Journal of Membrane Science 318 (2008) 422–428

Contents lists available at ScienceDirect

Journal of Membrane Science journal homepage: www.elsevier.com/locate/memsci

Poly(vinylidene fluoride)–polydiphenylamine composite electrospun membrane as high-performance polymer electrolyte for lithium batteries Anantha Iyengar Gopalan a,b , Kwang-Pill Lee a,b,∗ , Kalayil Manian Manesh a , Padmanabhan Santhosh a a b

Department of Chemistry Graduate School, Kyungpook National University, Daegu 702-701, South Korea Nano Practical Application Center, Daegu 704-230, South Korea

a r t i c l e

i n f o

Article history: Received 11 October 2007 Received in revised form 6 March 2008 Accepted 7 March 2008 Available online 15 March 2008 Keywords: Polymer electrolyte Lithium batteries Poly(vinylidene fluoride) Polydiphenylamine Electrospun membrane

a b s t r a c t Composites of poly(vinylidene fluoride) (PVdF)–polydiphenylamine (PDPA) are electrospun as nanofibrous membranes, PVdF–PDPA-CFM (CFM represents composite nanofibrous membrane). Polymer electrolytes are prepared by loading lithium salts into PVdF–PDPA-CFM. Field emission scanning electron microscope (FESEM) images clearly inform that PVdF–PDPA-CFM has interconnected multi-fibrous layers with ultrafine porous structures. The average diameter of fibers in PVdF–PDPA-CFM is ∼200 nm for a loading of 0.5 wt.% of PDPA in the composite, which is far lesser than pristine PVdF membrane. There is inter-fiber twisting in the PVdF–PDPA-CFM that generates microcavities. These interconnected morphological features of PVdF–PDPA-CFM result in higher ionic conductivity, effective lithium ion transport and good interfacial characteristics with lithium electrode. Higher ionic conductivity, lithium ion transport number of about 0.48, higher liquid electrolyte uptake (>280%) with dimensional stability, lower interfacial resistance and higher electrochemical stability window of 5.18 V vs. Li for PVdF–PDPA-CFM electrolyte are witnessed. With these improved performance characteristics, PVdF–PDPA-CFM finds its suitability as polymer electrolyte for high-performance lithium rechargeable batteries. © 2008 Elsevier B.V. All rights reserved.

1. Introduction Polymer electrolytes consist of salts dissolved in solid polymers hold a key role in realizing the major goal of an all-solid-state rechargeable lithium battery [1–3]. Amorphous polymer electrolytes have been studied intensively for more than 30 years, however, the performance of the electrolyte has not increased substantially over that period. Conductivity in the order of ∼mS cm−1 was achieved in these systems. To enable high performance of the lithium battery, a polymer electrolyte should possess ionic conductivity greater than mS cm−1 , good dimensional and thermal stability, an electrochemical stability window of ≥5.0 V, chemical compatibility with Li electrodes, ability to afford Li cycling (recharge) at an efficiency of greater than 99%. Therefore, it is essential to develop polymer electrolytes to suit for the modern needs in the power sources. Poly(vinylidene fluoride), PVdF with its high mechanical stability and chemical inertness has been used as polymer electrolytes in lithium batteries [4]. PVdF is inherently polar due to the presence of electronegative fluorine atoms in the backbone structure

∗ Corresponding author at: Department of Chemistry Graduate School, Kyungpook National University, Daegu 702-701, South Korea. Tel.: +82 53 950 5901; fax: +82 539528104. E-mail address: [email protected] (K.-P. Lee). 0376-7388/$ – see front matter © 2008 Elsevier B.V. All rights reserved. doi:10.1016/j.memsci.2008.03.007

[5]. The crystalline domains in PVdF hinder migration of lithium ions and lower the ionic conductivity. Polydiphenylamine (PDPA), a polymer of N-substituted aniline, is more soluble in common organic solvents [6], exhibits different redox characteristics than other poly(N-substituted anilines) [7,8]. Recent studies show that backbone units of PDPA can be grafted with other polymeric chains to have novel functional properties [9]. Electrospinning is an efficient fabrication process that gives fibrous and porous membranes with an average diameter ranging from 100 nm to 5 ␮m [10], which are at least one or two order of magnitude smaller than the fibers produced from melt or solution spinning. Electrospinning technology has recently been extended in various fields like preparation of porous filters, biomedical materials, reinforcing components, cloths for electromagnetic wave shielding, sensors, electronic devices, etc. [11–13]. Reports on using electrospinning for making non-woven mats of polymer composites consisting of conducting polymer and a conventional polymer are scarce. Blends of polyaniline with poly(ethylene oxide) in chloroform were electrospun to produce filaments in the range of 4–20 nm [14]. PEO provides adequate viscosity to the composite solution to achieve electrospinning. In the present study, PVdF–PDPA composite nanofibrous membranes (PVdF–PDPA-CFM) were prepared though electrospinning. ␤-Naphthalene sulfonic acid (NSA)-doped PDPA is fairly soluble in DMF in which PVdF (10% w/w) can form viscous solution and

