Next Article in Journal
Polydopamine-Supported Lipid Bilayers
Previous Article in Journal
Spectroscopy of Deep Traps in Cu2S-CdS Junction Structures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Properties of PEMA-NH4CF3SO3 Added to BMATSFI Ionic Liquid

by
Norwati Khairul Anuar
1,2,
Ri Hanum Yahaya Subban
1 and
Nor Sabirin Mohamed
2,*
1
Faculty of Applied Sciences, Universiti Teknologi MARA, Shah Alam Selangor 40450, Malaysia
2
Center for Foundation Studies in Science, University of Malaya, Kuala Lumpur 50603, Malaysia
*
Author to whom correspondence should be addressed.
Materials 2012, 5(12), 2609-2620; https://doi.org/10.3390/ma5122609
Submission received: 12 September 2012 / Revised: 31 October 2012 / Accepted: 2 November 2012 / Published: 4 December 2012

Abstract

:
Polymer electrolyte films, comprising ammonium trifluoromethanesulfonate salt and butyl-trimethyl ammonium bis(trifluoromethylsulfonyl)imide ionic liquid immobilized in poly (ethyl methacrylate) was studied. Structural, morphological, thermal and electrical properties of the polymer electrolyte films were investigated by differential scanning calorimetry, scanning electron microscopy, and impedance spectroscopy, respectively. Interactions of the salt and ionic liquid with the host polymer were investigated by Fourier transform infra-red spectroscopy. Electrochemical stability of the electrolytes was determined using linear sweep voltammetry and transference numbers corresponding to ionic transport has been evaluated by means of the Wagner polarization technique. The highest conductivity achieved is in the order of 10−4 S cm−1 for the film added with 35 wt % butyl trimethylammonium bis (trifluoromethanesulfonyl)imide. The film has high amorphicity and low glass transition temperature of 2 °C. The film is electrochemically stable up to 1.8 V. The ion transference number in the polymer film is 0.82 and the conductivity behavior obeys Vogel-Tamman-Fulcher equation.

1. Introduction

Proton conducting polymer electrolytes containing ammonium salts complexed with PEO and PPO were reported in the mid-eighties [1,2]. In such systems, the charge carriers are H+ ions that come from ammonium ions of the salts. The conduction of the H+ ions occurs through exchange of H+ ions between complexed sites [3,4]. In the present study, proton conducting polymer electrolytes were prepared using PEMA as the host. The use of PEMA as a host polymer was first reported by Han et al. and Fahmy et al. [3,5,6]. Few studies of PEMA as a host revealed it has an ionic conductivity of the order of 10−3 S cm−1 and electrochemical stability up to 4.3 V [3,7]. PEMA based polymer electrolytes also have been explored in blending with other polymers namely PVC and PVdF-HFP to obtain mechanically stable films that show ionic conductivity up to 10−3 S cm−1 [3,8,9,10,11]. In an earlier paper [12], we reported on our study of a PEMA-NH4SO3CF3 system. The system containing 35 wt % ammonium salt showed the highest conductivity of 1.02 × 10−5 S cm−1. The present study focuses on our efforts to increase the conductivity of the PEMA-NH4SO3CF3 system by adding different amounts of butyl trimethylammonium bis (trifluoromethanesulfonyl)imide (BMATFSI) ionic liquid. Ionic liquid (IL) meet the requirements of plasticizing salts and offer an improved thermal and mechanical properties to flexible polymers. Different polymer electrolytes containing IL have been reported to possess high conductivity. In addition, the incorporation of ILs into polymer electrolytes distinctively improves their electrochemical stability and increases the ionic conductivity of the polymer electrolytes [13,14,15,16,17,18,19,20,21,22]. Although PEMA is hydrophilic, and ionic liquid BMATFSI is hydrophobic, they can interact and form transparent film. The structural, morphological, thermal and electrical properties of the ionic liquid-added PEMA-NH4SO3CF3 system are reported in this paper.

2. Experimental Section

2.1. Materials

Poly ethyl methacrylate (PEMA, Mw ~ 515,000 g/mol), ammonium trifluoromethane sulfonate (NH4SO3CF3, 99%) and butyl trimethylammonium bis (trifluoromethanesulfonyl)imide (BMATFSI) were purchased from Aldrich, Germany.

