid electrolyte in lithium-ion cells. However, no significant change in lithium-ion transference number was observed with the introduction of LAGP in PEO-based electrolyte. On the contrary, LAGP added in poly(propylene carbonate)-based electrolyte increased lithium-ion transference number to 0.77 with conductivity of 0.56 × 10−3 S cm−1 and improved cyclability and stability in Li-ion battery [160].
Li et al. [13] employed electrospinning technique to synthesize Li0.33La0.557TiO3 nanorodsbefore adding in PVDF–HFP electrolyte containing LiTFSI salt. The CSPE was more thermally stable and showed decreased crystallinity with higher ambient ionic conductivity (1.21 × 10−4 S cm−1) and tensile strength (39.77 MPa) than filler-free electrolyte (σ = 4.72 × 10−5 S cm−1; tensile strength = 38.41 MPa) [13]. Ionic conductivity of PAN-based SPE has been shown to increase with the help of an active filler namely lithium titanate nanotubes (LiTNT) [161]. From the vibrational studies, Pignanelli and co-workers observed that the presence of LiTNT in SPE helped in the increment of lithium perchlorate dissociation, whereas the impedance spectroscopy showed two semicircles that are related to lithium-ion conductivity [161]. The semicircle in the higher frequency region is accredited to the bulk resistance of lithium transport across the polymer matrix, whereas the second is ascribed to lithium transport through the interaction with the LiTNT [161]. Some authors have proposed that lithium transport can take place via three possible routes, i.e., polymer—salt matrix, polymer—filler interface and inside filler grain [162,163]. Zheng et al. [163] showed that Li-ion movement took place mainly via ceramic filler LLZO grain for PEO-LiClO4-LLZO system, using NMR results. On the other hand, for PEO-LiClO4-TEGDME-LLZO electrolyte, it was found that Li-ion motion via TEGDME-related phase was preferred over filler or polymer matrix or polymer—filler interface [164]. In our opinion, a complete understanding on the conduction mechanism of such SPEs remain challenging as there are many influential factors behind it. Other than the abovementioned ceramic fillers, Li6.4La3Zr2Al0.2O12 and LiSnZr(PO4)3 have also been incorporated into PEO-LiTFSI [165] and PVDF–LiTFSI [166] electrolytes, respectively.
Another category of active fillers in CSPEs are the metal organic frameworks (MOFs) which display excellent properties such as microporous structures with high surface area, good mechanical and thermal properties, and controllable pore structure. MOFs have been credited to an increased lithium-ion transference number, ionic conductivity and improve the cycling stability of batteries [167]. Several examples of MOFs as active fillers in CSPEs for battery application include aluminum(III)-1,3,5-benzenetricarboxylate (Al-BTC) [168], D-UiO-66-NH2 [169], copper-1,4-benzenedicarboxylate (Cu-BDC) [170], HKUST-1(Cu) [171] and magnesium-benzene tricarboxylate (Mg-BTC) [172] to name a few. Overall, MOFs having 3D framework and controllable pore structures improve the interfacial contact between electrode and electrolyte. This is reflected with the addition of Zn4O(1,4-benzenedicarboxylate) in PEO-LiTFSI electrolyte which stabilized the interfacial resistance and improved the cyclability up to 100 cycles [173]. As usual, its ionic conductivity was enhanced too [173]. CSPE having Cu-BDC showed enhanced thermal stability which led to safer battery at high temperatures as compared to filler-free SPE [170]. More detailed reviews on CSPEs are available elsewhere [151,174,175,176]. Some other CSPEs systems that have been reported by many researchers are tabulated in Table 7.

3.2.7. SPEs Containing Other Kind of Additives

It is an undeniable fact that ionic conductivity plays a crucial role in ensuring the performance of electrochemical cells, but the cationic transference number also influences the efficacy of devices in particular batteries. Enhancement in lithium-ion transference number has been observed in PEO-LiI SPEs upon the addition of calix [4] arene with values around 0.8 to 1 at elevated temperatures between 55 and 90 °C [189]. In most cases, the cationic transference number values are found to be below 0.4 without supramolecular additives in PEO-salt SPE systems as mentioned earlier. Based on ab initio computational work, Johansson [190] predicted that the cryptands, i.e., organic compounds having large 3D molecular structure, can trap anions, such as Cl, ClO4, BF4 and F, and form coordination complexes, thereby reducing their tendency to re-associate with Li+ cation and consequently increasing the cationic transference number. However, the supramolecular compounds are unable to confine large PF6, TFSI or [(CF3SO2)2N] anions within a 3D network and cannot form complexes [190].
Mazor et al. [191] investigated the influence of calix[6]pyrrole on PEO-LiCF3SO3 SPEs. Although the ambient conductivity was found to decrease with the addition of calix, the conductivity at 60 °C was higher than that without calix. Nevertheless, the tLi+ of the calix-added SPE enhanced by 185% to 0.74 from 0.26 of the calix-free SPE at 60 °C [191]. Ambient conductivity decrement is said to be attributed to the increased rigidity of polymer network, whereas an improvement in lithium-ion transference number is due to increased amount of free ions from dissociated ion aggregates. The SPE has been employed in Li/MoOxSy cells that exhibited a power level of 2.1 mW cm−2 at 90 °C and a discharge capacity of ~24 μAh on the 60th cycle [191]. It is said that calix[6]arene can minimize the growth of lithium dendrites and subsequently improves cell performance [192]. A similar observation wherein an increased lithium-ion transference number (tLi+ = 0.95 at 70 °C) was accompanied by conductivity decrement has been reported in PEO-LiBF4-calix[6]pyrrole system [193]. In addition, the lithium diffusion-coefficient values were 2.0 × 10−7 cm2 s−1 and 2.5 × 10−7 cm2 s−1 for calix-free and calix-incorporated PEO-based electrolytes, respectively [193].
As aforementioned, the incorporation of supramolecular additives, such as calix-based compounds, is used to increase cationic transference number by trapping/immobilizing anion of the salt [194]. Usually, this will lead to reduce the anionic mobility and unwittingly decrease the ionic conductivity, as stated in preceding paragraph. Nevertheless, there are some controversial reports that showed otherwise. Won and co-workers [195] observed that the SPE containing PVC doped with lithium chloride (LiCl) salt and calix[4]pyrrole exhibited higher ionic conductivity, at around 10−3 S cm−1 at 25 °C, compared to that calix-free SPE (10−5 S cm−1). Using density functional theory, they obtained that the dissociation energies for LiCl-calix and PVC-LiCl-calix complexes were 110.8 and −18.1 kcal mol−1, whereas, for LiCl salt, the dissociation energy value was 145.6 kcal mol−1 [195]. A lower dissociation energy value implies that the cation of salt is more free to move, while its anion has complexed with calix [195].
Plastic crystal SN has also been employed as a polymer matrix in electrolyte, besides serving as a plasticizer and ionizer for both Li- and Na-based salts [131,196]. In order to improve the transference number of Na+, Chen et al. [196] synthesized boron-containing 2-hydroxyethyl methacrylate (B-HEMA) before incorporating it into SN-based polymer electrolyte that had non-woven polypropylene-cellulose composite framework and NaClO4 salt. This electrolyte having B- (boron anion) acceptor was prepared via a UV-assisted curing method, using 2-hydroxy-2-methyl-1-phenyl-1-propanone as photo-initiator. This electrolyte demonstrated superior properties in terms of ambient conductivity (3.60 × 10−4 S cm−1), tensile strength (28.2 MPa) and Na+ transference number of (tNa+ = 0.62) [196]. Promising results were also obtained after applying the electrolyte in Na-ion batteries (initial capacity of ~105 mAh g−1 at 0.1 C and retaining 80% capacity after 120 cycles) [196]. The boron moiety is said to complex with ClO4- ion of the salt, thus lowering the anionic mobility and subsequently enhancing the Na+ ion transference number. Meanwhile, the authors have stated that the polypropylene–cellulose structure did not contribute in ionic conduction, but instead improved the mechanical property of the plastic crystal polymer electrolyte [196].
Polyaniline (PANI), which is a conducting polymer that is usually used as counter electrode in DSSCs [197,198], has been incorporated in SPE containing PEO, tetrapropylammonium iodide (TPAI) and I2 to trap the large cations ((C3H7)4N+ or simply TPA+) so that the anions can contribute more to conduction, which is beneficial in DSSC [199]. The authors also believed that PANI has performed the role of plasticizing agent in decreasing crystallinity of PEO and its melting point based on DSC analysis. Upon PANI addition, the degrees of crystallinity and Tm were reduced to 3.37% and 66.2 °C from 8.93% and 66.5 °C [199]. The PANI-added SPE attained higher values of room temperature conductivity (8.61 × 10−5 S cm−1) and ionic transference number (0.99) than PANI-free SPE (σ = 6.81 × 10−5 S cm−1; tion= 0.96). The former also gave better DSSC performance (5.01% efficiency) than the latter (efficiency 3.5%) [199].
A special additive, i.e., hydroxypropyl trimethylammonium bis(trifluoromethane) sulfonamide chitosan salt (HACC-TFSI), was synthesized via ion-exchange route, using hydroxypropyl trimethylammonium chloride chitosan salt and LiTFSI before HACC-TFSI, PEO and LiTFSI were mixed in DMF solvent to form SPE via the solution-casting technique [200]. Ionic conductivity values for HACC-TFSI-added PEO electrolyte at 30 and 60 °C were 1.77 × 10−5 S cm−1 and 5.01 × 10−4 S cm−1, respectively [200]. In contrast, conductivities of 3.80 × 10−6 S cm−1 and 1.90 × 10−4 S cm−1 were recorded at the same temperature by PEO SPE without HACC-TFSI [200]. Besides better conductivity, Li+ ion transference number and tensile strength increased by 1.7 and 3.6 times. Moreover, the SPE with HACC-TFSI was found to be resistant against elevated temperature (150 °C) by retaining its shape after exposure at high temperature for one hour [200]. On the contrary, the electrolyte without HACC-TFSI has melted under similar conditions. LiFePO4/Li cell with HACC-TFSI added SPE demonstrated good cyclability and stability over 100 cycles maintaining 73% capacity of its initial capacity of ~108 mAh g−1 [200]. It is worthy to mention that this is considered quite an achievement, since there are not many cells with SPEs that can withstand and operate under such a high temperature.

4. Summary and Outlook

Solid-state electrochemical devices are the focus for the next-generation energy sources, especially for large power-consumption machines, such as electric vehicles. SPEs in electrochemical devices offer flexibility, low cost, harmless and cell cycling improvement. Basic SPEs which contain polymer and salt typically exhibit the ambient ionic conductivity of ~10−5 S cm−1. The conductivity must be improved at least to 10−3 S cm−1 for application in devices. Additives such as plasticizers and fillers are able to improve the conductivity of the SPEs. Plasticizers are able to reduce the crystallinity of the SPEs and enhance the elasticity and the segmental motion of the polymer chain, leading to an increase the overall ionic conductivity. Fillers, on the other hand, contribute either directly (active) or indirectly (inactive) towards conductivity enhancement. In contrast to SPEs with the plasticizers, fillers are able to improve the mechanical strength of the SPEs. Blending polymers is another approach that is used specially to improve the mechanical strength, apart from conductivity enhancement. Although ionic conductivity is one significant factor determining the feasibility of SPEs in electrochemical devices, other factors, such as ionic transport and transference number of ions, also influence the performance. It is appreciable that various efforts have been undertaken in developing suitable SPEs for practical use in electrochemical devices. In our opinion, there is a promising future for SPEs to compete and substitute liquid electrolytes in electrochemical devices. Among the abovementioned approaches, we think that there is much room for developments and improvements on SPEs having active fillers, in particular, MOFs, since research on this is relatively new and less than the others. Moreover, so far, the research only focuses on MOF-containing SPEs in application for Li-ion batteries. We believe that SPEs containing MOFs as fillers have the potential to be applied in other electrochemical devices, including Li-air batteries, DSSCs, PEMFCs and supercapacitors.