A.I. Gopalan et al. / Journal of Membrane Science 318 (2008) 422–428

this suits for electrospinning the composite [15]. The migration of lithium ions in PVdF is expected to be enhanced by forming composites with PDPA that can have physicochemical interactions with C–F groups in PVdF. Improved ion mobility and ionic conductivity are therefore expected. Further, the presence of PDPA can also reduce the bulk resistance of the electrolyte. 2. Experimental 2.1. Preparation of polymer electrolytes Polydiphenylamine was prepared by the oxidative polymerization of diphenylamine (50 mM in 1 M NSA) with potassium peroxodisulfate (0.2 M in 1 M NSA) at 5 ◦ C [16]. The green colored precipitate (NSA-doped PDPA) was filtered, washed with 1 M NSA and dried in vacuum oven. Adequate amounts of PVdF and NSAdoped PDPA were dissolved in DMF/acetone mixture (7:3 v/v). Electrospun composite membranes were prepared by using the methodology, as described elsewhere [15]. Electrospinning of the composite solution was performed at a flow rate of 10 mL/h with a potential difference of 25 kV. A distance of 15 cm was kept between the syringe tip and collector. Membranes were accumulated on the collector (drum) over the aluminium foil. Electrospun membranes with different amounts of PDPA (0.5%, 1% and 2% w/w) were made. Also, membranes with different thickness ranging from 20 ␮m to 60 ␮m were made by manipulating the time for the electrospinning. Polymer electrolyte was prepared by soaking the electrospun PVdF–PDPA-CFM in 1 M LiClO4 –PC solution at 25 ◦ C in the glove box. The methodology adapted for the fabrication of PVdF–PDPA-CFM is presented in Scheme 1. 2.2. Membrane characterization The morphology of the membranes was examined by field emission scanning electron microscope (FESEM)-Hitachi S-4300 with a field emission gun operated at 200 kV. Fourier transform infrared spectra were recorded at ambient temperature using Bruker IFS 66v FT-IR spectrophotometer with a wave number resolution of 4 cm−1 .

423

Samples for FT-IR were made by casting the membrane directly on KBr pellets and then simultaneously dried at 120 ◦ C for 48 h. 2.3. Electrochemical characterization The polymer electrolyte membranes for electro chemical measurements were obtained by soaking the electrospun PVdF–PDPA-CFM in 1 M LiClO4 –PC solution for 24 h at 25 ◦ C. Ionic conductivities of the PVdF–PDPA-CFM were measured by recording the ac impedance spectra of the symmetric SUS316 cell (area: 3.14 cm2 ) in the temperature range of 0–80 ◦ C (sweep: 100 Hz to 0.1 MHz and ac amplitude: 5 mV). The symmetrical cell was then fixed in an airtight double wall glass tube, through the outer jacket of which thermostated water was circulated for measurements at different temperatures. All electrochemical measurements were performed using EG and G PAR 283 Potentiostat/Galvanostat with FRA 1025. For electrochemical measurements, a cell was constructed by sandwiching the electrospun membrane electrolyte between two symmetrical lithium metal electrodes. Cyclic voltammetry (CV) and linear sweep voltammetry (LSV) were performed at a sweep rate of 1 mV s−1 . For dc polarization measurement, a small constant potential difference of about 10 mV was applied across the cell and the current was measured as a function of time until it reaches a constant value. By considering the potential drop occurring at surface layers on the electrode, the cation transport number, t+ was determined. ac impedance spectra were recorded before and after the current relaxation measurement without interruption of the dc bias [17]. For charge–discharge studies, pouch cell battery prototypes were assembled by laminating in sequence the three components, i.e. the graphite anode, PVdF–PDPA-CFM as a electrolyte (typically of 60-␮m thickness) and the LiCoO2 cathode. 3. Results and discussion 3.1. Morphology and structure Fig. 1 shows the FESEM images of electrospun fibrous membranes depicting the morphological variations between pristine

Scheme 1. Schematic representation for the fabrication of PVdF–PDPA-CFM.

424

A.I. Gopalan et al. / Journal of Membrane Science 318 (2008) 422–428

Fig. 2. FT-IR spectra of electrospun PVdF (a), pristine PDPA and PVdF–PDPA-CFM (0.5% PDPA).

Fig. 1. FESEM images of electrospun PVdF (a), PVdF–PDPA-CFM with PDPA wt.%: 0.5 (b) and 1 (c).

PVdF membrane and PVdF–PDPA-CFM. PVdF membrane is white in color, whilst PVdF–PDPA-CFM is green in color (viewed through optical microscope). FESEM images clearly show the morphological variations between pristine PVdF membrane and PVdF–PDPA-CFM. Electrospun PVdF fibrous membranes have a nearly straightened and tubular structure with an average diameter of ∼500 nm (Fig. 1a). However, PVdF–PDPA-CFM has interconnected multifibrous layers with ultrafine porous structures (Fig. 1b and c). The average diameter of PVdF–PDPA-CFM (∼200 nm) with 0.5 wt.% of PDPA in the composite (Fig. 1b) is far lesser than pristine PVdF membrane (Fig. 1a). There is inter-fiber twisting in the PVdF–PDPACFM that generates microcavities. PVdF–PDPA-CFM with 1 wt.% of PDPA in the composite has much lower (<200 nm) fiber diameters and more entanglements between the fibers (Fig. 1c). The fibers