2.2. Characterization

Polymer-salt-ionic liquid films with thickness ranging between 100 and 300 μm were prepared by solution casting technique using tetrahydrofuran (THF) as the solvent. PEMA and NH4SO3CF3 at fixed ratio of 65:35 wt % added with different ratios of BMATFSI were mixed and stirred for 24 h to achieve a homogeneous, viscous solution. The solution obtained was cast on a glass plate and allowed to evaporate slowly at room temperature. The films was further dried under vacuum at 40 °C for 24 h. Infrared spectra were collected at room temperature using Perkin Elmer FTIR Spectrometer; Spectrum 400. Glass transition temperatures (Tg) were measured with a METTLER TOLEDO DSC822 differential scanning calorimeter under nitrogen environment at scanning rate of 10 °C/min over a temperature range from −65 to 300 °C. The Tg was determined using the mid-point method on the DSC curve. Ionic conductivity was measured with a computer controlled HIOKI 3532-50 LCR HITESTER frequency response analyzer. Bulk resistance was determined from the x-intercept of the imaginary versus real impedance plot. The conductivity values were calculated using the equation
σ = t A R b
where σ is conductivity (S cm−1) and t and A are the film’s thickness and cross section area, respectively. The values reported are an average of six measurements. The electrochemical stability of the electrolytes was determined using linear sweep voltammetry (LSV) at scanning rate of 1 mV/s from −2.0 V to 4.0 V. Silver paint was applied on each side of samples. The transference numbers corresponding to ionic (tion) and electronic (te) transport have been evaluated by means of the Wagner polarization technique [23] for constant dc voltage of 2 mV. Voltage was applied across the blocking electrodes and current passing through the cells was measured as a function of time to allow the samples to become fully polarized. The experimental values of the total current (IT), which is the sum of ionic (Ii) and electronic (Ie) currents on immediate voltage application and saturated electronic current (Ie) give an estimate of ionic and electronic transport numbers in accordance with relation
t ion = ( I T I e ) I T

3. Results and Discussion

3.1. Differential Scanning Calorimetry

Differential Scanning Calorimetry (DSC) has been carried out on the pure PEMA, PEMA-NH4SO3CF3 and PEMA-NH4SO3CF3-BMATFSI films. DSC curves of the films are shown in Figure 1. Figure 1a reveals that Tg of the pure PEMA film is 72 °C while its melting temperature is 273 °C. This suggests that pure PEMA film is semi-crystalline in nature. It can be seen that with addition of salt and ionic liquid, the Tg shifts to lower temperatures. The glass transition temperature value of PEMA containing ammonium salt is observed to be 68 °C. This value is lower than that of pure PEMA. The glass transition temperature decreases further with ionic liquid addition as shown in Table 1. The decrease in Tg with salt addition is due to dissolved ion being accommodated in the PEMA phase [24]. Meanwhile, the decrease in Tg upon addition of BMATFSI is due to the presence of BMATFSI that has acted as a plasticizer and increased the chain mobility by spacing out the host polymer chains [25]. The DSC results reveals that the addition of BMATFSI can indeed weaken the interaction among the polymer chains.
Table 1. Glass transition temperature of PEMA-NH4SO3CF3-BMATFSI polymer electrolyte films. Where PEMA is poly ethyl methacrylate, BMATFSI is butyl trimethylammonium bis (trifluoromethanesulfonyl)imide.
Table 1. Glass transition temperature of PEMA-NH4SO3CF3-BMATFSI polymer electrolyte films. Where PEMA is poly ethyl methacrylate, BMATFSI is butyl trimethylammonium bis (trifluoromethanesulfonyl)imide.
Polymer filmGlass transition temperature, Tg (°C)
Pure PEMA72
(PEMA-NH4SO3CF3)68
(PEMA-NH4SO3CF3)-BMATFSI 15 wt %43
(PEMA-NH4SO3CF3)-BMATFSI 25 wt %29
(PEMA-NH4SO3CF3)-BMATFSI 35 wt %2
Figure 1. Differential Scanning Calorimetry curve for (a) PEMA; (b) PEMA-NH4SO3CF3 and PEMA-NH4SO3CF3 containing; (c) 15 wt % BMATFSI; (d) 25 wt % BMATFSI; (e) 35 wt % BMATFSI.
Figure 1. Differential Scanning Calorimetry curve for (a) PEMA; (b) PEMA-NH4SO3CF3 and PEMA-NH4SO3CF3 containing; (c) 15 wt % BMATFSI; (d) 25 wt % BMATFSI; (e) 35 wt % BMATFSI.
Materials 05 02609 g001