Author Contributions

Conceptualization, L.P.T. and M.H.B.; validation, A.K.A.; formal analysis, L.P.T. and M.H.B.; writing—original draft preparation, L.P.T. and M.H.B.; writing—review and editing, A.K.A.; supervision, A.K.A.; funding acquisition, M.H.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

M.H. Buraidah would like to thank Malaysia Ministry of Higher Education for the FRGS grant (No. FRGS/1/2019/STG07/UM/02/3).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fenton, D.E.; Parker, J.M.; Wright, P.V. Complexes of alkali metal ions with poly (ethylene oxide). Polymer 1973, 14, 589. [Google Scholar] [CrossRef]
  2. Armand, M.B.; Gabano, J.M.; Duclot, M. Second International Meeting on Solid Electrolytes, St. Andrews, Scotland, UK, 20–22 September 1978; Extended Abstract No. 6.5 in Fast Ion Transport in, Solids; Vashista, P., Mundy, J.N., Shenoy, G.K., Eds.; Elsevier: Amsterdam, North Holland, The Netherlands, 1979; p. 131. [Google Scholar]
  3. Voropaeva, D.Y.; Novikova, S.A.; Yaroslavtsev, A.B. Polymer electrolytes for metal-ion batteries. Russ. Chem. Rev. 2020, 89, 1132–1155. [Google Scholar] [CrossRef]
  4. Wu, H.; Wick, C.D. Computational Investigation on the Role of Plasticizers on Ion Conductivity in Poly(ethylene Oxide) LiTFSI Electrolytes. Macromolecules 2010, 43, 3502–3510. [Google Scholar] [CrossRef]
  5. Webb, M.A.; Savoie, B.M.; Wang, Z.-G.; Miller III, T.F. Chemically Specific Dynamic Bond Percolation Model for Ion Transport in Polymer Electrolytes. Macromolecules 2015, 48, 7346–7358. [Google Scholar] [CrossRef] [Green Version]
  6. Duarte, P.; Pereira, S.; Cunha, I.; Pimentel, A.; Dionísio, M.; Fortunato, E.; Martins, R.; Pereira, L. Cellulose-based solid electrolyte membranes through microwave assisted regeneration and application in electrochromic displays. Front. Mater. 2020, 7, 269. [Google Scholar] [CrossRef]
  7. Wu, Y.; Li, Y.; Wang, Y.; Liu, Q.; Chen, Q.; Chen, M. Advances and prospects of PVDF based polymer electrolytes. J. Energy Chem. 2022, 64, 62–84. [Google Scholar] [CrossRef]
  8. Arof, A.K.; Naeem, M.; Hameed, F.; Jayasundara, W.J.M.J.S.R.; Careem, M.A.; Teo, L.P.; Buraidah, M.H. Quasi solid state dye-sensitized solar cells based on polyvinyl alcohol (PVA) electrolytes containing I/I3 redox couple. Opt. Quant. Electron. 2014, 46, 143–154. [Google Scholar] [CrossRef]
  9. Arof, A.K.; Amirudin, S.; Yusof, S.Z.; Noor, I.M. A method based on impedance spectroscopy to determine transport properties of polymer electrolytes. Phys. Chem. Chem. Phys. 2014, 16, 1856–1867. [Google Scholar] [CrossRef] [Green Version]
  10. Buraidah, M.H.; Teo, L.P.; Majid, S.R.; Arof, A.K. Ionic conductivity by correlated barrier hopping in NH4I doped chitosan solid electrolyte. Phys. B 2009, 404, 1373–1379. [Google Scholar] [CrossRef]
  11. Kurapati, S.; Gunturi, S.S.; Nadella, K.J.; Erothu, H. Novel solid polymer electrolyte based on PMMA:CH3COOLi effect of salt concentration on optical and conductivity studies. Polym. Bull. 2019, 76, 5463–5481. [Google Scholar] [CrossRef]
  12. Teoh, K.H.; Lim, C.-S.; Ramesh, S. Lithium ion conduction in corn starch based solid polymer electrolytes. Measurement 2014, 48, 87–95. [Google Scholar] [CrossRef]
  13. Li, J.; Zhu, L.; Zhang, J.; Jing, M.; Yao, S.; Shen, X.; Li, S.; Tu, F. Approaching high performance PVDF-HFP based solid composite electrolytes with LLTO nanorods for solid-state lithium-ion batteries. Int. J. Energy Res. 2021, 45, 7663–7674. [Google Scholar] [CrossRef]
  14. Golodnitsky, D.; Strauss, E.; Peled, E.; Greenbaum, S. Review-On order and disorder in polymer electrolytes. J. Electrochem. Soc. 2015, 162, A2551–A2566. [Google Scholar] [CrossRef]
  15. Noor, I.M. Determination of charge carrier transport properties of gellan gum–lithium triflate solid polymer electrolyte from vibrational spectroscopy. High Perform. Polym. 2020, 32, 168–174. [Google Scholar] [CrossRef]
  16. Ahmed, H.T.; Abdullah, O.G. Impedance and ionic transport properties of proton-conducting electrolytes based on polyethylene oxide/methylcellulose blend polymers. J. Sci.-Adv. Mater. Dev. 2020, 5, 125–133. [Google Scholar] [CrossRef]
  17. Iftikhar, H.; Sonai, G.G.; Hashmi, S.G.; Nogueira, A.F.; Lund, P.D. Progress on electrolytes development in dye-sensitized solar cells. Materials 2019, 12, 1998. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Aziz, S.B.; Brza, M.A.; Brevik, I.; Hafiz, M.H.; Asnawi, A.S.F.M.; Yusof, Y.M.; Abdulwahid, R.T.; Kadir, M.F.Z. Blending and Characteristics of Electrochemical Double-Layer Capacitor Device Assembled from Plasticized Proton Ion Conducting Chitosan:Dextran:NH4PF6 Polymer Electrolytes. Polymers 2020, 12, 2103. [Google Scholar] [CrossRef] [PubMed]
  19. Stephan, A.M. Review on gel polymer electrolytes for lithium batteries. Eur. Polym. J. 2006, 42, 21–42. [Google Scholar] [CrossRef]
  20. Stallworth, P.E.; Li, J.; Greenbaum, S.G.; Croce, F.; Slane, S.; Salomon, M. Sodium-23 NMR and complex impedance studies of gel electrolytes based on poly(acrylonitrile). Solid State Ion. 1994, 73, 119–126. [Google Scholar] [CrossRef]
  21. Lewandowska, K. Miscibility and thermal stability of poly(vinyl alcohol)/chitosan mixtures. Thermochim. Acta 2009, 493, 42–48. [Google Scholar] [CrossRef]
  22. Gómez-Carracedo, A.; Alvarez-Lorenzo, C.; Gómez-Amoza, J.L.; Concheiro, A. Chemical structure and glass transition temperature of non-ionic cellulose ethers. J. Therm. Anal. Calorim. 2003, 73, 587–596. [Google Scholar] [CrossRef]
  23. Sivakumar, M.; Subadevi, R.; Rajendran, S.; Wu, H.-C.; Wu, N.-L. Compositional effect of PVdF-PEMA blend gel polymer electrolytes for lithium polymer batteries. Eur. Polym. J. 2007, 43, 4466–4473. [Google Scholar] [CrossRef]
  24. Sakurai, K.; Maegawa, T.; Takahashi, T. Glass transition temperature of chitosan and miscibility of chitosan/poly(N-vinyl pyrrolidone) blends. Polymer 2000, 41, 7051–7056. [Google Scholar] [CrossRef]
  25. Lu, D.R.; Xiao, C.M.; Xu, S.J. Starch-based completely biodegradable polymer materials. EXPRESS Polym. Lett. 2009, 3, 366–375. [Google Scholar] [CrossRef]
  26. Ue, M. Secondary batteries—Lithium rechargeable systems. Electrolytes: Nonaqueous. Encylcopedia Electrochem. Power Sources 2009, 71–84. [Google Scholar]
  27. Arof, A.K.; Teo, L.P.; Yahya, R. Ionic conductivity in some chitosan-based polymer electrolytes. Curr. Trends Polym. Sci. 2010, 14, 29–44. [Google Scholar]
  28. Izutsu, K. Electrochemistry in Nonaqueous Solutions; Wiley-VCH: Weinheim, Germany, 2002; pp. 3–24. [Google Scholar]
  29. Yu, Z.; Vlachopoulos, N.; Gorlov, M.; Kloo, L. Liquid electrolytes for dye-sensitized solar cells. Dalton Trans. 2011, 40, 10289–10303. [Google Scholar] [CrossRef]
  30. Fawcett, W.R. Acidity and Basicity Scales for Polar Solvents. J. Phys. Chem. 1993, 97, 9540–9546. [Google Scholar] [CrossRef]
  31. Solangi, M.Y.; Aftab, U.; Ishaque, M.; Bhutto, A.; Nafady, A.; Ibupoto, Z.H. Polyvinyl fibers as outperform candidature in the solid polymer electrolytes. J. Ind. Text. 2020, 1–13. [Google Scholar] [CrossRef]
  32. Vijayakumar, G.; Karthick, S.N.; Subramania, A. A new class of P(VdF-HFP)-CeO2-LiClO4 based composite microporous membrane electrolytes for Li-ion batteries. Int. J. Electrochem. 2011, 2011, 926383. [Google Scholar] [CrossRef] [Green Version]
  33. Röchow, E.T.; Coeler, M.; Pospiech, D.; Kobsch, O.; Mechtaeva, E.; Vogel, R.; Voit, B.; Nikolowski, K.; Wolter, M. In situ preparation of crosslinked polymer electrolytes for lithium ion batteries: A comparison of monomer systems. Polymers 2020, 12, 1707. [Google Scholar] [CrossRef]
  34. Liu, L.; Wang, Z.; Zhao, Z.; Zhao, Y.; Li, F.; Yang, L. PVDF/PAN/SiO2 polymer electrolyte membrane prepared by combination of phase inversion and chemical reaction method for lithium ion batteries. J. Solid State Electrochem. 2016, 20, 699–712. [Google Scholar] [CrossRef]
  35. Eriksson, T.; Mace, A.; Manabe, Y.; Yoshizawa-Fujita, M.; Inokuma, Y.; Brandell, D.; Jonas, M. Polyketones as Host Materials for Solid Polymer Electrolytes. J. Electrochem. Soc. 2020, 167, 070537. [Google Scholar] [CrossRef]
  36. Gupta, S.; Singh, P.K.; Bhattacharya, B. Charge carriers dynamics in PEO + NaSCN polymer electrolytes. Ionics 2018, 24, 163–167. [Google Scholar] [CrossRef]
  37. Ma, Q.; Liu, J.; Qi, X.; Rong, X.; Shao, Y.; Feng, W.; Nie, J.; Hu, Y.; Li, H.; Huang, X.; et al. New Na[(FSO2)(n-C4F9SO2)N]-Based Polymer Electrolyte for Solid-State Sodium Batteries. J. Mater. Chem. A 2017, 5, 7738–7743. [Google Scholar] [CrossRef]
  38. 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]
  39. Sim, L.N.; Sentanin, F.C.; Pawlicka, A.; Yahya, R.; Arof, A.K. Development of polyacrylonitrile-based polymer electrolytes incorporated with lithium bis(trifluoromethane)sulfonimide for application in electrochromic device. Electrochim. Acta 2017, 229, 22–30. [Google Scholar] [CrossRef]
  40. Sohaimy, M.I.H.A.; Isa, M.I.N. Natural Inspired Carboxymethyl Cellulose (CMC) Doped with Ammonium Carbonate (AC) as Biopolymer Electrolyte. Polymers 2020, 12, 2487. [Google Scholar] [CrossRef]
  41. Baharun, N.N.S.; Mingsukang, M.A.; Buraidah, M.H.; Woo, H.J.; Teo, L.P.; Arof, A.K. Development of solid polymer electrolytes based on sodium-carboxymethylcellulose (NaCMC)-polysulphide for quantum dot-sensitized solar cells (QDSSCs). Ionics 2020, 26, 1365–1378. [Google Scholar] [CrossRef]
  42. Perumal, P.; Selvasekarapandian, S.; Abhilash, K.P.; Sivaraj, P.; Hemalatha, R.; Selvin, P.C. Impact of lithium chlorate salts on structural and electrical properties of natural polymer electrolytes for all solid state lithium polymer batteries. Vacuum 2018, 159, 277–281. [Google Scholar] [CrossRef]
  43. 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]
  44. Singh, R.; Baghel, J.; Shukla, S.; Bhattacharya, B.; Rhee, H.-W.; Singh, P.K. Detailed electrical measurements on sago starch biopolymer solid electrolyte. Phase Transit. 2014, 87, 1237–1245. [Google Scholar] [CrossRef]
  45. Singh, R.; Singh, P.K.; Tomar, S.K.; Bhattacharya, B. Synthesis, characterization, and dye-sensitized solar cell fabrication using solid biopolymer electrolyte membranes. High Perform. Polym. 2016, 28, 47–54. [Google Scholar] [CrossRef]
  46. Hassan, M.F.; Azimi, N.S.N.; Kamarudin, K.H.; Sheng, C.K. Solid Polymer Electrolytes Based on Starch-Magnesium Sulphate: Study on Morphology and Electrical Conductivity. ASM Sci. J. 2018, 1, 17–28. [Google Scholar]
  47. Singh, V.K.; Bhattacharya, B.; Shukla, S.; Singh, P.K. Dye-sensitized solar cell comprising polyethyl methacrylate doped with ammonium iodide solid polymer electrolyte. Appl. Phys. A 2015, 118, 877–883. [Google Scholar] [CrossRef]