appeared to be uniform in composition without having any phase separated microstructure or beads which are due to the miscibility of PVdF and PDPA. Now, it becomes relevant to analyze the reason for the differences in morphology between PVdF membrane and PVdF–PDPACFM. It is generally known that the distance between the nozzle of the syringe and the collector, applied voltage, viscosity of the polymer solution and dielectric constant of the solvent are the parameters that influence the morphology of electrospun fibers. In the present study, the first three factors in the above list were maintained constant while electrospinning and hence cannot be the reason for the morphological differences. Hence, we envisage the following reasons. The incorporation of PDPA can increase the dielectric of the medium for electrospinning. Electrospin jets are therefore easily formed (in comparison to a lesser dielectric PVdF solution) at the nozzle of the syringe for PVdF–PDPA composite solution and cause formation of fibers with lower diameter without beads from the composite solution [15]. Interconnected fibrous morphology for the PVdF–PDPA-CFM is expected to arise from the probable intermolecular interactions between the protonated amine or imine groups in PDPA [7,8] and electronegative fluorine atoms in PVdF. FT-IR spectrum of PVdF–PDPA-CFM (Fig. 2) shows main bands corresponding to the oxidized form of PDPA consisting of diphenoquinodimine (quinoid) and diphenyl benzidine (benzenoid) structures [7,8]. However, the benzenoid and quinoid C N vibrational bands that appeared 1494 cm−1 and 1595 cm−1 , respectively for PDPA are found to be shifted to 1498 cm−1 and 1598 cm−1 in the PVdF–PDPA-CFM. Similarly, the band corresponding to CF2 bending mode of vibration that appeared in 1400 cm−1 for PVdF [18,19] is found to be shifted to 1375 cm−1 in the PVdF–PDPA-CFM. Further, new vibrational bands are found around 1670 cm−1 and 1734 cm−1 for the composite. The above observations indicate that there are molecular level interactions between the nitrogen atoms in benzenoid or quiniod structures of PDPA and electronegative fluorine atoms in PVdF in the PVdF–PDPA-CFM. 3.2. Swelling behaviour PVdF–PDPA-CFM was transformed into polymer electrolyte membrane by soaking the porous mats into an electrolyte solution (1 M LiClO4 in propylene carbonate) (Fig. 3). PVdF–PDPA-CFM with 2% PDPA (w/w) has higher uptake (>280 wt.%) of liquid electrolyte

A.I. Gopalan et al. / Journal of Membrane Science 318 (2008) 422–428

425

Fig. 3. Swelling characteristics of electrospun PVdF (a) and PVdF–PDPA-CFM (2% PDPA) (b) in 1 M LiClO4 –PC.

over PVdF membrane (∼200 wt.%). To the best of our knowledge, PVdF–PDPA-CFM has the highest liquid uptake for PVdF-based polymer electrolyte [20–24]. A maximum of 70% liquid uptake for PVdF and PVdF–HFP membranes has been reported [20,21]. For PVdF–HFP and PVdF–HFP-g-PMMA membranes, a liquid uptake of about 45% and 75%, respectively have been reported [22,23]. A maximum liquid uptake of about 260 wt.% for electrospun PVdF has been reported [24]. At the swelled state, PVdF–PDPA-CFM showed good dimensional stability. The higher liquid uptake for PVdF–PDPA-CFM mainly arises from the interconnected network morphology that increased the stability of pores. 3.3. Ionic conductivity The ionic conductivity of a polymer electrolyte membrane depends on the effective number of carrier ions and the ion mobility. The effective number of carrier ions is related to the concentration of the dissolved ions. The ion mobility in a polymer electrolyte formed by the dissolution of ions in polymer is facilitated by the segment mobility of the polymer chains. Ionic conductivities of the PVdF–PDPA-CFM were determined by ac impedance spectroscopy as a function of temperature. The mechanism of ionic conduction for PVdF–PDPA-CFM is mainly in accordance with Vogel–Tammann–Fulcher (VTF) relationship. VTF relationship to ion transport in PVdF–PDPA-CFM signifies the coupling of the charge carriers with the segmental motion of the polymer chains. PVdF–PDPA-CFM with 2% PDPA (60-␮m thickness) exhibits an ionic conductivity of 3.6 mS cm−1 at 25 ◦ C, the highest reported value for PVdF-based electrolyte [24–28]. The high ionic conductivity for PVdF–PDPA-CFM (2% PDPA) is attributed to arise from the combined influence of higher content of electrolyte (>280 wt.%) incorporated into the pores of the membrane and augmented lithium ion mobility in the membrane. Doped PDPA has positively charged nitrogen sites (protonated diphenoquinone diimine units) [7,8] that have molecular level interactions with the electronegative fluorine atoms present in PVdF. This environment provides (i) new path for Li+ ion migration in the composite, (ii) interconnected network morphology between PDPA and PVdF and (iii) compact porous structure to hold more amount of the liquid electrolyte. The recombination possibility of Li+ ion with ClO4 − ion is expected to be hindered in the presence of PDPA and hence facile Li+ ion mobility becomes possible. A portion of the ClO4 − ions is expected to be immobilized at the protonated amine or imine sites of PDPA by replacing the organic dopant, naphthalene sulfonate anions. The bulky naphthalene sulfonate anion has lesser mobility than ClO4 − ion. As a result, Li+ ions in the PVdF–PDPA membrane matrix can freely move to result higher ionic conductivity.

