You are on page 1of 23

Spectroscopic Investigation of Interactions among Components and Ion Transport Mechanism in Polyacrylonitrile Based Electrolytes

Zhaoxiang Wang, Biying Huang, Rongjian Xue, Xuejie Huang, and Liquan Chen* Laboratory for Solid state Ionics, Institute of Physics, Chinese Academy of Sciences P.O.Box 603-29, Beijing 100080, China

*Correspondence author

Abstract
Raman, infrared (IR), fluorescence, X-ray photoelectron (XPS), and nuclear magnetic resonance (NMR) spectra of various combinations of the components of the polyacrylonitrile-based electrolytes have been studied. It is found that the so-called composite- or mixed-phase polymer electrolytes are not simple mixtures of electrolytic salt, plasticizer and polymer; instead there are very strong interactions between each two of the components of the electrolytes though different systems may show different interaction characteristics. By analyzing the important functions of the plasticizer, a possible mechanism of the ion transport in the polymer electrolyte is proposed.

Introduction

Since the discovery that a number of polymer-salt systems exhibit considerable conductivity1-3, much research effort has been made to find the optimal combination of host polymer and dopant salt for fast ionic transport. In 1990, Abraham et al. 4 synthesized polyacrylonitrile (PAN)-based lithium-salt electrolytes with ionic conductivities higher than 10-3 S/cm at room temperature, breaking through the prediction of 10-4 S/cm. These electrolytes are ionically conducting, plasticizer-dissolved salt in a polymer matrix. The polymer contributes the mechanical strength and the plasticizer-dissolved salt allows the relatively high conductivity at ambient temperature. These systems combine the desirable mechanical properties of polymers (e.g. low density, high flexibility, ease of fabrication, etc.) and the potential of high conductivity of the liquid electrolytes. This laboratory has been studying PAN-based electrolytes by way of co-complex or plasticizing. A good deal of knowledge of the synthesis of these new polymer electrolytes has thus been obtained, especially concerning the factors that promote ionic mobility. Huang5 has synthesized a dozen of PAN-based solid electrolytes containing plasticizers, including propylene carbonate (PC), ethylene carbonate (EC), dimethylformamide (DMF), dimethylsulfoxide (DMSO), sulfolane (SL), -butyrolactone (BL), etc., and their mixtures. It was found that these electrolytes had conductivities as high as 2.5x10-3 S/cm at room temperature and were compatible with metallic lithium quite well. Moreover, these polymer electrolytes still show good elasticity after kept in water-proof environment for one year. Such polymer electrolytes are promising candidates of electrolytes in lithium batteries used with many applications6. As to the functions of the plasticizer in the polymer electrolyte system, there have been various suppositions. Some authors suppose that the only functions of the plasticizer are to decrease the glass transition temperature, dissolve the salt, make the polymer become amorphous and therefore allow the ions or ionic carriers to move freely so as to raise the ionic conductivity of the electrolyte. Other authors believe, however, that besides the plasticizing effect, one more important function of the plasticizer is to associate with the ionic carriers and allow them to move faster. To date, most researchers treat the polymer electrolytes containing plasticizer as mixed phases or composite electrolytes. This means that the interactions between the components of the polymer electrolytes have been neglected. Strong Coulombic attractions between the dissolved salt ions result in formation of ion pairs, triplets, and even multiples with higher degrees of association in concentrated (polymer) electrolytes. Since neutral ion pairs or neutral multiples do not contribute to the conductivity, the number of charge carriers will not be proportional to the stoichiometric concentration of the salt. As a result, the conductivity is reduced. In

addition, the solution becomes more viscous with the addition of the salt. In turn, the higher viscosity reduces the mobility of charge carriers, leading to a further drop of the ionic conductivity. In this paper, the recent experimental results of this laboratory are summarized on the interactions between each two components of the PAN-based electrolytes by Raman, infrared (IR), fluorescence, X-ray photoelectron (XPS) and nuclear magnetic resonance (NMR) spectroscopy. By analyzing the functions of the plasticizers, a possible transport mechanism of the ions in the PAN-based electrolytes is proposed.