  48. Whba, R.A.G.; TianKhoon, L.; Su’ait, M.S.; Rahman, M.Y.A.; Ahmad, A. Influence of binary lithium salts on 49% poly (methyl methacrylate) grafted natural rubber based solid polymer electrolytes. Arab. J. Chem. 2020, 13, 3351–3361. [Google Scholar] [CrossRef]
  49. Ramesh, S.; Arof, A.K. Electrical conductivity studies of polyvinyl chloride-based electrolytes with double salt system. Solid State Ion. 2000, 136–137, 1197–1200. [Google Scholar] [CrossRef]
  50. Tao, R.; Fujinami, T. Application of mix-salts composed of lithium borate and lithium aluminate in PEO-based polymer electrolytes. J. Power Sources 2005, 146, 407–411. [Google Scholar] [CrossRef]
  51. Yang, H.; Farrington, G.C. Poly(ethylene oxide) electrolytes containing mixed salts. J. Polym. Sci. B Polym. Phys. 1993, 31, 157–163. [Google Scholar] [CrossRef]
  52. Moryoussef, A.; Bonat, M.; Fouletier, M.; Hicter, P. Proceedings, 6th Riso International Symposium on Metallurgy and Materials Science; Poulsen, F.W., Andersen, N.H., Clausen, K., Skaarup, S., Sorensen, O.T., Eds.; Riso National Lab.: Roskilde, Denmark, 1985; p. 335. [Google Scholar]
  53. Henderson, W.A.; Passerini, S.; Smyrl, W.H. Lithium Batteries: Proceedings of the International Symposium (Eds. S. Surampudi, R.A. Marsh, Z. Ogumi, J. Prakash), Mixed Salt Polymer Electrolytes—PEOn (x) LiCF3SO3 (1−x) LiClO4 (n = 12), Electrochemical Society Proceedings; The Electrochemical Society: Pennington, NJ, USA, 2000; Volume 99–25, pp. 515–523. [Google Scholar]
  54. Deepa, M.; Sharma, N.; Agnihotry, S.A.; Chandra, R.; Sekhon, S.S. Effect of mixed salts on the properties of gel polymeric electrolytes. Solid State Ion. 2002, 148, 451–455. [Google Scholar] [CrossRef]
  55. Giua, M.; Panero, S.; Scrosati, B.; Cao, X.; Greenbaum, S.G. Investigation of mixed cation effects in PEO9Zn1-xCux(CF3SO3)2 polymer electrolytes. Solid State Ion. 1996, 83, 73–78. [Google Scholar] [CrossRef]
  56. Dissanayake, M.A.K.L.; Jaseetharan, T.; Senadeera, G.K.R.; Mellander, B.-E.; Albinsson, I.; Furlani, M.; Kumari, J.M.K.W. Solid-state solar cells co-sensitized with PbS/CdS quantum dots and N719 dye and based on solid polymer electrolyte with binary cations and nanofillers. J. Photochem. Photobio. A Chem. 2021, 405, 112915. [Google Scholar] [CrossRef]
  57. Zhao, Q.; Chen, P.; Li, S.; Liu, X.; Archer, L.A. Solid-state Polymer Electrolytes Stabilized by Task-specific Salt Additives. J. Mater. Chem. A 2019, 7, 7823–7830. [Google Scholar] [CrossRef]
  58. Aziz, N.A.; Majid, S.R.; Arof, A.K. Synthesis and characterizations of phthaloyl chitosan-based polymer electrolytes. J. Non. Cryst. Solids 2012, 358, 1581–1590. [Google Scholar] [CrossRef]
  59. Romanazzi, G.; Gabler, F.M.; Margosan, D.; Mackey, B.E.; Smilanick, J.L. Effect of chitosan dissolved in different acids on its ability to control postharvest gray mold of table grape. Phytopathology 2009, 99, 1028–1036. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Demarger-Andre, S.; Domard, A. Chitosan carboxylic acid salts in solution and in the solid state. Carbohydr. Polym. 1994, 23, 211–219. [Google Scholar] [CrossRef]
  61. Kurita, K.; Ikeda, H.; Shimojoh, M.; Yang, J. N-Phthaloylated chitosan as an essential precursor for controlled chemical modifications of chitosan: Synthesis and evaluation. Polym. J. 2007, 39, 945–952. [Google Scholar] [CrossRef] [Green Version]
  62. Yoksan, R.; Akashi, M.; Biramontri, S.; Chirachanchai, S. Hydrophobic chain conjugation at hydroxyl group onto γ–ray irradiated chitosan. Biomacromolecules 2001, 2, 1038–1044. [Google Scholar] [CrossRef]
  63. Rahman, N.A.; Hanifah, S.A.; Mobarak, N.N.; Su’ait, M.S.; Ahmad, A.; Loh, K.S.; Lee, T.K. Synthesis and characterizations of o-nitrochitosan based biopolymer electrolyte for electrochemical devices. PLoS ONE 2019, 14, e0212066. [Google Scholar] [CrossRef]
  64. Jing, B.; Wang, X.; Shi, Y.; Zhu, Y.; Gao, H.; Fullerton-Shirey, S.K. Combining Hyperbranched and Linear Structures in Solid Polymer Electrolytes to Enhance Mechanical Properties and Room-Temperature Ion Transport. Front. Chem. 2021, 9, 563864. [Google Scholar] [CrossRef]
  65. Imperiyka, M.; Ahmad, A.; Hanifah, S.A.; Rahman, M.Y.A. Preparation and Characterization of Polymer Electrolyte of Glycidyl Methacrylate-Methyl Methacrylate-LiClO4 Plasticized with Ethylene Carbonate. Int. J. Polym. Sci. 2014, 2014, 63879. [Google Scholar] [CrossRef] [Green Version]
  66. Lehmann, M.L.; Yang, G.; Nanda, J.; Saito, T. Well-designed Crosslinked Polymer Electrolyte Enables High Ionic Conductivity and Enhanced Salt Solvation. J. Electrochem. Soc. 2020, 167, 070539. [Google Scholar] [CrossRef] [Green Version]
  67. Youcef, H.B.; Garcia-Calvo, O.; Lago, N.; Shanmukaraj, D.; Armand, M. Cross-Linked Solid Polymer Electrolyte for All-Solid-State Rechargeable Lithium Batteries. Electrochim. Acta 2016, 220, 587–594. [Google Scholar] [CrossRef]
  68. Porcarelli, L.; Gerbaldi, C.; Bella, F.; Nair, J.R. Super Soft All-Ethylene Oxide Polymer Electrolyte for Safe All Solid Lithium Batteries. Sci. Rep. 2016, 6, 19892. [Google Scholar] [CrossRef] [Green Version]
  69. Li, S.; Zuo, C.; Jo, Y.H.; Li, S.; Jiang, K.; Yu, L.; Zhang, Y.; Wang, J.; Li, L.; Xue, Z. Enhanced ionic conductivity and mechanical properties via dynamic-covalent boroxine bonds in solid polymer electrolytes. J. Membr. Sci. 2020, 608, 118218. [Google Scholar] [CrossRef]
  70. Lim, J.Y.; Kang, D.A.; Kim, N.U.; Lee, J.M.; Kim, J.H. Bicontinuously crosslinked polymer electrolyte membranes with high ion conductivity and mechanical strength. J. Membr. Sci. 2019, 589, 117250. [Google Scholar] [CrossRef]
  71. Vijayalekshmi, V.; Khastgir, D. Eco-friendly methanesulfonic acid and sodium salt of dodecylbenzene sulfonic acid doped cross-linked chitosan based green polymer electrolyte membranes for fuel cell applications. J. Membr. Sci. 2017, 523, 45–59. [Google Scholar] [CrossRef]
  72. Su’ait, M.S.; Ahmad, A.; Hamzah, H.; Rahman, M.Y.A. Preparation and characterization of PMMA-MG49-LiClO4 solid polymer electrolyte. J. Phys. D Appl. Phys. 2009, 42, 055410. [Google Scholar] [CrossRef]
  73. Su’ait, 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]
  74. Ali, A.M.M.; Subban, R.H.Y.; Bahron, H.; Yahya, M.Z.A.; Kamisan, A.S. Investigation on modified natural rubber gel polymer electrolytes for lithium polymer battery. J. Power Sources 2013, 244, 636–640. [Google Scholar] [CrossRef]
  75. Sutton, P.; Airoldi, M.; Porcarelli, L.; Olmedo-Martínez, J.L.; Mugemana, C.; Bruns, N.; Mecerreyes, D.; Ullrich, S.; Gunkel, I. Tuning the Properties of a UV-Polymerized, Cross-Linked Solid Polymer Electrolyte for Lithium Batteries. Polymers 2020, 12, 595. [Google Scholar] [CrossRef] [Green Version]
  76. Raghu, S.; Kilarkaje, S.; Sanjeev, G.; Nagaraja, G.K.; Devendrappa, H. Effect of electron beam irradiation on polymer electrolytes: Change in morphology, crystallinity, dielectric constant and AC conductivity with dose. Radiat. Phys. Chem. 2014, 98, 124–131. [Google Scholar] [CrossRef]
  77. Singh, D.; Singh, P.K.; Bhattacharya, B. Ion irradiation on polymer electrolyte films: Comparative study on conductivity. High Perform. Polym. 2016, 28, 1059–1063. [Google Scholar] [CrossRef]
  78. Chowdhury, F.I.; Khandaker, M.U.; Amin, Y.M.; Arof, A.K. Effect of gamma radiation on the transport and structural properties of polyacrylonitrile-lithium bis(oxalato)borate films. Solid State Ion. 2017, 304, 27–39. [Google Scholar] [CrossRef]
  79. Rahaman, M.H.A.; Khandaker, M.U.; Khan, Z.R.; Kufian, M.Z.; Noor, I.S.M.; Arof, A.K. Effect of gamma irradiation on poly(vinyledene difluoride)–lithium bis(oxalato)borate electrolyte. Phys. Chem. Chem. Phys. 2014, 16, 11527–11537. [Google Scholar] [CrossRef] [Green Version]
  80. Hema, M.; Tamilselvi, P.; Pandaram, P. Conductivity enhancement in SiO2 doped PVA:PVDF nanocomposite polymer electrolyte by gamma ray irradiation. Nucl. Instrum. Methods Phys. Res. B 2017, 403, 13–20. [Google Scholar] [CrossRef]
  81. Raghu, S.; Archana, K.; Sharanappa, C.; Ganesh, S.; Devendrappa, H. Electron beam and gamma ray irradiated polymer electrolyte films: Dielectric properties. J. Radiat. Res. Appl. Sci. 2016, 9, 117–124. [Google Scholar] [CrossRef] [Green Version]
  82. Heffny, N.A.; Khandaker, M.U.; Osman, Z.; Woo, H.J.; Rabir, M.H.; Arof, A.K. Effect of kGy neutron doses on polymer composite consists of poly(vinylidene difluoride)–lithium bis(oxalato)borate. Radiat. Phys. Chem. 2021, 189, 109747. [Google Scholar] [CrossRef]
  83. Sinha, M.; Goswami, M.M.; Mal, D.; Middya, T.R.; Tarafdar, S.; De, U.; Chaudhuri, S.K.; Das, D. Effect of gamma irradiation on the polymer electrolyte PEO-NH4ClO4. Ionics 2008, 14, 323–327. [Google Scholar] [CrossRef]
  84. Rajeswari, N.; Selvasekarapandian, S.; Prabu, M.; Karthikeyan, S.; Sanjeeviraja, C. Lithium ion conducting solid polymer blend electrolyte based on bio-degradable polymers. Bull. Mater. Sci. 2013, 36, 333–339. [Google Scholar] [CrossRef] [Green Version]
  85. Buraidah, M.H.; Arof, A.K. Characterization of chitosan/PVA blended electrolyte doped with NH4I. J. Non-Cryst. Solids 2011, 357, 3261–3266. [Google Scholar] [CrossRef]
  86. Buraidah, M.H.; Teo, L.P.; Au Yong, C.M.; Shah, S.; Arof, A.K. Performance of polymer electrolyte based on chitosan blended with poly(ethylene oxide) for plasmonic dye-sensitized solar cell. Opt. Mater. 2016, 57, 202–211. [Google Scholar] [CrossRef]
  87. Sim, L.N.; Majid, S.R.; Arof, A.K. FTIR studies of PEMA/PVdF-HFP blend polymer electrolyte system incorporated with LiCF3SO3 salt. Vib. Spectrosc. 2012, 58, 57–66. [Google Scholar] [CrossRef]
  88. Zhu, L.; Li, J.; Jia, Y.; Zhu, P.; Jing, M.; Yao, S.; Shen, X.; Li, S.; Tu, F. Toward high performance solid-state lithium-ion battery with a promising PEO/PPC blend solid polymer electrolyte. Int. J. Energy Res. 2020, 25, 10168–10178. [Google Scholar] [CrossRef]
  89. Kesavan, K.; Mathew, C.M.; Rajendran, S.; Subbu, C.; Ulaganathan, M. Solid Polymer Blend Electrolyte Based on Poly(ethylene oxide) and Poly(vinyl pyrrolidone) for Lithium Secondary Batteries. Braz. J. Phys. 2014, 45, 19–27. [Google Scholar] [CrossRef]
  90. Kesavan, K.; Mathew, C.M.; Rajendran, S.; Ulaganathan, M. Preparation and characterization of novel solid polymer blend electrolytes based on poly (vinyl pyrrolidone) with various concentrations of lithium perchlorate. Mater. Sci. Eng. B 2014, 184, 26–33. [Google Scholar] [CrossRef]
  91. Saadiah, M.A.; Nagao, Y.; Samsudin, A.S. Proton (H+) transport properties of CMC-PVA blended polymer solid electrolyte doped with NH4NO3. Inter. J. Hydrog. Energy 2020, 45, 14880–14896. [Google Scholar] [CrossRef]
  92. Balian, S.R.C.; Ahmad, A.; Mohamed, N.S. The Effect of Lithium Iodide to the Properties of Carboxymethyl κ-Carrageenan/Carboxymethyl Cellulose Polymer Electrolyte and Dye-Sensitized Solar Cell Performance. Polymers 2016, 8, 163. [Google Scholar] [CrossRef] [Green Version]