Fig. 4. Electrochemical stability of the PVdF–PDPA-CFM performed by linear sweep voltammetry, PVdF membrane (a), PVdF–PDPA-CFM with PDPA wt.%: 0.5 (b), 1 (c) and (d) 2.

3.4. Anodic stability The electrochemical stability window of the polymer electrolyte is one of the important parameters that must be considered for application in rechargeable lithium batteries. This parameter can be determined by means of a linear sweep voltammetry. Generally, LSV is performed by the fabrication of a laminated electrode cell with two inert blocking electrodes at an ambient temperature. The blocking electrode is polarized anodically and the electrochemical stability is identified by the rapid increase of the anodic current observed when the decomposition of the polymer electrolyte, i.e. typically the oxidation of the electrolyte anion, takes place. The working potential range or practical lithium rechargeable batteries are generally between +1.8 V and +3.5 V vs. Li. In the present study, to ascertain the electrochemical stability of PVdF–PDPA-CFM, LSV was performed. Fig. 4 shows the current–voltage response obtained for PVdF–PDPA-CFM electrolytes. The onset of current flow is associated with the decomposition voltage of the electrolyte. From the magnitude of the current response, the decomposition voltage of PVdF–PDPA-CFM was found to be 5.0 V, 5.1 V and 5.18 V vs. Li (PDPA: 0.5%, 1% and 2%, respectively). It can be seen that the electrochemical stability of the membrane is influenced by the amount of PDPA content. LSV measurements of PVdF–PDPA-CFM reveals an increased anodic limit voltage (5.18 V), which is higher than the electrospun PVdF membrane (4.6 V; Fig. 4a) and reported values of PVdF-based electrolytes [24,29–31]. The composite nanofibrous structure increases the oxidation potential of liquid electrolyte by ∼1.6 V (generally 3.6 V vs. Li) and gives a higher anodic stability to the electrolyte membrane. The negligible contribution of current beyond 4.3 V for the PVdF–PDPACFM arises from the electronic contribution of PDPA. In the present study, the presence of PDPA has a beneficial effect to widen the electrochemical stability window of the resultant electrolyte membrane. 3.5. Reversibility of Li/Li+ couple Cyclic voltammetry was performed to study the reversibility of Li deposition–stripping process of the PVdF–PDPA-CFM. Fig. 5 shows the repetitive cyclic voltammograms for Li/PVdF–PDPACFM/Li electrode cell performed at an ambient temperature. The sweep rate was kept as 1 mV s−1 . The onset potential for Li depo-

426

A.I. Gopalan et al. / Journal of Membrane Science 318 (2008) 422–428 Table 1 Lithium transport number of PVdF–PDPA-CFM Membranes

Ii (mA)

Iss (mA)

Ri (k)

Rss (k)

t+

PVdF

0.89

0.42

2.90

2.98

0.40

PVdF–PDPA-CFM PDPA: 0.5% PDPA: 1.0% PDPA: 2.0%

0.80 0.60 0.42

0.40 0.30 0.21

2.80 2.32 2.10

2.85 2.40 2.35

0.43 0.46 0.48

ing between 0.2 and 0.4 has been reported for PVdF–HFP-based electrolyte membrane [32–35]. 3.7. Interfacial behaviour

Fig. 5. Cyclic voltammograms of PVdF–PDPA-CFM with 2% PDPA.

sition was about −0.4 V and that can be taken as cathodic limit for the membrane electrolyte. On sweeping the electrode potential, a cathodic peak was observed at around −0.4 V, which corresponds to the plating of lithium onto the electrode. On the reverse scan, stripping of lithium was observed at around 0.38 V. The high currents along with peaks in both anodic and cathodic direction indicate the establishment of reversibility of Li/Li+ couple. The reaction Li+ + e−  Li

Compatibility with the electrode materials is an essential parameter in the cyclability and the reliability of rechargeable lithium battery. ac impedance studies performed for the symmetrical non-blocking Li/PVdF–PDPA-CFM/Li and Li/PVdF membrane/Li cells under open-circuit conditions provide useful information on the charge transfer and diffusion of lithium ions [36]. The intercept at high frequency is assigned to bulk resistance (Rb ) and the intercept at the lower frequency side corresponds to resistance at the interface arising from the passivation of membrane (Ri ) at the electrode. PVdF electrospun membrane has Ri as 0.46 k (at 25 ◦ C), whereas the PVdF–PDPA-CFM (2% PDPA) shows a lower Ri of about 0.32 k (at 25 ◦ C) (Fig. 6a). Further, it can be seen that the Ri increases as a function of time (aging time) and decreases with PDPA content in the membrane. The increase is Ri can be attributed to the growth of a passivating film on the lithium surface.