3.2. Scanning Electron Microscopy (SEM)

The obtained SEM micrographs of pure PEMA, PEMA-NH4SO3CF3 and PEMA-NH4SO3CF3-BMATFSI polymer electrolyte films measured at 300 K are depicted in Figure 2. Pure PEMA electrolytes show large and well-defined spherulites. The existence of the well defined spherulites of 3–11 μm in diameter indicates the presence of crystalline region in the PEMA film. This observation is consistent with the DSC result discussed earlier.
When NH4SO3CF3 is complexed with polymer (Figure 2b), the size of the spherulites decreases to 0.4–3 μm. When the BMATFSI is added to this system (Figure 2c), an increase in amorphous region (light grey) is clearly seen. This is the reason for the shifting of the glass transition peaks, which is noticed in DSC. Figure 2 also shows a decrease in surface roughness upon addition BMATFSI. This could help in enhancing contact at the electrolyte–electrode interface [26].
Figure 2. Scanning Electron micrograph of (a) PEMA; (b) PEMA-NH4SO3CF3 and (c) PEMA-NH4SO3CF3-BMATFSI.
Figure 2. Scanning Electron micrograph of (a) PEMA; (b) PEMA-NH4SO3CF3 and (c) PEMA-NH4SO3CF3-BMATFSI.
Materials 05 02609 g002

3.3. Fourier Transform Infrared

Presented in Figure 3 are the FTIR spectra of PEMA-NH4SO3CF3 containing 35 wt % BMATFSI, PEMA-NH4SO3CF3 containing 25 wt % BMATFSI, PEMA-NH4SO3CF3 and pure PEMA. The appearance of a strong band in the spectrum at 1723 cm−1 which corresponds to C=O stretching frequency of pure PEMA is slightly shifted to a lower wave number of 1721–1718 cm−1 in the polymer complex (Figure 3a).
Figure 3. Fourier Transform Infrared spectrum of (a) pure PEMA; (b) PEMA-NH4SO3CF3 and (c) PEMA-NH4SO3CF3 added with 35 wt % BMATFSI.
Figure 3. Fourier Transform Infrared spectrum of (a) pure PEMA; (b) PEMA-NH4SO3CF3 and (c) PEMA-NH4SO3CF3 added with 35 wt % BMATFSI.
Materials 05 02609 g003
This effect is due to the coordination of the cation of NH4SO3CF3 with the oxygen, which results in the weakening of the C=O bond. Similar results have been reported by Weihua Zhu et al. [27] and Selvasekarapandian et al. [28] for the PEG-PU/NaClO4 complexes and PVAc-NH4SCN respectively. Moreover, the addition of NH4SO3CF3 causes a shift of the C–O–C stretching band at 1237 cm−1 down to lower wave number due to the coordination of the ether oxygen with the cation of the salt (Figure 3b). A similar effect was reported by Wieczorek et al. [29] for the polyether-poly (methyl methacrylate) blend-based system. The addition of ionic liquid also made the C=O and C–O–C bonds slightly move to (1721–1718 cm−1) and (1226–1227 cm−1) respectively. This shows that the BMATFSI interacts with the host polymer.