  93. Muhammad, F.H.; Subban, R.H.Y.; Winie, T. Solid solutions of hexanoyl chitosan/poly(vinyl chloride) blends and NaI for all-solid-state dye-sensitized solar cells. Ionics 2009, 25, 3373–3386. [Google Scholar] [CrossRef]
  94. Bao, J.; Qu, X.; Qi, G.; Huang, Q.; Wu, S.; Tao, C.; Gao, M.; Chen, C. Solid electrolyte based on waterborne polyurethane and poly(ethylene oxide) blend polymer for all-solid-state lithium ion batteries. Solid State Ion. 2018, 320, 55–63. [Google Scholar] [CrossRef]
  95. Tao, C.; Gao, M.-H.; Yin, B.-H.; Lia, B.; Huang, Y.-P.; Xua, G.; Bao, J.-J. A promising TPU/PEO blend polymer electrolyte for all-solid-state lithium ion batteries. Electrochim. Acta 2017, 257, 31–39. [Google Scholar] [CrossRef]
  96. Koduru, H.K.; Marino, L.; Scarpelli, F.; Petrov, A.G.; Marinov, Y.G.; Hadjichristov, G.B.; Iliev, M.T.; Scaramuzza, N. Structural and dielectric properties of NaIO4 e Complexed PEO/PVP blended solid polymer electrolytes. Curr. Appl. Phys. 2017, 17, 1518–1531. [Google Scholar] [CrossRef]
  97. Anilkumar, K.M.; Jinisha, B.; Manoj, M.; Jayalekshmi, S. Poly(ethylene oxide) (PEO)—Poly(vinyl pyrrolidone) (PVP) blend polymer based solid electrolyte membranes for developing solid state magnesium ion cells. Eur. Polym. J. 2017, 89, 249–262. [Google Scholar] [CrossRef]
  98. Senthil, R.A.; Theerthagiri, J.; Madhavan, J.; Arof, A.K. Performance characteristics of guanine incorporated PVDF-HFP/PEO polymer blend electrolytes with binary iodide salts for dye-sensitized solar cells. Opt. Mater. 2016, 58, 357–364. [Google Scholar] [CrossRef]
  99. Chandra, M.V.L.; Karthikeyan, S.; Selvasekarapandian, S.; Premalatha, M.; Monisha, S. Study of PVAc-PMMA-LiCl polymer blend electrolyte and the effect of plasticizer ethylene carbonate and nanofiller titania on PVAc-PMMA-LiCl polymer blend electrolyte. J. Polym. Eng. 2017, 37, 617–631. [Google Scholar] [CrossRef]
  100. Sengwa, R.J.; Choudhary, S.; Dhatarwal, P. Influences of ultrasonic- and microwave-irradiated preparation methods on the structural and dielectric properties of (PEO–PMMA)–LiCF3SO3–x wt% MMT nanocomposite electrolytes. Ionics 2014, 21, 95–109. [Google Scholar] [CrossRef]
  101. Guo, M.; Zhang, M.; He, D.; Hua, J.; Wang, X.; Gong, C.; Xie, X.; Xue, Z. Comb-like solid polymer electrolyte based on polyethylene glycol-grafted sulfonated polyether ether ketone. Electrochim. Acta 2017, 255, 396–404. [Google Scholar] [CrossRef]
  102. Liu, T.-M.; Saikia, D.; Ho, S.-Y.; Chen, M.-C.; Kao, H.-M. High ion-conducting solid polymer electrolytes based on blending hybrids derived from monoamine and diamine polyethers for lithium solid-state batteries. RSC Adv. 2017, 7, 20373–20383. [Google Scholar] [CrossRef] [Green Version]
  103. Ibrahim, S.; Ahmad, R.; Johan, M.R. Conductivity and optical studies of plasticized solid polymer electrolytes doped with carbon nanotube. J. Lumin. 2012, 132, 147–152. [Google Scholar] [CrossRef]
  104. Imperiyka, M.; Ahmad, A.; Hanifah, S.A.; Mohamed, N.S.; Rahman, M.Y.A. Investigation of plasticized UV-curable glycidyl methacrylate based solid polymer electrolyte for photoelectrochemical cell (PEC) application. Int. J. Hydrog. Energy 2014, 39, 3018–3024. [Google Scholar] [CrossRef]
  105. Fan, L.-Z.; Wang, X.-L.; Long, F. All-solid-state polymer electrolyte with plastic crystal materials for rechargeable lithium-ion battery. J. Power Sources 2009, 189, 775–778. [Google Scholar] [CrossRef]
  106. Machado, G.O.; Ferreira, H.C.A.; Pawlicka, A. Influence of plasticizer contents on the properties of HEC-based solid polymeric electrolytes. Electrochim. Acta 2005, 50, 3827–3831. [Google Scholar]
  107. Kim, J.S.; Lim, J.K.; Park, J.S. Enhancement of Mechanical Stability and Ionic Conductivity of Chitosan-based Solid Polymer Electrolytes Using Silver Nanowires as Fillers. Bull. Korean Chem. Soc. 2019, 40, 898–905. [Google Scholar] [CrossRef]
  108. Aziz, S.B.; Asnawi, A.S.F.M.; Mohammed, P.A.; Abdulwahid, R.T.; Yuhanees, M.; Yusof, Y.M.; Abdullah, R.M.; Kadir, M.F.Z. Impedance, circuit simulation, transport properties and energy storage behavior of plasticized lithium ion conducting chitosan based polymer electrolytes. Polym. Test. 2021, 101, 107286. [Google Scholar] [CrossRef]
  109. Li, Y.; Wang, J.; Tang, J.; Liu, Y.; He, Y. Conductive performances of solid polymer electrolyte films based on PVB/LiClO4 plasticized by PEG200, PEG400 and PEG600. J. Power Sources 2009, 187, 305–311. [Google Scholar] [CrossRef]
  110. Jibreel, U.M.; Bhattacharya, B.; Singh, P.K. Synthesis, Characterization, and Detailed Studies on Plasticized Poly(ethyl methacrylate): NH4I Polymer Electrolyte. Adv. Polym. Technol. 2018, 37, 542–546. [Google Scholar] [CrossRef]
  111. Anuar, N.K.; Subban, R.H.Y.; Mohamed, N.S. Properties of PEMA-NH4CF3SO3 added to BMATSFI ionic liquid. Materials 2012, 5, 2609–2620. [Google Scholar] [CrossRef]
  112. Wang, H.; Lin, C.; Yan, X.; Wu, A.; Shen, S.; Wei, G.; Zhang, J. Mechanical property-reinforced PEO/PVDF/LiClO4/SN blend all solid polymer electrolyte for lithium ion batteries. J. Electroanaly. Chem. 2020, 869, 114156. [Google Scholar] [CrossRef]
  113. Reddy, M.J.; Kumar, J.S.; Rao, U.V.S.; Chu, P.P. Structural and ionic conductivity of PEO blend PEG solid polymer electrolyte. Solid State Ion. 2006, 177, 253–256. [Google Scholar] [CrossRef]
  114. Widstrom, M.D.; Ludwig, K.B.; Matthews, J.E.; Jarry, A.; Erdi, M.; Cresce, A.V.; Rubloff, G.; Kofinas, P. Enabling high performance all-solid-state lithium metal batteries using solid polymer electrolytes plasticized with ionic liquid. Electrochim. Acta 2020, 345, 136–156. [Google Scholar] [CrossRef]
  115. Wang, W.; Fang, Z.; Zhao, M.; Peng, Y.; Zhang, J.; Guan, S. Solid polymer electrolytes based on the composite of PEO–LiFSI and organic ionic plastic crystal. Chem. Phys. Lett. 2020, 747, 137335. [Google Scholar] [CrossRef]
  116. Abdullah, O.G.; Ahmed, H.T.; Tahir, D.A.; Jamal, G.M.; Mohamad, A.H. Influence of PEG plasticizer content on the proton-conducting PEO:MC-NH4I blend polymer electrolytes based films. Results Phys. 2021, 23, 104073. [Google Scholar] [CrossRef]
  117. Ahmed, H.T.; Abdullah, O.G. Structural and ionic conductivity characterization of PEO:MC-NH4I proton-conducting polymer blend electrolytes based films. Results Phys. 2020, 16, 102861. [Google Scholar] [CrossRef]
  118. Pal, P.; Ghosh, A. Investigation of ionic conductivity and relaxation in plasticized PMMA-LiClO4 solid polymer electrolytes. Solid State Ion. 2018, 319, 117–124. [Google Scholar] [CrossRef]
  119. Rajendran, S.; Sivakumar, M.; Subadevi, R. Li-ion conduction of plasticized PVA solid polymer electrolytes complexed with various lithium salts. Solid State Ion. 2004, 167, 335–339. [Google Scholar] [CrossRef]
  120. Johan, M.R.; Oon, H.S.; Ibrahim, S.; Yassin, S.M.M.; Tay, Y.H. Effects of Al2O3 nanofiller and EC plasticizer on the ionic conductivity enhancement of solid PEO–LiCF3SO3 solid polymer electrolyte. Solid State Ion. 2011, 196, 41–47. [Google Scholar] [CrossRef]
  121. Liew, C.-W.; Ramesh, S.; Ramesh, K.; Arof, A.K. Preparation and characterization of lithium ion conducting ionic liquid-based biodegradable corn starch polymer electrolytes. J. Solid State Electrochem. 2012, 16, 1869–1875. [Google Scholar] [CrossRef]
  122. Gupta, S.; Varshney, P.K. Effect of plasticizer on the conductivity of carboxymethyl cellulose-based solid polymer electrolyte. Polym. Bull. 2019, 76, 6169–6178. [Google Scholar] [CrossRef]
  123. Liew, C.-W.; Arifin, K.H.; Kawamura, J.; Iwai, Y.; Ramesh, S.; Arof, A.K. Effects of halide anions in ionic liquid added poly(vinyl alcohol)-based ion conductors for electrical double layer capacitors. J. Non-Cryst. Solids 2017, 458, 97–106. [Google Scholar] [CrossRef]
  124. Buraidah, M.H.; Teo, L.P.; Arof, A.K. Determining the potential of 55 wt.% chitosan-45 wt.% NH4I biopolymer electrolyte for application in dye-sensitized solar cells. Mol. Cryst. Liq. Cryst. 2019, 695, 97–106. [Google Scholar] [CrossRef]
  125. Polu, A.R.; Singh, P.K. Improved ion dissociation and amorphous region of PEO based solid polymer electrolyte by incorporating tetracyanoethylene. Mat. Today Proc. 2020. [Google Scholar] [CrossRef]
  126. Singh, P.K.; Bhattacharya, B.; Nagarale, R.K.; Kim, K.-W.; Rhee, H.-W. Synthesis, characterization and application of biopolymer-ionic liquid composite membranes. Synth. Met. 2010, 160, 139–142. [Google Scholar] [CrossRef]
  127. Senthil, R.A.; Theerthagiri, J.; Madhavan, J. Optimization of performance characteristics of 2-mercaptopyridine-doped polyvinylidene fluoride (PVDF) polymer electrolytes for dye-sensitized solar cells. J. Non-Cryst. Solids 2014, 406, 133–138. [Google Scholar] [CrossRef]
  128. Senthil, R.A.; Theerthagiri, J.; Madhavan, J.; Ganesan, S.; Arof, A.K. Influence of organic additive to PVDF-HFP mixed iodide electrolytes on the photovoltaic performance of dye-sensitized solar cells. J. Phys. Chem. Solids 2017, 101, 18–24. [Google Scholar] [CrossRef]
  129. Tuhania, P.; Singh, P.K.; Bhattacharya, B.; Dhapola, P.S.; Yadav, S.; Shukla, P.K.; Gupta, M. PVDF-HFP and 1-ethyl-3-methylimidazolium thiocyanate–doped polymer electrolyte for efficient supercapacitors. High Perform. Polym. 2018, 30, 911–917. [Google Scholar] [CrossRef]
  130. Xue, Z.; He, D.; Xie, X. Poly(ethylene oxide)-based electrolytes for lithium-ion batteries. J. Mater. Chem. A 2015, 3, 19218–19253. [Google Scholar] [CrossRef]
  131. He, R.; Echeverri, M.; Ward, D.; Zhu, Y.; Kyu, T. Highly conductive solvent-free polymer electrolyte membrane for lithium-ion batteries: Effect of prepolymer molecular weight. J. Membr. Sci. 2016, 498, 208–217. [Google Scholar] [CrossRef]
  132. Tan, C.G.; Siew, W.O.; Pang, W.L.; Osman, Z.; Chew, K.W. The effects of ceramic fillers on the PMMA-based polymer electrolyte systems. Ionics 2007, 13, 361–364. [Google Scholar] [CrossRef]
  133. Ahmad, S.; Bohidar, H.B.; Ahmad, S.; Agnihotry, S.A. Role of fumed silica on ion conduction and rheology in nanocomposite polymeric electrolytes. Polymer 2006, 47, 3583–3590. [Google Scholar] [CrossRef]
  134. Wimalaweera, K.K.; Seneviratne, V.A.; Dissanayake, M.A.K.L. Effect of Al2O3 ceramic filler on thermal and transport properties of poly(ethylene oxide)-lithium perchlorate solid polymer electrolyte. Procedia Eng. 2017, 215, 109–114. [Google Scholar] [CrossRef]
  135. Croce, F.; Persi, L.; Scrosati, B.; Serraino-Fiory, F.; Plichta, E.; Hendrickson, M.A. Role of the ceramic fillers in enhancing the transport properties of composite polymer electrolytes. Electrochim. Acta 2001, 46, 2457–2461. [Google Scholar] [CrossRef]