is facile at the interface between the lithium electrode and PVdF–PDPA-CFM. The separation in peak potentials being large may be due to the fact that there is no reference electrode used. It is observed that, although the process remains reversible upon cycling, the amount of cycled lithium progressively decreases. This phenomenon may be attributed to the formation of a passive layer on the electrode, which is similar to those commonly experienced in liquid organic electrolytes. The formation of passive layer is also envisaged from the impedance response of the lithium electrode with PVdF–PDPA-CFM. The reversibility of the membrane demonstrates that the PVdF–PDPA-CFM is electrochemically stable and hence can be safely used as the polymer electrolyte in the rechargeable lithium batteries. 3.6. Transport number The transport number of an electrolyte is an important index of its conductive behaviour. For intercalation/de-intercalation of lithium ions throughout the host compound lattice, electrolytes with a Li+ transference number (t+ ) close to unity are desirable to avoid a concentration gradient during charge–discharge cycling. Therefore, evaluation of the transference number is important for characterizing electrolyte materials for lithium polymer battery applications. Cation transference number, t+ of the composite electrolyte was determined by the application of 10 mV dc potential across the test cell (Li/(PVdF–PDPA-CFM)/Li). The current decays immediately and asymptotically approaches steady state. The decrease in current may be due to the growth of passivating layers at the electrode [2,3]. The transference number of Li+ ions was determined and presented in Table 1. The t+ value of PVdF–PDPA-CFM with 2% PDPA was determined to be 0.48. The reported t+ values for SPEs range from 0.06 to 0.2 [21]. For gel polymer systems, t+ values in the order of 0.4–0.5 have been reported. Cationic transport number rang-

Fig. 6. (i) Plot of interfacial resistance (Ri ) with PDPA content in PVdF–PDPA-CFM; (ii) ac impedance spectra of Li/ PVdF–PDPA-CFM (2% PDPA)/Li cell at different temperatures: (a) 25 ◦ C, (b) 30 ◦ C, (c) 40 ◦ C, (d) 50 ◦ C, (e) 60 ◦ C and (f) 70 ◦ C. Inset shows the influence of temperature on Ri and i0 .

A.I. Gopalan et al. / Journal of Membrane Science 318 (2008) 422–428

427

Table 2 Electrochemical properties of PVdF–PDPA-CFM Ionic conductivity at 25 ◦ C (mS cm−1 ) Anodic stability (vs. Li/Li+ ) (V) Transport number (t+ )

0.56 [25] 4.5 [24] 0.3 [32]

2.0 [26] 4.3 [27] 0.28 [33]

Generally, the C–F bonds in PVdF react with lithium or lithiated graphite to form stable LiF and/or >C CF-type unsaturated bonds [37] and cause instability to the negative electrode of the lithium and lithium ion batteries and also result thermal runaway problems. In the PVdF–PDPA-CFM, a portion of C–F bonds is intermolecularly bound to protonated amine or imine part of PDPA units and hence suppress the adverse effect of C–F interaction with Li electrode. Further, to understand the effect of temperature on the Li/PVdF–PDPA-CFM interface, impedance measurements were conducted at different temperatures between 25 ◦ C and 70 ◦ C (Fig. 6b). The Nyquist plots reveal that the interfacial resistance (Ri ) decreases with temperature and consequently the exchange current density, i0 , increases with temperature. This indicates that the properties of the interfacial film at higher temperatures are conducive for a facile charge transfer and this may be based on the breakdown of the passivating film. It is obvious from the impedance results that the PVdF–PDPA-CFM exhibits good compatibility with lithium metal electrode. PVdF–PDPA-CFM exhibited improved performance characteristics than electrospun PVdF and other PVdF-based polymer electrolyte membranes. A comparison with respect to the electrochemical characteristics of PVdF–PDPA-CFM with other PVdF-based electrolytes was made and presented in Table 2. 3.8. Battery performances In order to explore the usefulness of the PVdF–PDPA-CFM as electrolyte in rechargeable lithium batteries, pouch cells consisting of graphite as anode, LiCoO2 as cathode and PVdF–PDPA-CFM as an electrolyte were constructed and the performance was evaluated. A typical charge and discharge curves of the pouch cell with PVdF–PDPA-CFM (2% PDPA) at 25 ◦ C is shown in Fig. 7. The cell was cycled between the cut-off voltages, 4.2 V and 3.0 V at 1 C rate. The charge–discharge profile is similar to what is observed in lithium ion batteries in general [38], which ensures a good contact between the electrodes and PVdF–PDPA-CFM. An open-circuit voltage as high as 4.15 V is achieved at 25 ◦ C and the charging–discharging voltages are quite reproducible over the entire set of cycles. Fig. 8a shows the discharge curves for the cells with PVdF–PDPACFM (PDPA: 1% and 2%) as electrolyte at various rates. A high-rate capability is realized in the cell with PVdF–PDPA-CFM as an electrolyte, which is due to the high ionic conductivity and highly porous morphology of PVdF–PDPA-CFM. A good fraction of the capacity is delivered at 0.2 C and the cell is still able to operate even

1.74 [27] 4.7 [30] 0.34 [34]

2.0 [28] 5.0 [31] 0.34 [35]

3.6 [PVdF–PDPA-CFM] 5.18 [PVdF–PDPA-CFM] 0.48 [PVdF–PDPA-CFM]