3.4. Conductivity Study

3.4.1. Composition Dependence of Conductivity

The bulk conductivity (σ) of the studied PEMA based electrolyte materials was evaluated using complex impedance technique. Figure 4 shows typical complex impedance plot for the PEMA-NH4SO3CF3 and PEMA-NH4SO3CF3-BMATFSI films. In this plot, a semicircle corresponding to the bulk impedance can be clearly seen and the bulk resistance Rb can be very easily determined. The conductivity of PEMA, PEMA-NH4SO3CF3 and PEMA-NH4SO3CF3-BMATFSI films is given in Table 2.
Figure 4. Complex impedance plot for (a) PEMA-NH4SO3CF3 and (b) PEMA-NH4SO3CF3 containing 35 wt % BMATFSI films.
Figure 4. Complex impedance plot for (a) PEMA-NH4SO3CF3 and (b) PEMA-NH4SO3CF3 containing 35 wt % BMATFSI films.
Materials 05 02609 g004
Table 2. Conductivity of PEMA-NH4SO3CF3-BMATFSI polymer electrolyte films.
Table 2. Conductivity of PEMA-NH4SO3CF3-BMATFSI polymer electrolyte films.
Polymer filmConductivity, σ (S cm−1)
PEMA8.60 × 10−11
(PEMA-NH4SO3CF3)1.02 × 10−5
(PEMA-NH4SO3CF3)-BMATFSI 15 wt %4.05 × 10−5
(PEMA-NH4SO3CF3)-BMATFSI 25 wt %7.47 × 10−5
(PEMA-NH4SO3CF3)-BMATFSI 35 wt %8.35 × 10−4
The conductivity of PEMA is 8.60 × 10−11 S cm−1 while PEMA-NH4SO3CF3 has a conductivity value of 1.02 × 10−5 S cm−1. Addition of 5 wt % of BMATFSI increases the conductivity to 4.05 × 10−5 S cm−1. The conductivity increases further with further increase of BMATFSI. The sample with the highest BMATFSI loading had the highest conductivity and reaches a maximum value of 8.35 × 10−4 S cm−1 at 35% BMATFSI. However, the complexes with BMATFSI more than 40 wt % were found to be mechanically unstable and difficult to handle for conductivity measurement.

3.4.2. Temperature Dependence of Conductivity

Temperature dependence of conductivity for polymer electrolyte of different compositions of BMATFSI has been studied in the temperature range 30–100 °C. Figure 5 illustrates the dependence of ionic conductivity on temperature for the polymer films.
Figure 5. VTF plot of PEMA-NH4SO3CF3 containing (a) 35 wt %; (b) 25 wt %; (c) 15 wt % BMATFSI; and (d) PEMA-NH4SO3CF3.
Figure 5. VTF plot of PEMA-NH4SO3CF3 containing (a) 35 wt %; (b) 25 wt %; (c) 15 wt % BMATFSI; and (d) PEMA-NH4SO3CF3.
Materials 05 02609 g005
As can be seen in this figure, increase in temperature leads to an increase in conductivity. This is expected because as the temperature increases the polymer expands to produce free volumes, which leads to enhanced ionic mobility. BMATFSI, which acts as plastisizer, may contribute to conductivity enhancement by opening up narrow rivulets of ionic liquid-rich phases for greater ionic transport [30,31]. The ionic conduction in this polymer electrolyte system obeys the Vogel-Tamman-Fulcher (VTF) relation.
σ ( T ) = A T 1 / 2 exp [ B R ( T T o ) ]
In this equation, A is a constant proportional to the number of charge carriers, B is the conduction activation energy, R is the universal gas constant, To is the glass transition temperature. The increase in conductivity with temperature can also be interpreted as being due to a hopping mechanism between coordination sites, local structure relaxations and segmental motion of polymer. As the amorphous region progressively increases, the polymer chain acquires faster internal modes in which bond rotation produces segmental motion which helps the inter-chain and intra-chain ion movements [5]. From the VTF equation, it is clear that ionic conductivity could be improved by lowering the Tg. The conductivity result is in agreement with DSC result, where the sample with the lowest Tg value shows the highest conductivity.

3.5. Activation Energy for Proton Ion Transport

The activation energy for ion transport, EA can be determined from the gradient of the VTF plots (Table 3). The trend of the EA value suggests that the activation energy for ion transport decreases with the increase in BMATFSI. The addition of BMATFSI decreases the glass transition temperature, Tg value, hence improves segmental motion of the polymer electrolyte and helps in ion movement.
Table 3. Activation energy of PEMA-NH4SO3CF3-BMATFSI polymer electrolyte films.
Table 3. Activation energy of PEMA-NH4SO3CF3-BMATFSI polymer electrolyte films.
Polymer filmTg (K)EA (kJ)
(PEMA-NH4SO3CF3)340.932.5
(PEMA-NH4SO3CF3)-BMATFSI 15 wt %315.652.3
(PEMA-NH4SO3CF3)-BMATFSI 25 wt %302.312.2
(PEMA-NH4SO3CF3)-BMATFSI 35 wt %275.251.8