  136. Singh, P.; Saroj, A.L. Effect of SiO2 Nano-particles on Plasticized Polymer Blend Electrolytes: Vibrational, Thermal, and Ionic Conductivity Study. Polym. Plast. Technol. Mater. 2021, 6, 298–305. [Google Scholar]
  137. Li, J.; Lian, K. The effect of SiO2 additives on solid hydroxide ion-conducting polymer electrolytes: A Raman microscopy study. Phys. Chem. Chem. Phys. 2018, 20, 7148–7155. [Google Scholar] [CrossRef]
  138. Kim, S.-H.; Choi, K.-H.; Cho, S.-J.; Kil, E.-H.; Lee, S.-Y. Mechanically compliant and lithium dendrite growth-suppressing composite polymer electrolytes for flexible lithium-ion batteries. J. Mater. Chem. A 2013, 1, 4949–4955. [Google Scholar] [CrossRef] [Green Version]
  139. Cui, J.; Zhou, Z.; Jia, M.; Chen, X.; Shi, C.; Zhao, N.; Guo, X. Solid polymer electrolytes with flexible framework of SiO2 nanofibers for highly safe solid lithium batteries. Polymers 2020, 12, 1324. [Google Scholar] [CrossRef] [PubMed]
  140. Jiang, Y.; Yan, X.; Ma, Z.; Mei, P.; Xiao, W.; You, Q.; Zhang, Y. Development of the PEO based solid polymer electrolytes for all-solid state lithium ion batteries. Polymers 2018, 10, 1237. [Google Scholar] [CrossRef] [Green Version]
  141. Jung, Y.-C.; Lee, S.-M.; Choi, J.-H.; Jang, S.S.; Kim, D.-W. All solid-state lithium batteries assembled with hybrid solid electrolytes. J. Electrochim. Soc. 2015, 162, A704–A710. [Google Scholar] [CrossRef]
  142. Wang, W.; Yi, F.; Fici, A.J.; Laine, R.M.; Kieffer, J. Lithium ion conducting poly(ethylene oxide)-based solid electrolytes containing active or passive ceramic nanoparticles. J. Phys. Chem. C 2017, 121, 2563–2573. [Google Scholar] [CrossRef]
  143. 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]
  144. Liu, W.; Liu, N.; Sun, J.; Hsu, P.-C.; Li, Y.; Lee, H.-W.; Cui, Y. Ionic conductivity enhancement of polymer electrolytes with ceramic nanowire fillers. Nano Lett. 2015, 15, 2740–2745. [Google Scholar] [CrossRef] [PubMed]
  145. Zhang, Q.; Liu, K.; Li, J.; Ma, C.; Zhou, L.; Du, Y. Study of a composite solid electrolyte made from a new pyrrolidone-containing polymer and LLZTO. J. Colloid Interface Sci. 2020, 580, 389–398. [Google Scholar] [CrossRef]
  146. Wang, H.; Zhao, N.; Bi, Z.; Gao, S.; Dai, Q.; Yang, T.; Wang, J.; Jia, Z.; Peng, Z.; Huang, J.; et al. Clear representation of surface pathway reactions at Ag nanowire cathodes in all-solid Li-O2 batteries. ACS Appl. Mater. Interfaces 2021, 13, 39157–39164. [Google Scholar] [CrossRef] [PubMed]
  147. Shen, L.; Wang, L.; Wang, Z.; Jin, C.; Peng, L.; Pan, X.; Sun, J.; Yang, R. Preparation and characterization of Ga and Sr co-doped Li7La3Zr2O12 garnet-type solid electrolyte. Solid State Ion. 2019, 339, 114992. [Google Scholar] [CrossRef]
  148. Huang, W.L.; Zhao, N.; Bi, Z.J.; Shi, C.; Guo, X.X.; Fan, L.-Z.; Nan, C.-W. Can we find solution to eliminate Li penetration through solid garnet electrolytes? Mater. Today Nano 2020, 10, 100075. [Google Scholar] [CrossRef]
  149. Ren, Y.; Shen, Y.; Lin, Y.; Nan, C.-W. Direct observation of lithium dendrites inside garnet-type lithium-ion solid electrolyte. Electrochem. Commun. 2015, 57, 27–30. [Google Scholar] [CrossRef]
  150. Shen, F.; Guo, W.; Zeng, D.; Sun, Z.; Gao, J.; Li, J.; He, B.; Han, X. A simple and highly efficient method toward high-density garnet-type LLZTO solid-state electrolyte. ACS Appl. Mater. Interfaces 2020, 12, 30313–30319. [Google Scholar] [CrossRef] [PubMed]
  151. Karabelli, D.; Birke, K.P.; Weeber, M. A performance and cost overview of selected solid-state electrolytes: Race between polymer electrolytes and inorganic sulfide electrolytes. Batteries 2021, 7, 18. [Google Scholar] [CrossRef]
  152. Nguyen, Q.H.; Luu, V.T.; Nguyen, H.L.; Lee, Y.-W.; Cho, Y.; Kim, S.Y.; Jun, Y.-S.; Ahn, W. Li7La3Zr2O12 garnet solid polymer electrolyte for highly stable all-solid-state batteries. Front. Chem. 2021, 8, 619832. [Google Scholar] [CrossRef]
  153. Zhang, W.; Wang, X.; Zhang, Q.; Wang, L.; Xu, Z.; Li, Y.; Huang, S. Li7La3Zr2O12 ceramic nanofiber-incorporated solid polymer electrolytes for flexible lithium batteries. ACS Appl. Energy Mater. 2020, 3, 5238–5246. [Google Scholar] [CrossRef]
  154. Cha, J.H.; Didwal, P.N.; Kim, J.M.; Chang, D.R.; Park, C.-J. Poly(ethylene oxide)-based composite solid polymer electrolyte containing Li7La3Zr2O12 and poly(ethylene glycol) dimethyl ether. J. Membr. Sci. 2020, 595, 117538. [Google Scholar] [CrossRef]
  155. Keller, M.; Appetecchi, G.B.; Kim, G.-T.; Sharova, V.; Schneider, M.; Schuhmacher, J.; Roters, A.; Passerini, S. Electrochemical performance of a solvent-free hybrid ceramic polymer electrolyte based on Li7La3Zr2O12 in P(EO)15LiTFSI. J. Power Sources 2017, 353, 287–297. [Google Scholar] [CrossRef]
  156. Zhou, D.; Zhang, M.; Su, F.; Arlt, T.; Frerichs, J.E.; Dong, K.; Wang, J.; Hilger, A.; Wilde, F.; Kolek, M.; et al. Performance and beghavior of LLZO-based composite polymer electrolyte for lithium metal electrode with high capacity utilization. Nano Energy 2020, 77, 105196. [Google Scholar] [CrossRef]
  157. Zhang, X.; Xu, B.-Q.; Lin, Y.-H.; Shen, Y.; Li, L.; Nan, C.-W. Effects of Li6.75La3Zr1.75Ta0.25O12 on chemical and electrochemical properties of polyacrylonitrile-based solid electrolytes. Solid State Ion. 2018, 327, 32–38. [Google Scholar] [CrossRef]
  158. Zhang, J.; Zang, X.; Wen, H.; Dong, T.; Chai, J.; Li, Y.; Chen, B.; Zhao, J.; Dong, S.; Ma, J.; et al. High-voltage and free-standing poly(propylene carbonate)/Li6.75La3Zr1.75Ta0.25O12 composite solid electrolyte for wide temperature range and flexible solid lithium ion battery. J. Mater. Chem. A 2017, 5, 4940–4948. [Google Scholar] [CrossRef]
  159. Piana, G.; Bella, F.; Geobaldo, F.; Meligrana, G.; Gerbaldi, C. PEO/LAGP hybrid solid polymer electrolytes for ambient temperature lithium batteries by solvent-free, “one-pot” preparation. J. Energy Storage 2019, 26, 100947. [Google Scholar] [CrossRef]
  160. Sung, B.-J.; Didwal, P.N.; Verma, R.; Nguyen, A.-G.; Chang, D.R.; Park, C.-J. Composite solid electrolyte comprising poly(propylene carbonate) and Li1.5Al0.5Ge1.5(PO4)3 for long-life all-solid-state Li-ion batteries. Electrochim. Acta 2021, 392, 139007. [Google Scholar] [CrossRef]
  161. Pignanelli, F.; Romero, M.; Esteves, M.; Luciana, F.-W.; Faccio, R.; Mombrú, A.W. Lithium titanate nanotubes as active fillers for lithium-ion polyacrylonitrile solid polymer electrolytes. Ionics 2019, 25, 2607–2614. [Google Scholar] [CrossRef]
  162. Boaretto, N.; Meabe, L.; Martinez-Ibanez, M.; Armand, M.; Zhang, H. Polymer electrolytes for rechargeable batteries: From nanocomposite to nanohybrid. J. Electrochem. Soc. 2020, 167, 070524. [Google Scholar] [CrossRef]
  163. Zheng, J.; Tang, M.; Hu, Y.-Y. Lithium ion pathway within Li7La3Zr2O12–polyethylene oxide composite electrolytes. Angew. Chem. 2016, 55, 12538–12542. [Google Scholar] [CrossRef] [PubMed]
  164. Zheng, J.; Dang, H.; Feng, X.; Chien, P.-H.; Hu, Y.-Y. Li-ion transport in a representative ceramic-polymer-plasticizer composite electrolyte: Li7La3Zr2O12–polyethylene oxide-tetraethylene glycol dimethyl ether. J. Mater. Chem. A. 2017, 5, 18457–18463. [Google Scholar] [CrossRef]
  165. Cai, D.; Wang, D.; Chen, Y.; Zhang, S.; Wang, X.; Xia, X.; Tu, J. A highly ion-conductive three-dimensional LLZAO-PEO/LiTFSI solid electrolyte for high-performance solid-state batteries. Chem. Eng. J. 2020, 394, 124993. [Google Scholar] [CrossRef]
  166. Pareek, T.; Dwivedi, S.; Ahmad, S.A.; Badole, M.; Kumar, S. Effect of NASICON-type LiSnZr(PO4)3 ceramic filler on the ionic conductivity and electrochemical behavior of PVDF based composite electrolyte. J. Alloys Compd. 2020, 824, 153991. [Google Scholar] [CrossRef]
  167. Zhao, R.; Wu, Y.; Liang, Z.; Gao, L.; Xia, W.; Zhao, Y.; Zou, R. Metal-organic frameworks for solid-state electrolytes. Energy Environ. Sci. 2020, 13, 2386–2403. [Google Scholar] [CrossRef]
  168. Gerbaldi, C.; Nair, J.R.; Kulandainathan, M.A.; Kumar, R.S.; Ferrara, C.; Mustarellic, P.; Stephan, A.M. Innovative high performing metal organic framework (MOF)-laden nanocomposite polymer electrolytes for all-solid-state lithium batteries. J. Mater. Chem. A 2014, 2, 9948–9954. [Google Scholar] [CrossRef]
  169. Angulakshmi, N.; Zhou, Y.; Suriyakumar, S.; Dhanalakshmi, R.B.; Satishrajan, M.; Alwarappan, S.; Alkordi, M.H.; Stephan, A.M. Microporous metal-organic framework (MOF)-based composite polymer electrolyte (CPE) mitigating lithium dendrite formation in all-solid-state-lithium batteries. ACS Omega 2020, 5, 7885–7894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Kumar, R.S.; Raja, M.; Kulandainathan, M.A.; Stephan, A.M. Metal organic framework-laden composite polymer electrolytes for efficient and durable all-solid-state-lithium batteries. RSC Adv. 2014, 4, 26171–26175. [Google Scholar] [CrossRef]
  171. Wang, Z.; Zhou, H.; Meng, C.; Xiong, W.; Cai, Y.; Hu, P.; Pang, H.; Yuan, A. Enhancing ion transport: Function of ionic liquid decorated MOFs in polymer electrolytes for all solid-state lithium batteries. ACS Appl. Energy Mater. 2020, 3, 4265–4274. [Google Scholar] [CrossRef]
  172. Angulakshmi, N.; Kumar, R.S.; Kulandainathan, M.A.; Stephan, A.M. Composite polymer electrolytes encompassing metal organic frameworks: A new strategy for all solid-state lithium batteries. J. Phys. Chem. C 2014, 118, 24240–24247. [Google Scholar] [CrossRef]
  173. Yuan, C.; Li, J.; Han, P.; Lai, Y.; Zhang, Z.; Liu, J. Enhanced electrochemical performance of poly(ethylene oxide) based composite polymer electrolyte by incorporation of nano-sized metal-organic framework. J. Power Sources 2013, 240, 653–658. [Google Scholar] [CrossRef]
  174. 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]
  175. Li, L.; Deng, Y.; Chen, G. Status and prospect of garnet/polymer solid composite electrolytes for all-solid-state lithium batteries. J. Energy Chem. 2020, 50, 154–177. [Google Scholar] [CrossRef]
  176. Barbosa, J.C.; Goncalves, R.; Costa, C.M.; Bermudez, V.d.Z.; Fidalgo-Marijuan, A.; Zhang, Q.; Lanceros-Mendez, S. Metal-organic frameworks and zeolite materials as active fillers for lithium-ion battery solid polymer electrolytes. Mater. Adv. 2021, 2, 3790–3805. [Google Scholar] [CrossRef]
  177. Capiglia, C.; Mustarelli, P.; Quartarone, E.; Tomasi, C.; Magistris, A. Effects of nanoscale SiO on the thermal and transport properties of solvent-free, poly(ethylene oxide) (PEO)-based polymer electrolytes. Solid State Ion. 1999, 118, 73–79. [Google Scholar] [CrossRef]
  178. Wei, W.; Li, L.; Zhang, L.; Hong, L.; He, G. An all-solid-state Li-organic battery with quinone-based polymer cathode and composite polymer electrolyte. Electrochem. Commun. 2018, 90, 21–25. [Google Scholar] [CrossRef]