Fig. 8. (a) Discharge voltage profiles of pouch cell at various rates with PVdF–PDPACFM (PDPA: 1% and 2%) as electrolyte; anode, graphite; cathode, LiCoO2 ; temperature, 25 ◦ C. The pouch cell was charged at 0.5 C rate. (b) Capacity vs. no. of cycles of the pouch cell PVdF–PDPA-CFM (PDPA 2%) as an electrolyte; anode, graphite; cathode, LiCoO2 ; temperature, 25 ◦ C. Cell was cycled at the 1 C rate between 3.0 V and 4.2 V.

at 2 C. Typically, at 1 C rate, the cell can deliver about 90% of its 0.2 C capacity. Even at a high 2 C rate, the cell can still deliver about ∼70% of its 0.2 C capacity. Further, the cell was subjected to cycle tests with cut-off voltages of 4.2 V (upper limit) and 3.0 V (lower limit) at 1 C rate. Fig. 8b reveals that the cell with PVdF–PDPACFM as electrolyte has a very steady charge–discharge behaviour without significant loss in capacity under the voltage conditions (1C rate at 25 ◦ C). Also, a high charge–discharge efficiency (ratio between charge and discharge capacity) is observed, which confirms the good interfacial stability between the electrode and the PVdF–PDPA-CFM. Thus, the pouch cell with PVdF–PDPA-CFM as an electrolyte shows low impedance, excellent rate capability and good cyclability and displays a superior potential for rechargeable lithium batteries. 4. Conclusions PVdF–PDPA composite nanofibrous membranes were prepared by electrospinning and the polymer electrolytes were prepared by soaking the porous mats into an electrolyte solution. The polymer electrolyte based on PVdF–PDPA-CFM shows superior performances in terms of ionic conductivity, electrochemical stability window and good interfacial behaviour with electrode and proved to be a promising material component for high-performance lithium batteries. Acknowledgements

Fig. 7. Charge–discharge voltage profiles of graphite/PVdF–PDPA-CFM (2% PDPA)/LiCoO2 cell.

This work was supported by Korean Research Foundation Grant (KRF-2006-J02402 and KRF-2006-C00001). The authors acknowl-

428

A.I. Gopalan et al. / Journal of Membrane Science 318 (2008) 422–428

edge the Korea Basic Science Institute (Daegu) and Kyungpook National University Center for providing Scientific Instrumentation. We also thank Samsung SDI, CRD Energy Lab for recording charge-discharge profiles. References [1] D.E. Fenton, J.M. Parker, P.V. Wright, Complexes of alkali metal ions with poly(ethylene oxide), Polymer 14 (1973) 589. [2] P. Santhosh, T. Vasudevan, A. Gopalan, K.P. Lee, Preparation and characterization of polyurethane/poly(vinylidene fluoride) composites and evaluation as polymer electrolytes, Mater. Sci. Eng. B 135 (2006) 65. [3] P. Santhosh, T. Vasudevan, A. Gopalan, K.P. Lee, Preparation and properties of new cross-linked polyurethane acrylate electrolytes for lithium batteries, J. Power Sources 160 (2006) 609. [4] S.S. Sekhon, H.P. Singh, Ionic conductivity of PVdF-based polymer gel electrolytes, Solid State Ionics 152–153 (2002) 169. [5] E. Quartarone, P. Mustarelli, A. Magistris, Transport properties of porous PVDF membranes, J Phys. Chem. B 106 (2002) 10828. [6] T.C. Wen, J.B. Chen, A. Gopalan, Soluble and methane sulfonic acid doped poly(diphenylamine)-synthesis and characterization, Mater. Lett. 57 (2002) 280. [7] A.V. Orlov, S.Z. Ozkan, G.P. Karpacheva, Oxidative polymerization of diphenylamine: a mechanistic study, Polym. Sci. Ser. B 48 (2006) 11. [8] P. Santhosh, M. Sankarasubramanian, M. Thanneermalai, A. Gopalan, T. Vasudevan, Electrochemical, spectroelectrochemical and spectroscopic evidences for copolymer formation between diphenylamine and m-toluidine, Mater. Chem. Phys. 85 (2004) 316. [9] F. Hua, E. Ruckenstein, Water-soluble conducting poly(ethylene oxide)-grafted polydiphenylamine: synthesis through a “graft onto” process, Macromolecule 36 (2003) 9971. [10] S. Adam, A fine set of threads, Nature 411 (2001) 236. [11] K.M. Manesh, H.T. Kim, P. Santhosh, A.I. Gopalan, K.P. Lee, A novel glucose biosensor based on immobilization of glucose oxidase into multiwall carbon nanotubes-polyelectrolyte-loaded electrospun nanofibrous membrane, Biosens. Bioelectron. 23 (2008) 771. [12] K.M. Manesh, A.I. Gopalan, K.P. Lee, P. Santhosh, K.D. Song, D.D. Lee, Fabrication of functional nanofibrous ammonia sensor, IEEE Trans. Nanotechnol. 6 (2007) 513. [13] H.S. Kim, H.J. Jin, S.J. Myung, M. Kang, I.J. Chin, Carbon nanotube-adsorbed electrospun nanofibrous membranes of Nylon 6, Macromol. Rapid Commun. 27 (2006) 146. [14] I.D. Norris, M.M. Shaker, K.K. Frank, A.G. MacDiarmid, Electrostatic fabrication of ultrafine conducting fibers: polyaniline/polyethylene oxide blends, Synth. Met. 114 (2000) 109. [15] K.M. Manesh, P. Santhosh, A. Gopalan, K.P. Lee, Electrospun poly(vinylidene fluoride)/poly(aminophenylboronic acid) composite nanofibrous membrane as a novel glucose sensor, Anal. Biochem. 360 (2007) 189. [16] S. Nagarajan, P. Santhosh, M. Sankarasubramanian, T. Vasudevan, A. Gopalan, K.P. Lee, UV–vis spectroscopy for following the kinetics of homogeneous polymerization of diphenylamine in p-toluene sulphonic acid, Spectrochim. Acta A 62 (2005) 420. [17] P. Santhosh, A. Gopalan, T. Vasudevan, K.P. Lee, Evaluation of a cross-linked polyurethane acrylate as polymer electrolyte for lithium batteries, Mater. Res. Bull. 41 (2006) 1023. [18] Y. Bormashenko, R. Pogreb, O. Stanevsky, E. Bormashenko, Vibrational spectrum of PVDF and its interpretation, Polym. Test. 23 (2004) 791.