3.6. Electrochemical Stability Determination

As mentioned earlier, the PEMA-NH4SO3CF3 system containing the highest amount (35 wt %) of BMATFSI shows the highest conductivity of the order of 10−4 S cm−1. This system was then subjected to an electrochemical stability window. The electrochemical stability window of the electrolyte with 35 wt % BMATFSI was analyzed using linear sweep voltammetry (LSV) and the voltammogram is shown in Figure 6.
Figure 6. Linear sweep voltammetry curve for the film of PEMA-NH4SO3CF3 containing 35 wt % BMATFSI.
Figure 6. Linear sweep voltammetry curve for the film of PEMA-NH4SO3CF3 containing 35 wt % BMATFSI.
Materials 05 02609 g006
The onset current of the sample are detected about 1.8 V at temperature 300 K. The onset current is assumed to be the film’s breakdown voltage. This voltage is high enough to allow the safe use of a PEMA-based solid polymer electrolyte for fabrication of protonic batteries, since the electrochemical window standard of protonic battery is about ~1 V [32,33].

3.7. Ionic Transference Number Measurements

Figure 7 illustrates the current-time plot for PEMA-NH4SO3CF3containing 35 wt % BMATFSI polymer electrolytes. The experimental values of tion for samples PEMA-NH4SO3CF3 containing 0, 25 and 35 wt % BMATFSI polymer electrolyte are listed in Table 4. The incorporation of BMATFSI into PEMA-NH4SO3CF3 electrolyte led to a decrease in tion. However, the value of tion indicates that the majority of charge carriers in the electrolytes system are ions which are expected to be protons.
Figure 7. Transference number plot for the film of PEMA-NH4SO3CF3 containing 35 wt % BMATFSI film.
Figure 7. Transference number plot for the film of PEMA-NH4SO3CF3 containing 35 wt % BMATFSI film.
Materials 05 02609 g007
Table 4. Ionic transference number of PEMA-NH4SO3CF3-BMATFSI polymer electrolyte films.
Table 4. Ionic transference number of PEMA-NH4SO3CF3-BMATFSI polymer electrolyte films.
Polymer filmtion
(PEMA-NH4SO3CF3)0.999
(PEMA-NH4SO3CF3)-BMATFSI 25 wt %0.930
(PEMA-NH4SO3CF3)-BMATFSI 35 wt %0.820

4. Conclusions

The addition of BMATFSI ionic liquid increases the conductivity of PEMA-NH4SO3CF3. The ionic conductivity is due to decreases in crystallinity and the improvement of segmental motion of the polymer electrolyte. Addition of the ionic liquid also decreases roughness of the PEMA-NH4SO3CF3 surface. The conductivity shows VTF behavior, indicating that the ion transport is controlled by the polymer segmental motion of the amorphous state in the polymer electrolyte. Linear sweep voltammetry result reveals that the electrochemical stability of the BMATFSI ionic liquid-added PEMA is up to ~1.8V.