  179. Jeyabanua, K.; Sundaramahalingam, K.; Devendran, P.; Manikandan, A.; Nallamuthu, N. Effect of electrical conductivity studies for CuS nanofillers mixed magnesium ion based PVA-PVP blend polymer solid electrolyte. Phys. B Condens. Matter. 2019, 572, 129–138. [Google Scholar] [CrossRef]
  180. Dhatarwala, P.; Choudhary, S.; Sengwa, R.J. Electrochemical performance of Li+-ion conducting solid polymer electrolytes based on PEO–PMMA blend matrix incorporated with various inorganic nanoparticles for the lithium ion batteries. Compos. Commun. 2018, 10, 11–17. [Google Scholar] [CrossRef]
  181. Choudhary, S.; Sengwa, R.J. Effects of different inorganic nanoparticles on the structural, dielectric and ion transportation properties of polymers blend based nanocomposite solid polymer electrolytes. Electrochim. Acta 2017, 247, 924–941. [Google Scholar] [CrossRef]
  182. Li, L.; Shan, Y.; Yang, X. New insights for constructing solid polymer electrolytes with ideal lithium-ion transfer channels by using inorganic filler. Mater. Today Commun. 2020, 26, 101910. [Google Scholar] [CrossRef]
  183. Dhatarwala, P.; Sengwa, R.J.; Choudhary, S. Effect of intercalated and exfoliated montmorillonite clay on the structural, dielectric and electrical properties of plasticized nanocomposite solid polymer electrolytes. Compos. Commun. 2017, 5, 1–7. [Google Scholar] [CrossRef]
  184. Zhang, X.; Wang, X.; Liu, S.; Tao, Z.; Chen, J. A novel PMA/PEG-based composite polymer electrolyte for all-solid-state sodium ion batteries. Nano Res. 2018, 11, 62244–66251. [Google Scholar] [CrossRef]
  185. Koduru, H.K.; Iliev, M.T.; Kondamareddy, K.K.; Karashanova, D.; Vlakhov, T.; Zhao, X.Z.; Scaramuzza, N. Investigations on Poly(ethylene oxide) (PEO)—blend based solid polymer electrolytes for sodium ion batteries. J. Phys. Conf. Ser. 2016, 764, 012006. [Google Scholar] [CrossRef]
  186. Hu, J.; Chen, K.; Zhenguo Yao, Z.; Li, C. Unlocking solid-state conversion batteries reinforced by hierarchical microsphere stacked polymer electrolyte. Sci. Bull. 2021, 66, 694–707. [Google Scholar] [CrossRef]
  187. Zhao, Y.; Wu, C.; Peng, G.; Chen, X.; Yao, X.; Bai, Y.; Wu, F.; Chen, S.; Xu, X. A new solid polymer electrolyte incorporating Li10GeP2S12 into a polyethylene oxide matrix for all-solid-state lithium batteries. J. Power Sources 2016, 301, 47–53. [Google Scholar] [CrossRef]
  188. Suriyakumar, S.; Kanagaraj, M.; Angulakshmi, N.; Kathiresan, M.; Nahm, K.S.; Walkowiak, M.; Wasinski, K.; Poltrolniczak, P.; Stephan, A.M. Charge-discharge studies of all-solid-state Li/LiFePO4 cells with PEO-based composite electrolytes encompassing metal organic frameworks. RSC Adv. 2016, 6, 97180–97186. [Google Scholar] [CrossRef]
  189. Blazejczyk, A.; Wieczorek, W.; Kovarsky, R.; Golodnitsky, D.; Peled, E.; Scanlon, L.G.; Appetecchi, G.B.; Scrosati, B. Novel Solid Polymer Electrolytes with Single Lithium-Ion Transport. J. Electrochem. Soc. 2004, 151, A1762–A1766. [Google Scholar] [CrossRef]
  190. Johansson, P. Anion cryptands: A way to increased cation transport in polymer electrolytes. Electrochim. Acta 2003, 48, 2291–2294. [Google Scholar] [CrossRef]
  191. Mazor, H.; Golodnitsky, D.; Peled, E.; Wieczorek, W.; Scrosati, B. A search for a single-ion-conducting polymer electrolyte: Combined effect of anion trap and inorganic filler. J. Power Sources 2008, 178, 736–743. [Google Scholar] [CrossRef]
  192. Kufian, M.Z.; Arof, A.K. Effect of calix[6]arene on charge discharge behaviour of lithium air cell based on PAN-LiBOB gel polymer electrolyte. Mater. Technol. 2014, 29, A114–A117. [Google Scholar] [CrossRef]
  193. Kalita, M.; Bukat, M.; Ciosek, M.; Siekierski, M.; Chung, S.H.; Rodr’ıguez, T.; Greenbaum, S.G.; Kovarsky, R.; Golodnitsky, D.; Peled, E.; et al. Effect of calixpyrrole in PEO–LiBF4 polymer electrolytes. Electrochim. Acta 2005, 50, 3942–3948. [Google Scholar] [CrossRef]
  194. Panero, S.; Scrosati, B.; Sumathipala, H.H.; Wieczorek, W. Dual-composite polymer electrolytes with enhanced transport properties. J. Power Sources 2007, 167, 510–514. [Google Scholar] [CrossRef]
  195. Won, J.; Lee, K.M.; Kang, Y.S.; Chang, S.-K.; Kim, C.K.; Kim, C.K. Anion complexation by calix[4]pyrrole in solid polymer electrolytes. Macromol. Res. 2006, 14, 404–407. [Google Scholar] [CrossRef]
  196. Chen, S.; Feng, F.; Yin, Y.; Lizo, X.; Ma, Z. Plastic crystal polymer electrolytes containing boron based anion acceptors for room temperature all-solid-state sodium-ion batteries. Energy Storage Mater. 2019, 22, 57–65. [Google Scholar] [CrossRef]
  197. Eslah, S.; Nouri, M. Synthesis and Characterization of tungsten trioxide/polyaniline/polyacrylonitrile composite nanofibers for application as a counter electrode of DSSCs. Russ. J. Electrochem. 2019, 55, 291–304. [Google Scholar] [CrossRef]
  198. Li, Q.; Wu, J.; Tang, Q.; Lan, Z.; Li, L.; Lin, J.; Fan, L. Application of microporous polyaniline counter electrode for dye-sensitized solar cells. Electrochem. Commun. 2008, 10, 1299–1302. [Google Scholar] [CrossRef]
  199. Dissanayake, M.A.K.L.; Kumari, J.M.K.W.; Senadeera, G.K.R.; Jaseetharan, T.; Mellander, B.-E.; Albinson, I.; Furlani, M. Polyaniline (PANI) mediated cation trapping effect on ionic conductivity enhancement in poly(ethylene oxide) based solid polymer electrolytes with application in solid state dye sensitized solar cells. React. Funct. Polym. 2020, 155, 104683. [Google Scholar] [CrossRef]
  200. Tan, J.; Ao, X.; Dai, A.; Yuan, Y.; Zhuo, H.; Lu, H.; Zhuang, L.; Ke, Y.; Su, C.; Peng, X.; et al. Polycation ionic liquid tailored PEO-based solid polymer electrolytes for high temperature lithium metal batteries. Energy Storage Mater. 2020, 33, 173–180. [Google Scholar] [CrossRef]
Table 1. Typical polymers used in polymer electrolytes, along with their characteristics.
Table 1. Typical polymers used in polymer electrolytes, along with their characteristics.
PolymerFunctional Group(s)Glass Transition Temperature, Tg (°C)Reference
Poly(ethylene oxide) (PEO)Ether−64 [19]
Polyacrylonitrile (PAN)Nitrile125[19,20]
Poly(vinyl alcohol) (PVA)Hydroxyl82[21]
CelluloseHydroxyl, Ether220[22]
Poly(methyl methacrylate) (PMMA)Ester, Carbonyl105[19]
Polyethyl methacrylate (PEMA)Ester, Carbonyl63[23]
Polyvinylidene fluoride (PVDF)Difluoromethylene−40[19]
ChitosanHydroxyl, Amine, Ether203[24]
Poly(vinylidene fluoride-hexafluoropropylene) (PVDF-HFP)Difluoromethylene−90[19]
StarchHydroxy, Ether70[25]
Table 2. Some popular salts used in polymer electrolytes, along with their properties [18,26,27].
Table 2. Some popular salts used in polymer electrolytes, along with their properties [18,26,27].
SaltCationic Radii
(pm)
Anionic Radii
(pm)
Lattice Energy
(kJ mol−1)
Lithium bis(trifluoromethanesulfonyl)imide ((LiN(CF3SO2)2) or (LiTFSI)60328564
Lithium trifluoromethanesulfonate or lithium triflate (LiCF3SO3)60268730
Lithium perchlorate (LiClO4)60234718
Lithium tetrafluoroborate (LiBF4)60227735
Lithium hexafluorophosphate (LiPF6)60254681
Sodium iodide (NaI)102216674
Sodium thiocyanate (NaSCN)102213680
Ammonium iodide (NH4I)143216605
Lithium iodide (LiI)60216762
Table 3. Physical properties of some solvents [28,29,30].
Table 3. Physical properties of some solvents [28,29,30].
SolventDonor Number, DNDielectric Constant, ε at 25 °CViscosity (cP) at 25 °C Melting Point, Tm (°C)Boiling Point (°C)
Acetone17.020.60.30−94.756.1
Acetonitrile (ACN)14.135.90.37−43.881.6
N-methyl-2-pyrrolidone (NMP)27.332.21.67−24.4202
Tetrahydrofuran (THF) 7.580.46−108.460
N,N-Dimethylacetamide (DMAc)27.837.80.93−20166.1
Dimethyl sulfoxide (DMSO)29.846.51.9918.5189
Gamma-butyrolactone (gBL)18.039.0 1.70−44204
N,N-Dimethylformamide (DMF)26.636.10.80−61153
Ethylene carbonate (EC)16.489 *1.90 *36238
Propylene carbonate (PC)15.1642.50−49241
Ethanol19.2251.08−11478
Water18.0780.890100
* At 40 °C. At 20 °C.
Table 4. Basic SPEs films, along with their conductivity and significant properties.
Table 4. Basic SPEs films, along with their conductivity and significant properties.
Electrolyte SystemPreparation MethodConductivity, σ (S cm−1)Important FindingsReference
PEO-NaSCNSolution casting9.86 × 10−8 (RT)tion = 0.85
n = 1.21 × 1020 m−3;
μ = 5.10 × 10−7 m2 V−1 s−1
[36]
PEO-NaFNFSI
[EO]/[Na]=15
Solution casting3.16 × 10−6 (80 °C)
3.36 × 10−4 (30 °C)
Thermal stability > 300 °C; Tg = −36.3 °C
tNa+ = 0.24 (80 °C)
Oxidation potential = 4.87 V vs Na+/Na
NaCu1/9Ni2/9Fe1/3Mn1/3O2/SPE/Na cell:
Initial capacity = 122.4 mAh g−1 (0.1 C, 80 °C)70% capacity retention at 1 C after 150 cycles (80 °C)
[37]
PEO-LiCF3SO3Ball-milling and hot press1.00 × 10−6Tg = −64.4 °C; %χC = 37.3%[38]
PAN-LiTFSISolution casting2.54 × 10−4 (25 °C)C = 9.3%
ITO-WO3/SPE/CeO2-TiO2-ITO ECD:
Cathodic coloration at −1.25 V;
Anodic bleaching at −0.40 V
[39]
CMC-(NH4)2CO3Solution casting7.71 × 10−6 (30 °C)C = 30.9%; tion = 0.98
CMC film without salt: σ = 10−9 S cm−1
[40]
NaCMC-Na2S-SSolution casting2.79 × 10−5 (30 °C)EA = 0.38 eV
n = 15.24 × 1018 cm−3;
μ = 11.44 × 10−6 m2 V−1 s−1,
D = 2.94 × 10−7 cm2 s−1
TiO2-CdS-ZnS/SPE/Pt QDSSC (light intensity of 100 mW cm−2):
OCV = 0.41 V; Jsc = 4.92 mA cm−2;
FF = 0.45; η = 0.90%
[41]
Pectin-LiClSolution casting1.96 × 10−3 (30 °C)EA = 0.23 eV
Tensile strength = 4.61 MPa
Electrochemical stability window 3.77 V
Zn/SPE/LiFePO4 cell:
OCV = 1.25 V
[42]
PEO-LiCF3SO3
[EO]/[Li] = 9:1
Solution casting1.40 × 10−6 (30 °C)Tg = −39 °C; Tm = 64 °C[43]
Sago starch-KISolution casting3.41 × 10−4 (RT)Dielectric relaxation time = 2.33 × 10−7 s
N3/TiO2/SPE/Pt DSSC (light intensity of 100 mW cm−2):
OCV = 0.58 V; Jsc = 0.29 mA cm−2;
FF = 0.43; η = 0.57%
[44,45]
Corn starch-MgSO4Solution casting8.52 × 10−5 (30 °C)SEM micrograph of SPE shows rough surface, while that of starch film without salt shows otherwise.[46]
PEMA-NH4ISolution casting1.80 × 10−5 (30 °C)tion = 0.93
Ru-based dye/TiO2/SPE/Pt DSSC (light intensity of 100 mW cm−2):
OCV = 0.56 V; Jsc = 1.52 mA cm−2;
FF = 0.51; η = 0.43%
[47]
NaSCN—sodium thiocyanate; RT—room temperature; NaFNFSI—sodium (fluorosulfonyl)(n-nonafluorobutanesulfonyl)imide (Na[(FSO2)(n-C4F9SO2)N]; %χC—degree of crystallinity percentage; CMC—carboxymethyl cellulose; ITO—indium tin oxide; WO3—tungsten (VI) oxide; CeO2—cerium (IV) oxide; TiO2—titanium dioxide; (NH4)2CO3—ammonium carbonate; NaCMC—sodium-carboxymethylcellulose; Na2S—sodium sulfide; S—sulfur; OCV—open circuit voltage; Jsc—short circuit current density; FF—fill factor; η—efficiency; CdS—cadmium sulfide; ZnS—zinc sulfide; Pt—platinum; LiCl—lithium chloride; LiFePO4—lithium iron phosphate; N3—cis-bis(isothiocyanato)bis(2,2′-bipyridyl-4,4′-dicarboxylato)ruthenium(II); MgSO4—magnesium sulfate.
Table 5. Various blended SPEs systems, along with their conductivity values and other important parameters.
Table 5. Various blended SPEs systems, along with their conductivity values and other important parameters.