[19] Y.J. Shen, M.J. Reddy, P.P. Chu, Porous PVDF with LiClO4 complex as ‘solid’ and ‘wet’ polymer electrolyte, Solid State Ionics 175 (2004) 747. [20] A. Magistris, E. Quartarone, P. Mustarelli, Y. Saito, H. Kataoka, PVDF-based porous polymer electrolytes for lithium batteries, Solid State Ionics 152–153 (2002) 347. [21] T. Godz, C. Schumtz, J.M. Tarascon, P. Warran, US Patent 5,418,091. [22] Y. Liu, J.Y. Lee, L. Hong, Synthesis, characterization and electrochemical properties of poly(methyl methacrylate)-grafted-poly(vinylidene fluoridehexafluoropropylene) gel electrolytes, Solid State Ionics 150 (2002) 317. [23] J.M. Tarascon, A.S. Gozdz, C. Schmutz, F. Shokoohi, P.C. Warren, Performance of Bellcore’s plastic rechargeable Li-ion batteries, Solid State Ionics 86–88 (1996) 49. [24] S.W. Choi, S.M. Jo, W.S. Lee, Y.R. Kim, An electrospun poly(vinylidene fluoride) nanofibrous membrane and its battery applications, Adv. Mater. 15 (2003) 2027. [25] C.Y. Chiang, Y.J. Shen, M.J. Reddy, P.P. Chu, Complexation of poly(vinylidene fluoride):LiPF6 solid polymer electrolyte with enhanced ion conduction in ‘wet’ form, J. Power Sources 123 (2003) 222. [26] S. Panero, F. Ciuffa, A. D’Epifano, B. Scrosati, New concepts for the development of lithium and proton conducting membranes, Electrochim. Acta 48 (2003) 2009. [27] H.S. Choe, J. Giaccai, M. Alamgir, K.M. Abraham, Preparation and characterization of poly(vinyl sulfone)- and poly(vinylidene fluoride)-based electrolytes, Electrochim. Acta 40 (1995) 2289. [28] S.S. Choi, Y.S. Lee, C.W. Joo, S.G. Lee, J.K. Park, H.S. Han, Electrospun PVDF nanofiber web as polymer electrolyte or separator, Electrochim. Acta 50 (2004) 339. [29] J.R. Kim, S.W. Choi, S.M. Jo, W.S. Lee, B.C. Kim, Electrospun PVdF-based fibrous polymer electrolytes for lithium ion polymer batteries, Electrochim. Acta 50 (2004) 69. [30] G.B. Appetecchi, F. Croce, A.D. Paolis, B. Scrosati, A poly(vinylidene fluoride)based gel electrolyte membrane for lithium batteries, J. Electroanal. Chem. 463 (1999) 248. [31] Z. Wang, Z. Tang, Characterization of the polymer electrolyte based on the blend of poly(vinylidene fluoride-co-hexafluoropropylene) and poly(vinyl pyrrolidone) for lithium ion battery, Mater. Chem. Phys. 82 (2003) 16. [32] E. Quartarone, M. Brusa, P. Mustarelli, C. Tomasi, A. Magistris, Preparation and characterization of fluorinated hybrid electrolytes, Electrochim. Acta 44 (1988) 677. [33] M. Stolarska, L. Niedzicki, R. Borkowska, A. Zalewska, W. Wieczorek, Structure, transport properties and interfacial stability of PVdF/HFP electrolytes containing modified inorganic filler, Electrochim. Acta 53 (2007) 1512. [34] H. Dai, T.A. Zawodzinski, The dependence of lithium transference numbers on temperature, salt concentration and anion type in poly(vinylidene fluoride)hexafluoropropylene copolymer-based gel electrolytes, J. Electroanal. Chem. 459 (1998) 111. [35] H. Kataoka, Y. Saito, Y. Miyazaki, D. Deki, Ionic mobilities of PVDF-based polymer gel electrolytes as studied by direct current NMR, Solid State Ionics 152–153 (2002) 175. [36] K.P. Lee, A.I. Gopalan, K.M. Manesh, P. Santhosh, K.S. Kim, Influence of finely dispersed carbon nanotubes on the performance characteristics of polymer electrolytes for lithium batteries, IEEE Trans. Nanotechnol. 6 (2007) 362. [37] A.D. Pasquier, F. Disma, T. Bowma, A.S. Gozdz, G.J.M. Amatucci, Differential scanning calorimetry study of the reactivity of carbon anodes in plastic Li-ion batteries, J. Electrochem. Soc. 145 (1988) 472. [38] L.J. Fu, H. Liu, C. Li, Y.P. Wu, E. Rahm, R. Holze, H.Q. Wu, Electrode materials for lithium secondary batteries prepared by sol–gel methods, Prog. Mater. Sci. 50 (2005) 881.