References

  1. Stainer, M.; Hardy, L.C.; Whitmore, D.H.; Shriver, D. Stoichiometry of formation and conductivity response of amorphous and crystalline complexes formed between poly(ethylene oxide) and ammonium salts: PEOx·NH4SCN and PEOx·NH4SO3CF3. J. Electrochem. Soc. 1984, 131, 784–790. [Google Scholar] [CrossRef]
  2. Ansari, S.M.; Browdwin, M.; Stainer, M.; Druger, S.D.; Ratner, M.A.; Shriver, D.F. Conductivity and dielectric constant of the polymeric solid electrolyte, (PEO)8NH4SO3CF3, in the 100 Hz to 1010 Hz range. Solid State Ionics 1985, 17, 101–106. [Google Scholar] [CrossRef]
  3. Hashmi, S.A.; Kumar, A.; Maurya, K.K.; Chandra, S. Proton-conducting polymer electrolyte I: The polyethylene oxide + NH4ClO4 system. J. Phys. D Appl. Phys. 1990, 23, 1307–1314. [Google Scholar] [CrossRef]
  4. Khatijah, S.; Subban, R.H.Y.; Mohamed, N.S. Ionic conductivity of PVC-NH4I-EC proton conducting polymer electrolytes. Adv. Mater. Res. 2012, 545, 312–316. [Google Scholar] [CrossRef]
  5. Reiter, J.; Velicka, J.; Mika, M. Proton-conducting polymer electrolytes based on methacrylates. Electrochim. Acta 2008, 53, 7769–7774. [Google Scholar]
  6. Han, H.S.; Kang, H.R.; Kim, S.W.; Kim, H.T. Phase-separated polymer electrolyte based on poly(vinyl chloride)/poly(ethyl methacrylate) blend. J. Power Source 2002, 112, 461–468. [Google Scholar] [CrossRef]
  7. Fahmy, T.; Ahmed, M.T. Thermal induced structural change investigations in PVC/PEMA polymer blend. Polym. Test. 2001, 20, 477–484. [Google Scholar] [CrossRef]
  8. Rajendran, S.; Prabhu, M.R.; Rani, M.U. Ionic conduction in poly(vinyl chloride)/poly(ethyl methacrylate)-based polymer blend electrolytes complexed with different lithium salts. J. Power Sources 2008, 180, 880–883. [Google Scholar] [CrossRef]
  9. Zakaria, N.A.; Isa, M.I.N.; Mohamed, N.S.; Subban, R.H.Y. Characterization of polyvinyl chloride/polyethyl methacrylate polymer blend for use as polymer host in polymer electrolytes. J. Appl. Polym. Sci. 2012, 126, E419–E424. [Google Scholar] [CrossRef]
  10. Rudziah, S.; Ibrahim, S.; Mohamed, N.S. PVDF-HFP/PEMA-NH4CF3SO3-TiO2 polymer electrolyte and its application in proton batteries. Adv. Mater. Res. Appl. Eng. Mater. 2011, 287–290, 285–288. [Google Scholar]
  11. Rudhziah, S.; Muda, N.; Ibrahim, S.; Rahman, A.A.; Mohamed, N.S. Proton conducting polymer electrolytes based on PVDF-HFP and PVDF-HFP/PEMA blend. Sains Malays. 2011, 40, 707–712. [Google Scholar]
  12. Anuar, N.K.; Zainal, N.; Mohamed, N.S.; Subban, R.H.Y. Studies of poly(ethyl methacrylate) complexed with ammonium trifluoromethanesulfonate. Adv. Mater. Res. 2012, 501, 19–23. [Google Scholar] [CrossRef]
  13. Noda, A.; Watanabe, M. Highly conductive polymer electrolytes prepared by in situ polymerization of vinyl monomers in room temperature molten salts. Electrochim. Acta 2000, 45, 1265–1270. [Google Scholar] [CrossRef]
  14. Fuller, J.; Breda, A.C.; Carlin, R.T. Ionic liquid–polymer gel electrolytes from hydrophilic and hydrophobic ionic liquids. J. Electroanal. Chem. 1998, 459, 29–34. [Google Scholar] [CrossRef]
  15. Kim, K.S.; Park, S.Y.; Choi, S.; Lee, H. Ionic liquid–polymer gel electrolytes based on morpholinium salt and PVdF(HFP) copolymer. J. Power Sources 2006, 155, 385–390. [Google Scholar] [CrossRef]
  16. Bansal, D.; Cassel, F.; Croce, F.; Hendrickson, M.; Plichta, E.; Salomon, M. Conductivities and Transport Properties of Gelled Electrolytes with and without an Ionic Liquid for Li and Li-Ion Batteries. J. Phys. Chem. B 2005, 109, 4492–4496. [Google Scholar] [CrossRef] [PubMed]
  17. Shin, J.H.; Henderson, W.A.; Tizzani, C.; Passerini, S.; Jeong, S.S.; Kim, K.W. Characterization of Solvent-Free Polymer Electrolytes Consisting of Ternary PEO-LiTFSI-PYR14 TFSI. J. Electrochem. Soc. 2006, 153, A1649–A1654. [Google Scholar] [CrossRef]
  18. Singh, B.; Sekhon, S.S. Polymer electrolytes based on room temperature ionic liquid: 2,3-dimethyl-1-octylimidazolium triflate. J. Phys. Chem. B 2005, 109, 16539–16543. [Google Scholar] [CrossRef] [PubMed]