Electrolyte SystemPreparation MethodConductivity, σ (S cm−1)Important FindingsReference
PVA:PVP (70:30)-25 wt.% LiNO3Solution casting6.83 × 10−4EA = 0.27 eV; Tg = 80 °C
70 PVA:30 PVP film:
σ = 1.58 × 10−6 S cm−1; Tg = 130 °C
Degree of swelling lower than blend SPE
[84]
Chitosan:PVA (1:1)-NH4ISolution casting1.77 × 10−6 (30 °C)EA = 0.38 eV
Without PVA blend:
σ = 3.73 × 10−7 S cm−1 (30 °C)
EA = 0.46 eV
[85]
Chitosan:PEO (1:1)-NH4ISolution casting3.66 × 10−6 (30 °C)EA = 0.31 eV; tion = 0.85
N3/Ag-TiO2/SPE/Pt DSSC (light intensity of 100 mW cm−2):
OCV = 0.58 V; Jsc = 2.84 mA cm−2;
FF = 0.69; η = 1.13%
Without PEO blend:
σ = 3.73 × 10−7 S cm−1 (30 °C);
EA = 0.46 eV
[86]
PEMA:PVDF-HFP (70:30)-LiCF3SO3Reflux + Solution casting2.87 × 10−7 (RT)PEMA-LiCF3SO3:
σ = 9.18 × 10−8 S cm−1 (RT)
[87]
50% (PEO+LiTFSI):50% PPC [EO]/[Li] = 18:1Solution casting2.04 × 10−5 (25 °C)
2.82 × 10−4 (60 °C)
EA = 0.78 eV; Tg = −41.3 °C;
Tm = 49.7 °C; %χC = 22.10%;
tLi+ = 0.184 (60 °C);
Electrochemical stability window 4.90 V
LiFePO4/SPE/Li cell (0.5C, 60 °C):
Discharge specific capacity of
112 mAh g−1 after 100th cycle
Without PPC blend:
σ = 1.02 × 10−5 S cm−1 (25 °C);
σ = 1.77 × 10−4 S cm−1 (60 °C);
EA = 0.80 eV; Tg = −35.2 °C;
Tm = 58.1 °C; %χC = 48.07%;
tLi+ = 0.156 (60 °C);
Electrochemical stability window 4.25 V
Discharge specific capacity = 93 mAh g−1 after 100th cycle (0.5 C, 60 °C)
Decreased crystallinity with PPC addition
Improved interface stability
[88]
PEO:PVP (9:1)-8 wt.% LiClO4Solution casting0.23 × 10−5 (30 °C)Bandgap energy = 4.07 eV (indirect) and 4.35 eV (direct)
Thermally stable till 324 °C
Lowest intensity in photoluminescence results compared to that at other PEO:PVP ratios
[89,90]
CMC:PVA:NH4NO3
(56:14:30 wt.%)
Solution casting1.70 × 10−3Tg = 88.7 °C; tH+ = 0.42
Decomposition temperature = 340 °C
[91]
MG49:PMMA (70:30)-25 wt.% LiBF4
[O]/[Li] = 5:1
Solution casting8.60 × 10−6 (30 °C)MG49-PMMA film:
σ = 1.10 × 10−12 S cm−1 (30 °C)
[73]
MG49:PMMA (70:30)-25 wt.% LiClO4
[O]/[Li] = 6:1
Solution casting1.50 × 10−8 (30 °C)MG49-PMMA film:
σ = 1.10 × 10−12 S cm−1 (30 °C)
[73]
CMKC:CMC (60:40)-LiISolution casting3.89 × 10−3 (27 °C)Tg = −43.0 °C
N719/TiO2/SPE/Pt DSSC (light intensity of 100 mW cm−2):
OCV = 0.492 V; Jsc = 0.40 mA cm−2;
FF = 0.57; η = 0.11%
[92]
HCh:PVC (90:10)-NaISolution casting1.50 × 10−5 (RT)C = 24%; D = 1.10 × 10−8 cm2 s−1;
μ = 4.10 × 10−7 cm2 V−1 s−1;
n = 2.30 × 1020 cm−3
N3/TiO2/SPE/Pt DSSC (light intensity of 100 mW cm−2):
OCV = 0.58 V; Jsc = 8.62 mA cm−2;
FF = 0.59; η = 2.93%
[93]
PEO:WPU (3:1)-LiTFSI
[EO]/[Li] = 16:1
Solution casting3.10 × 10−3 (80 °C)Tg = −45.4 °C; thermally stable till 300 °C
LiFePO4/SPE/Li battery:
Discharge capacity = 122 mAh g−1 after 100th cycle (1C, 80 °C), 96% capacity retention
[94]
PEO:TPU (3:1)-LiTFSI
[EO]/[Li] = 16:1
Solution casting5.30 × 10−4 (60 °C)tLi+ = 0.31; Tensile stress 1.38 MPa
LiFePO4/SPE/Li battery:
Discharge capacity = 112 mAh g−1 after 100th cycle (1 C, 60 °C), 96% capacity retention
[95]
PEO-PVP-NaIO4
(65:25:10 wt.%)
Solution casting1.56 × 10−7 (30 °C)EA = 0.22 eV; Tg = 50.2 °C; %χC = 55.0%
D = 2.72 × 10−8 cm2 s−1;
μ = 1.0 × 10−4 cm2 V−1 s−1;
n = 9.35 × 1015 cm−3
Direct optical bandgap = 3.60 eV
Refractive index = 2.28
[96]
PEO:PVP (4:1)-Mg(NO3)2Solution casting5.80 × 10−4 (25 °C)EA = 0.31 eV; t+ = 0.33
Thermally stable till 220 °C
MgMn2O4/SPE/Mg battery:
OCV = 1.46 V
[97]
PVDF-HFP:PEO (4:1)-KI-TBAI-I2Solution casting4.53 × 10−5 (25 °C)Tm = 144.9 °C; %χC = 18.1%
N3/TiO2/SPE/Pt DSSC (light intensity of 60 mW cm−2):
OCV = 0.674 V; Jsc = 4.39 mA cm−2;
FF = 0.50; η = 2.46%
Without PEO blend:
σ = 9.99 × 10−6 S cm−1 (25 °C)
Tm = 146.5 °C; %χC = 23.5%
OCV = 0.648 V; Jsc = 3.78 mA cm−2;
FF = 0.46; η = 1.88%
[98]
PVAc:PMMA (70:30)-LiClSolution casting1.03 × 10−5 (30 °C)EA = 0.25 eV; Tg = 42.6 °C; t+ = 0.97
D+ = 6.93 × 10−11 cm2 s−1;
μ = 2.65 × 10−9 cm2 V−1 s−1;
n = 2.35 × 1022 cm−3
[99]
PVP—poly(vinyl pyrrolidone); LiNO3—lithium nitrate; PPC—propylene carbonate; CMKC—carboxymethyl κ-carrageenan; N719—di-tetrabutylammonium cis-bis(isothiocyanato)bis(2,2′-bipyridyl-4,4′-dicarboxylato) ruthenium(II); PVC—poly(vinyl chloride); WPU—waterborne polyurethane; TPU—thermoplastic polyurethane; NaIO4—sodium periodate; Mg(NO3)2—magnesium nitrate; TBAI—tetrabutylammonium iodide; PVAc—poly(vinyl acetate); LiCl—lithium chloride.
Table 6. Examples of plasticized SPEs systems found in the literature.
Table 6. Examples of plasticized SPEs systems found in the literature.
Electrolyte SystemPreparation MethodConductivity, σ (S cm−1)Important FindingsReference
PEO-LiPF6-ECSolution casting2.06 × 10−4 (25 °C)Bandgap energy = 5.63 eV (indirect) and 5.80 eV (direct)
Unplasticized SPE:
σ = 4.10 × 10−5 S cm−1 (25 °C)
Bandgap energy = 5.74 eV (indirect) and 5.90 eV (direct)
[103]
PGMA-LiClO4-ECSolution casting2.00 × 10−4 (RT)PGMA was synthesized via photo-polymerization route.
EA = 0.21 eV
Electrochemical stability window 3.00 V
Unplasticized SPE:
σ = 4.20 × 10−5 S cm−1 (RT);
EA = 0.26 eV
Electrochemical stability window 2.50 V
[104]
PEO-LiClO4-SNSolution casting~9.12 × 10−5 (20 °C)EA = 0.45 eV
Unplasticized SPE:
σ = 3.79 × 10−7 S cm−1 (20 °C)
EA = 0.89 eV
[105]
HEC-LiCF3SO3-glycerol
[EO]/[Li] = 6:1
Solution casting1.06 × 10−5 (30 °C)EA = 0.18 eV; Tg = −83 °C[106]
Chitosan-LiClO4-glycerolSolution casting~3.00 × 10−5 (25 °C)Too much glycerol, SPE easily torn off.
Without glycerol:
σ = 1.40 × 10−5 S cm−1 (25 °C)
[107]
Chitosan-LiClO4-glycerolSolution casting1.20 × 10−3 (25 °C)tion = 0.955
D = 4.19 × 10−9 cm2 s−1;
μ = 1.63 × 10−7 cm2 V−1 s−1;
n = 4.57 × 1022 cm−3
Specific capacitance = 98.99 F g−1 at scan rate 10 mV s−1
[108]
PVB-LiClO4-PEG400Solution casting2.15 × 10−6 (RT)Unplasticized SPE:
σ = 5.80 × 10−10 S cm−1
[109]
PEMA-NH4I-ECSolution casting1.16 × 10−5 (RT)tion = 0.85
n = 1.75 × 1021 m−3
[110]
PEMA-NH4CF3SO3-BMATSFISolution casting8.35 × 10−4 (30 °C)Tg = 2 °C; tion = 0.82
Unplasticized SPE:
σ = 1.02 × 10−5 S cm−1; Tg = 68 °C
[111]
PVDF-HFP-LiClO4-SNSolution casting~1.62 × 10−3 (20 °C)EA = 0.40 eV
Unplasticized SPE:
σ = ~6.38 × 10−9 S cm−1 (20 °C)
EA = 0.69 eV
[105]
PVDF-HFP-LiBETI-SNSolution casting~2.58 × 10−3 (20 °C)LiCoO2/SPE/Li4Ti5O12 cell (C/10, 25 °C):
Initial discharge capacity 116 mAh g−1, 78% capacity retention after 120th cycle
[105]
PVDF-PEO-LiClO4-SNSolution infiltration in PVDF film + drying2.80 × 10−5(25 °C)Tg = −75.6 °C; tLi+ = 0.37 (25 °C)
LiFePO4/SPE/Li:
OCV = 3.1 V, discharge capacity 109 mAh g−1 (100th cycle), 90% capacity retention
[112]
PEO-LiCF3SO3-TiO2-EC
[EO]/[Li] = 9:1
Solution casting1.60 × 10−4 (30 °C)EA = 0.60 eV; Tg = −50 °C; Tm = 50 °C[43]
PEO-KI-PEG Solution casting5.27 × 10−5 (25 °C)Unblended PEO-KI SPE:
σ = 1.96 × 10−5 S cm−1 (25 °C)
Low Mw of PEG (~4000) might be trapped or crosslink in PEO-KI network
[113]
PEO-LiTFSI-S2TFSIMixing, vacuum-sealed in pouch and hot press0.96 × 10−3 (22 °C)
4.00 × 10−3 (60 °C)
tLi+ = 0.31 (60 °C)
LiFePO4/SPE/Li:
Discharge capacity (1 C, 60 °C) = 160.1 mAh g−1 (500th cycle), 89.9% capacity retention
[114]
PEO-LiFSI-N1222FSI
[EO]/[Li] = 20:1
Solution casting1.78 × 10−5 (25 °C)
2.14 × 10−4 (50 °C)
Tg = −58.1 °C; %χC = 24.7%
High thermal stability with decomposition temperature above 220 °C
LiFePO4/SPE/Li:
Discharge capacity (0.2 C, 50°C) = 151.5 mAh g−1 (120th cycle), 95.4% capacity retention
Unplasticized SPE:
σ = 7.24 × 10−7 S cm−1 (25 °C)
Tg = −44.8 °C; %χC = 33.6%
[115]
PEO:MC (60:40)-NH4I-PEG
(Mw PEG 8000)
Solution casting3.37 × 10−3 (30 °C)EA = 0.0322 eV
Unplasticized SPE:
σ = 7.62 × 10−5 S cm−1 (30 °C)
EA = 0.0465 eV
[116,117]
PMMA-LiClO4-PC
[MMA]/[Li] = 4:1
Solution casting3.52 × 10−5 (30 °C)EA = 0.11 eV
Conduction and relaxation via hopping mechanism
[118]
PVA-LiClO4-DMP
(30:10:60)
Solution casting0.15 × 10−3 (29 °C)Conductivity-temperature relation obeys Vogel-Tamman-Fulcher (VTF) rule[119]
PVA-LiCF3SO3-DMP
(30:10:60)
Solution casting~6.35 × 10−5 (29 °C)Conductivity-temperature relation obeys Vogel-Tamman-Fulcher (VTF) rule[119]
PVA-LiBF4-DMP
(30:10:60)
Solution casting~1.66 × 10−5 (29 °C)Conductivity-temperature relation obeys Vogel-Tamman-Fulcher (VTF) rule[119]
PEO-LiCF3SO3-15 wt.% PEG (Mw PEG 6000)Ball-milling and hot press1.71 × 10−5Tg = −55.8 °C; %χC = 31.6%[38]
PEO-LiCF3SO3-20 wt.% DOPBall-milling and hot press7.60 × 10−4Tg = −63.9 °C; %χC = 23.1%[38]
PEO-LiCF3SO3-20 wt.% ECSolution casting8.12 × 10−5 (25 °C)Tg = −70.5 °C; %χC = 31.3%
Without plasticizer:
σ = 2.89 × 10−5 S cm−1 (25 °C)
Tg = −69.0 °C; %χC = 31.7%
[120]
Corn starch-LiPF6-BmImTfSolution casting6.00 × 10−4 (80 °C)EA = 0.01 eV; Tg = −29 °C
Without plasticizer:
σ = ~10−6 S cm−1 (80 °C)
EA = 0.13 eV; Tg = −19 °C
[121]
CMC-LiBF4-glycerolSolution casting3.70 × 10−3 (25 °C)Tg = 54°C
Tg of CMC film = 87 °C
Without glycerol:
σ = 8.20 × 10−6 S cm−1 (25 °C)
[122]
PVA-CH3COONH4-BmImISolution casting9.63 × 10−5 (25 °C)EA = 6.68 meV; %χC = 2%
D+ = 9.80 × 10−12 cm2 s−1
EDLC at scan rate 10 mV s−1, 200th cycle:
Specific capacitance = 44.46 F g−1
Power density = 47.66 kW kg−1
Without BmImI:
σ = 1.94 × 10−5 S cm−1 (25 °C)
[123]
Chitosan-NH4I-BmImISolution casting3.43 × 10−5 (27 °C)EA = 0.25 eV
N3/TiO2/SPE/Pt DSSC (light intensity of 100 mW cm−2):
OCV = 0.58 V; Jsc = 3.10 mA cm−2;
FF = 0.41; η = 0.74%
Without BmImI:
σ = 3.73 × 10−7 S cm−1 (27 °C)
[124]
PEO-LiTFSI-30 wt.% TCNE
[EO]/[Li] = 12:1
Solution casting1.11 × 10−4 (25 °C)Tg = −46.21 °C
Without TCNE:
σ = 1.21 × 10−5 S cm−1 (25 °C)
Tg = −43.03°C
[125]
Chitosan-NaI-I2-EMImSCNSolution casting2.60 × 10−4 (RT)N719/TiO2/SPE/Pt DSSC (light intensity of 100 mW cm−2):
OCV = 0.53 V; Jsc = 2.62 mA cm−2;
FF = 0.52; η = 0.73%
Without EMImSCN:
σ = 1.21 × 10−5 S cm−1 (25 °C)
OCV = 0.35 V; Jsc = 1.05 mA cm−2;
FF = 0.34; η = 0.13%
[126]
PVDF-KI-I2-30 wt.% MCPSolution casting4.58 × 10−5 (30 °C)Tm = 152.8 °C; %χC = 51.49%
N3/TiO2/SPE/Pt DSSC:
η = 2.65% at 60 mW cm−2
Without MCP:
σ = 4.79 × 10−6 S cm−1 (30 °C)
Tm = 170.5 °C; %χC = 63.80%
η = 1.18% at 60 mW cm−2
[127]
PVDF-HFP-KI-TBAI-I2-
4 wt.% ATDT
Solution casting2.82 × 10−4 (30 °C)Porous surface morphology
N3/TiO2/SPE/Pt DSSC:
η = 4.64% at 60 mW cm−2
Without ATDT:
σ = 9.99 × 10−6 S cm−1 (30 °C)
η = 1.88% at 60 mW cm−2
[128]
PVDF-HFP-NaI-EMImSCNSolution casting2.65 × 10−3 (30 °C)Electrochemical stability window 3.6 V
EDLC at scan rate 0.05 mV s−1:
Specific capacitance = 2.36 F g−1
[129]
P(GMA-co-MMA)-LiClO4-EC Solution casting3.00 × 10−4 (RT)Electrochemical stability window 3.8 V
Without EC:
σ = 8.70 × 10−6 S cm−1 (RT)
[65]
PVAc:PMMA (70:30)-LiCl-ECSolution casting1.03 × 10−4 (30 °C)Tg = 26.6 °C; t+ = 0.98
D+ = 7.00 × 10−11 cm2 s−1;
μ = 2.68 × 10−9 cm2 V−1 s−1;
n = 2.35 × 1022 cm−3
Without plasticizer:
σ = 1.03 × 10−5 S cm−1 (30 °C)
Tg = 42.6 °C; t+ = 0.97
D+ = 6.93 × 10−11 cm2 s−1; μ = 2.65 × 10−9 cm2 V−1 s−1;
n = 2.35 × 1022 cm−3
[99]
PGMA—poly glycidyl methacrylate; SN—succinonitrile; HEC—hydroxyethylcellulose; PVB—polyvinyl butyral; NH4CF3SO3—ammonium trifluoromethane sulfonate; BMATFSI—butyl trimethylammonium bis(trifluoromethanesulfonyl)imide; LiBETI—lithium bisperfluoroethylsulfonylimide (Li(C2F5SO2)2N); LiCoO2 –lithium cobalt oxide; S2TFSI—triethylsulfonium bis(trifluoromethylsulfonyl)imide; LiFSI—lithium bis(fluorosulfonyl)imide; N1222FSI—triethylmethylammonium bis(fluorosulfonyl)imide; DMP—dimethyl phthalate; DOP—dioxy phthalate; BmImTf—1-butyl-3-methylimidazolium trifluoromethanesulfonate; CH3COONH4—ammonium acetate; BmImI—1-butyl-3-methylimidazolium iodide; TCNE—tetracyanoethylene; EMImSCN—1-ethyl 3-methylimidazolium thiocyanate; MCP—2-mercatopyridine; ATDT—5-amino-1,3,4-thiadiazole-2-thiol.