polydiphenylamine composite electrospun membrane ...

a Department of Chemistry Graduate School, Kyungpook National University, Daegu 702-701, South Korea b Nano ... Electrospinning technology has recently been extended ..... cells under open-circuit conditions provide useful information on.

736KB Sizes 0 Downloads 220 Views

Recommend Documents

An inorganic composite membrane as the separator of ...
Available online 6 October 2004. Abstract ... However, high cost and safety concerns still set restrictions to the above ... +1 301 394 0981; fax: +1 301 394 0273.

Electrospun Nanofibers Modified with Phospholipid ...
purified by vacuum distillation before use. Lipase (from. Candida rugosa, protein concentration was 6.8 wt.-%), Brad- ford reagent, bovine serum albumin (BSA, ...

Electrospun for Redox Enzyme Immobilization
ever, for redox enzymes such as catalase, a direct electron- transfer path should be .... protoporphyrin ring and a central Fe atom, i.e., ferriproto- porphyrin, where ...

Composite ratchet wrench
Mar 2, 1999 - outwardly extending annular ?ange 33. Formed in the outer surface of the body 31 adjacent to the drive lug 32 is a circumferential groove 34.

Breakable composite drill screw
Apr. 11, 1991 .... first shank and at the same time to use a low-carbon steel as a material of the second ... kind, for instance of a low-carbon steel which is suscep.

Electrospun nanofibrous membranes filled with carbon ...
over their free forms, which include easy recovery and possible reuse, improved stability and ..... both partition and diffusion effects. The higher Km value of.

Membrane and Desalination Technologies.pdf
and bridges 80% 20%. 3. Whoops! There was a problem loading this page. Retrying... Main menu. Displaying Membrane and Desalination Technologies.pdf.

CO2 separation using bipolar membrane electrodialysis
Dec 7, 2010 - that the energy consumption required to regenerate CO2 gas from aqueous bicarbonate (carbonate) ... effective technologies for controlling the atmospheric CO2 ... CO2 emitted in the United States.3 Moreover, since combusting ... alterna

Addax Composite Couplings.pdf
Addax Composite Couplings.pdf. Addax Composite Couplings.pdf. Open. Extract. Open with. Sign In. Main menu. Displaying Addax Composite Couplings.pdf.

Multilayer reverse osmosis membrane of polyamide-urea
Oct 29, 1991 - support of polyester sailcloth. The solution is cast at a knife clearance of 5.5 mi]. The sailcloth bearing the cast polyethersulfone solution is ...

Addax Composite Couplings.pdf
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. Addax ...

An electrospun triphasic nanofibrous scaffold for bone ...
May 8, 2007 - 80 nm particle size) into the solution mixture for co- electrospinning [26]. ... synthesized by chemical precipitation at room temperature using Ca(OH)2 and ... were fully enclosed within a clear Lexan box in order to minimize ...

Advance-Membrane-Technology-and-Application.pdf
W. S. Winston Ho, and T. Matsuura. Page 3 of 989. Advance-Membrane-Technology-and-Application.pdf. Advance-Membrane-Technology-and-Application.pdf.

Membrane Technology and Applications 2nd Edition.pdf ...
3. Whoops! There was a problem loading this page. Retrying... Main menu. Displaying Membrane Technology and Applications 2nd Edition.pdf. Page 1 of 50.

Chapter 7: Membrane Structure and Function - WordPress.com
First, inspection of a variety of membranes revealed that membranes with different functions differ in structure and chemical composition. A second, more serious problem became apparent once membrane proteins were better characterized. Unlike protein

Cheap Pc Laptop Composite Video Tv Rca Composite S-Video Av ...
Cheap Pc Laptop Composite Video Tv Rca Composite S- ... pter Switch Box Free Shipping & Wholesale Price.pdf. Cheap Pc Laptop Composite Video Tv Rca ...

2003_Modeling of Cement Based Composite ...
2003_Modeling of Cement Based Composite Laminates_AnnArbor_HPFRCC4-AnytoPDF.pdf. 2003_Modeling of Cement Based Composite ...

A.S.T.M.25, Composite Materials.pdf
... including carbon-carbon composites (C-C) are. covered in Volume 4 and Volume 5 , respectively. 5. This standardization handbook has been developed and ...

May, 2004 CERAMICS AND COMPOSITE MATERIALS
7.a) Describe production of fiber reinforced polymers and their applications. b) What are the requirement of fibers and matrix? What is the effect of.