  19. Cheng, H.; Zhu, C.; Huang, B.; Lu, M.; Yang, Y. Synthesis and electrochemical characterization of PEO-based polymer electrolytes with room temperature ionic liquids. Electrochim. Acta 2007, 52, 5789–5794. [Google Scholar] [CrossRef]
  20. Singh, P.K.; Jadhav, N.A.; Mishra, S.K.; Singh, U.P.; Bhattacharya, B. Application of ionic liquid doped solid polymer electrolyte. Ionics 2010, 16, 645–648. [Google Scholar] [CrossRef]
  21. Singh, P.K.; Bhattacharya, B. Present scenario of solid state photoelectrochemical solar cell and dye sensitized solar cell using PEO-based polymer electrolytes. In Dye-Sensitized Solar Cells and Solar Cell Performance; Travino, M.R., Ed.; Nova Science Publishers: Hauppauge, New York, NY, USA, 2011; Chapter 10; pp. 235–266. [Google Scholar]
  22. Axenov, K.V.; Laschat, S. Thermotropic ionic liquid crystals. Materials 2011, 4, 206–259. [Google Scholar] [CrossRef]
  23. Wagner, J.B.; Wagner, C. Electrical Conductivity Measurements on Cuprous Halides. J. Chem. Phys. 1957, 26, 1597–1601. [Google Scholar] [CrossRef]
  24. Cai, H.; Hu, R.; Egami, T.; Farrington, C.G. The effect of salt concentration on the local atomic structure and conductivity of PEO-based NiBr2 electrolytes. Solid State Ionics 1992, 52, 333–338. [Google Scholar] [CrossRef]
  25. Ghiya, V.P.; Dave, V.; Gross, R.A.; McCarthy, S.P. Biodegradability of cellulose acetate plasticized with citrate esters. J. Macromol. Sci. A 1996, 33, 627–638. [Google Scholar] [CrossRef]
  26. Singh, P.K.; Nagarale, R.K.; Pandey, S.P.; Rhee, H.W.; Bhattacharya, B. Present status of solid state photoelectrochemical solar cells and dye sensitized solar cells using PEO-based polymer electrolytes. Adv. Nat. Sci. Nanosci. Nanotechnol. 2011, 2, 023002:1–13. [Google Scholar] [CrossRef]
  27. Zhu, W.; Wang, X.; Yang, B.; Tang, X. A novel ionic-conduction mechanism based on polyurethane electrolyte. J. Polymer Sci. B 2001, 39, 1246–1254. [Google Scholar] [CrossRef]
  28. Selvasekarapandian, S.; Baskaran, R.; Hema, M. Complex AC impedance, transference number and vibrational spectroscopy studies of proton conducting PVAc–NH4SCN polymerelectrolytes. Phys. B Condens. Matter 2005, 357, 412–419. [Google Scholar] [CrossRef]
  29. Wieczoreck, W.; Stevens, J.R. Impedance spectroscopy and phase structure of polyether-poly(methyl methacrylate)-LiCF3SO3 blend-based electrolytes. J. Phys. Chem. B 1997, 101, 1529–1534. [Google Scholar] [CrossRef]
  30. Rhoo, H.J.; Kim, H.T.; Park, J.M.; Hwang, T.S. Ionic conduction in plasticized PVC/PMMA blend polymer electrolytes. Electrochim. Acta 1997, 42, 1571–1579. [Google Scholar] [CrossRef]
  31. Aravindan, V.; Lakshmi, C.; Vickraman, P. Investigations on Na+ ion conducting polyvinylidenefluoride-co-hexafluoropropylene/poly ethylmetharcrylate blend polymer electrolytes. Curr. Appl. Phys. 2009, 9, 1106–1111. [Google Scholar] [CrossRef]
  32. Pratap, R.; Singh, B.; Chandra, S. Polymeric rechargeable solid-state proton battery. J. Power Sources 2006, 161, 702–706. [Google Scholar] [CrossRef]
  33. Ng, L.S.; Mohamad, A.A. Effect of temperature on the performance of proton batteries based on chitosan-NH4NO3-EC membrane. J. Membr. Sci. 2008, 325, 653–657. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Anuar, N.K.; Subban, R.H.Y.; Mohamed, N.S. Properties of PEMA-NH4CF3SO3 Added to BMATSFI Ionic Liquid. Materials 2012, 5, 2609-2620. https://doi.org/10.3390/ma5122609

AMA Style

Anuar NK, Subban RHY, Mohamed NS. Properties of PEMA-NH4CF3SO3 Added to BMATSFI Ionic Liquid. Materials. 2012; 5(12):2609-2620. https://doi.org/10.3390/ma5122609

Chicago/Turabian Style

Anuar, Norwati Khairul, Ri Hanum Yahaya Subban, and Nor Sabirin Mohamed. 2012. "Properties of PEMA-NH4CF3SO3 Added to BMATSFI Ionic Liquid" Materials 5, no. 12: 2609-2620. https://doi.org/10.3390/ma5122609

Article Metrics

Back to TopTop