Table 7. Various SPEs systems with fillers added.
Table 7. Various SPEs systems with fillers added.
Electrolyte SystemPreparation MethodConductivity, σ (S cm−1)Important FindingsReference
PEO-LiCF3SO3-TiO2
[EO]/[Li] = 9:1
Solution casting4.90 × 10−5 (30 °C)EA = 0.82 eV; Tg = −46 °C;
Tm = 60 °C
[43]
PEO-LiN(CF3SO2)2-SiO2
[EO]/[Li] = 8:1
Solution casting1.40 × 10−4 EA = 0.54 eV; Tg = −43 °C
Without filler:
σ = 1.50 × 10−5 S cm−1 (27 °C);
EA = 0.60 eV; Tg = −49 °C
[177]
PEO-LiClO4-SiO2
[EO]/[Li] = 8:1
Solution casting2.50 × 10−5 EA = 0.55 eV; Tg = −44 °C
Without filler:
σ = 9.70 × 10−7 S cm−1; Tg = −46 °C
[177]
PEO-LiClO4-LLTOSolution casting7.99 × 10−4 (70 °C)LLTO nanoparticles was synthesized via Pechini precursor method.
PCTB/SPE/Li cell (60 mA g−1; 70 °C):
Discharge capacity = 104 mAh g−1
90% capacity retention after 300th cycle
[178]
PEO-LiPF6-EC-αCNTSolution casting1.30 × 10−3 (25 °C)αCNT was synthesized via chemical route
Bandgap energy = 4.42 eV (indirect) and 4.60 eV (direct)
[103]
Chitosan-LiClO4-glycerol-Ag nanowiresSolution casting~2.07 × 10−5 (25 °C)Diameter of Ag nanowires = 20 nm
Elastic modulus = ~3.5 MPa
[107]
Chitosan-LiClO4-glycerol-Ag nanospheresSolution casting~1.00 × 10−4 (25 °C)Diameter of Ag nanospheres = 100 nm
Elastic modulus = ~1.0 MPa
[107]
PVA: PVP (1:1)-MgCl2-CuSSolution casting3.32 × 10−7 (RT)
1.85 × 10−3 (protonic σ at RT)
CuS nanoparticles (particle size 20 nm) was synthesized via microwave irradiation method
More amorphous after CuS addition
Suitable for solid state magnesium batteries, supercapacitors and PEMFC
[179]
PEO-PMMA (1:1)-LiClO4-3 wt.% SiO2
[EO + C=O]/[Li] = 9:1
Solution casting + melt press0.51 × 10−5 (27 °C)SiO2 particle size < 15 nm
EA = 0.32 eV; tLi+ = 0.9894
Electrochemical stability window 6.18 V
Without filler:
σ = 1.72 × 10−5 S cm−1 (27 °C);
EA = 0.37 eV
[180,181]
PEO-PMMA (1:1)-LiClO4-3 wt.% Al2O3
[EO + C=O]/[Li] = 9:1
Solution casting + melt press0.38 × 10−5 (27 °C)Al2O3 particle size < 50 nm
EA = 0.30 eV; tLi+ = 0.9836
Electrochemical stability window 6.00 V
[180,181]
PEO-PMMA (1:1)-LiClO4-3 wt.% ZnO
[EO + C=O]/[Li] = 9:1
Solution casting + melt press1.67 × 10−5 (27 °C)ZnO particle size < 100 nm
EA = 0.27 eV; tLi+ = 0.9857
Electrochemical stability window 6.06 V
[180,181]
PEO-PMMA (1:1)-LiClO4-3 wt.% SnO2
[EO + C=O]/[Li] = 9:1
Solution casting + melt press1.00 × 10−5 (27 °C)SnO2 particle size = 100 nm
EA = 0.38 eV; tLi+ = 0.9881
Electrochemical stability window 6.10 V
[180,181]
PEO-PMMA (1:1)-LiClO4-10 wt.% MMT
[EO + C=O]/[Li] = 12:1
Microwave irradiation + Solution casting and melt press3.79 × 10−6 (27 °C)MMT is layered nanosheet which is compatible with PEO polymer chain
Without filler:
σ = 1.99 × 10−6 S cm−1 (27 °C)
[100]
PVDF-CA-LiTFSI-OMMT 3.40 × 10−4 (25 °C)EA = 0.45 eV; tLi+ = 0.315
Tensile strength = 44.89 MPa
Electrochemical stability window 4.20 V
LiFePO4/SPE/Li:
Discharge capacity = 112.9 mAh g−1 (0.5 C; 100th cycle)
Without filler:
EA = 0.53 eV; tLi+ = 0.120
Tensile strength = 34.93 MPa
[182]
PEO-PMMA (1:1)-LiBF4-EC-MMT
PEO-PMMA (1:1)-LiBF4-EC-MMT
Solution cast method
ultrasonic-microwave irradiation solution cast method
9.17 × 10−6 (27 °C)
6.43 × 10−6 (27 °C)
Ion-dipolar complexes intercalated in MMT network in SPE prepared via solution cast method.
Exfoliated MMT structure in SPE by ultrasonic-microwave irradiation solution cast method
[183]
PEO-LiCF3SO3-20 wt.% Al2O3Ball-milling and hot press8.64 × 10−5 (RT)Tg = −65.3 °C; %χC = 18.6%[38]
PEO-LiCF3SO3-EC-15 wt.% Al2O3Solution casting5.07 × 10−4 (25 °C)Tg = −72.0 °C; %χC = 31.1%
Without filler:
σ = 8.12 × 10−5 S cm−1 (25 °C)
Tg = −70.5 °C; %χC = 31.3%
[120]
PMA:PEG (65:35)-NaClO4-nano α-Al2O3Solution casting1.76 × 10−4 (70 °C)EA = 0.38 eV
Tensile strength = 50.79 MPa
Electrochemical stability window 4.50 V
Na3V2(PO4)3/SPE/Na cell (0.5 C, 70 °C):
Discharge capacity = 85 mAh g−1
94.1% capacity retention after 350th cycle
Without filler:
σ around 10−5 S cm−1 (70 °C)
[184]
PEO-PVP-NaIO4-GOSolution casting1.00 × 10−6 (30 °C)Without filler:
σ = 1.57 × 10−7 S cm−1 (30 °C)
[185]
PEO-LiTFSI-g-C3N4
[EO]/[Li] = 20:1
Solution casting3.06 × 10−5 (25 °C)
2.50 × 10−4 (60 °C)
g-C3N4 was synthesized via thermal polycondensation route
tLi+ = 0.69; Tg = −41.1 °C
Tensile strength = 3.97 MPa
Electrochemical stability window 5.12 V at 60 °C
FeF3/SPE/Li cell (1 C, 60 °C):
Capacity 300 mAh g−1 after 200th cycle
Without g-C3N4:
σ = 2.32 × 10−6 S cm−1 (25 °C)
tLi+ = 0.25; Tg = −40.7 °C
Tensile strength = 0.92 MPa
Electrochemical stability window 4.60 V at 60 °C
[186]
PEO-LiTFSI-Li10GeP2S12
[EO]/[Li] = 18:1
Solution casting1.18 × 10−5 (25 °C)
1.21 × 10−3 (80 °C)
Li10GeP2S12 (particle size 2–3 μm) was synthesized via solid state reaction method
Tg = −41.6 °C; %χC = 42.14%
tLi+ = 0.26 (80 °C)
Electrochemical stability window 5.70 V
LiFePO4/SPE/Li cell (0.5 C, 60 °C):
Discharge capacity = 137.4 mAh g−1
92.5% capacity retention after 50th cycle
Without filler:
σ = 6.16 × 10−6 S cm−1 (25 °C);
= 7.98 × 10−4 S cm−1 (80 °C);
Tg = −39.6 °C; %χC = 67.07%
tLi+ = 0.22 (80 °C)
Electrochemical stability window 4.80 V
Discharge capacity = ~50 mAh g−1
50.4% capacity retention after 50th cycle
[187]
PVAc:PMMA (70:30)-LiCl-TiO2Solution casting4.45 × 10−4 (30 °C)Tg = 28.1 °C; t+ = 0.99
D+ = 7.07 × 10−11 cm2 s−1;
μ = 2.71 × 10−9 cm2 V−1 s−1;
n = 2.35 × 1022 cm−3
Electrochemical stability window 2.69 V
Without filler:
σ = 1.03 × 10−5 S cm−1 (30 °C)
Tg = 42.6 °C; t+ = 0.97
D+ = 6.93 × 10−11 cm2 s−1;
μ = 2.65 × 10−9 cm2 V−1 s−1;
n = 2.35 × 1022 cm−3
Electrochemical stability window 1.69 V
[99]
PEO-LiTFSI-Ni3-(BTC)2Hot pressing1.40 × 10-1 (30 °C)LiFePO4/SPE/Li cell (1 C):
Discharge capacity = 75 mAh g-1
[188]
LLTO—Li0.3La0.566TiO3; PCTB—poly(2-chloro-3,5,6-trisulfide-1,4-benzoquinone); αCNT—amorphous carbon nanotubes; CuS—copper sulfate; CA—cellulose acetate; OMMT—organic modified montmorillonite; MMT—montmorillonite; Al2O3—nano alumina; NaIO4—sodium periodate; GO—graphene oxide; PMA—poly(methacrylate); g-C3N4—polymeric graphitic carbon nitride; Ni3-(BTC)2—nickel-1,3,5-benzene trcarboxxylate metal organic framework.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Teo, L.P.; Buraidah, M.H.; Arof, A.K. Development on Solid Polymer Electrolytes for Electrochemical Devices. Molecules 2021, 26, 6499. https://doi.org/10.3390/molecules26216499

AMA Style

Teo LP, Buraidah MH, Arof AK. Development on Solid Polymer Electrolytes for Electrochemical Devices. Molecules. 2021; 26(21):6499. https://doi.org/10.3390/molecules26216499

Chicago/Turabian Style

Teo, Li Ping, Mohd Hamdi Buraidah, and Abdul Kariem Arof. 2021. "Development on Solid Polymer Electrolytes for Electrochemical Devices" Molecules 26, no. 21: 6499. https://doi.org/10.3390/molecules26216499

Article Metrics

Back to TopTop