Experimental
Preparation of the samples. The reagents used in the experiments, EC, PC, DMF, DMSO, and LiClO4 (all from Aldrich), and PAN (Fushun, China) were analytically pure. The solvents were distilled whereas the salt and the polymer reagents were dried in a vacuum oven prior to use. Then they were mixed thoroughly at the required molar ratios, sealed in a bottle and heated at 110C for 2 hours for the liquid electrolytes but for 12 hours for the polymer eletrolytes, and subsequently cooled down to room temperature (20C). Raman and IR spectrometers and the experimental geometry The IR and Raman spectra were all recorded with an IFS66+FRA106 type Fourier transform IRRaman spectrometer (Bruker Instruments Corporation, Germany). The laser power on the sample was 200 mW with a radiation spot of 0.1 mm in diameter. The excitation line was 1.064 m. The scattered light was collected in a direction of 180 to the incident light (back-scattering geometry). Each of the Raman spectra was the average of 200 scans while the IR spectra were average of 16 scans, at a speed of two seconds per scan. The resolution was 4 cm-1 for both the Raman and IR spectra. A germanium detector used to measure the scattered light was cooled in liquid nitrogen. An OS/2 computer operating system was used to collect and process the data. For the liquid samples, optical glass test tubes were used for the Raman spectra while for the gel samples the Raman spectra were recorded directly. The gel samples were pressed into very thin films (less than 10 m) and supported with a KBr pellet for the IR measurement.

Results and Discussion 1. Interactions between LiClO4 and plasticizers

A. Interactions between Li + ion and Plasticizers Fig.1 shows the Raman spectra of ethylene carbonate (EC) with different concentrations of LiClO4. It can be seen that the spectral changes of the ring bending mode of EC at 715 cm-1 are obvious. With the increase of the LiClO4 concentration, a shoulder appears at 727 cm-1 at first and its relative intensity increases with increasing solution concentration. Finally the shoulder becomes so strong that it dominates the ring bending band. This agrees with the IR spectra of the mode. Similar band splitting phenomena have also been observed in the IR and Raman spectra of the ring breathing mode (the mother peak at 893 cm-1 is split into 893 and 903 cm-1 components in concentrated solutions). It is well known that the solvation effect of the ClO4 anion is very weak in nonaqueous solutions7 due to its large size and will not influence the Raman or IR spectra of EC significantly. Therefore, the band splitting phenomena observed here should be ascribed to the interaction between the Li+ ion from LiClO4 and the oxygen atoms of the ring group of EC (including both the C=O and C-O bonds). Considering other spectral changes, such as the C=O stretching band (moves from 1771 cm-1 in pure EC to 1776 cm-1 in concentrated solutions) in the IR spectra, we assume that the interactions mainly occur between the C=O group and Li + ion. This viewpoint is also true of the following discussion of interaction between LiClO4 and other solvents. There are lone pairs of electrons on both the oxygen and the nitrogen atoms of the O=C-N group of dimethylformamide (DMF). When different contents of LiClO4 are added into DMF, it is found that a shoulder appears on the high-wavenumber side ( at 670 cm-1) in both the Raman and IR spectra of the O=C-N deformation band (at 658 cm-1 in pure DMF). This shoulder has a changing feature similar to the ring bending and ring breathing bands of EC containing LiClO4 but its change is more obvious (Fig.2). In addition, the CH3-N stretching band at 1095 cm-1 in pure DMF also moves to 1100 cm -1 in the Raman spectra of concentrated solutions. Changes in relative intensities of other bands related to the N or O atoms were also observed in the experiments. These indicate that Li+ ions can interact with both the oxygen and the nitrogen atoms of DMF in the system of LiClO4/DMF. Due to the severe self-association in dimethylsulfoxide (DMSO) by way of SLLO of neighboring DMSO molecules, most of the DMSO molecules exist in the form of dimers or polymers at room temperature. When LiClO 4 is introduced into DMSO, however, the interaction between LiClO4 and DMSO molecules destructs the cyclic dimers at first, making them become monomers. Then the Li+ ions interact

with the monomers, forming an ion-molecule associate. In the meantime, due to the interaction between Li+ ion and the S=O bond, the CH3 groups linked to the S atoms are also affected by way of an induction effect of Li+ ions transferred along the molecular chain. As a result, its Raman shift moves towards the highwavenumber side of the Raman spectrum with increasing content of LiClO4 (Fig.3). The above discussion has demonstrated that no matter the Li+ ions associate with the O or the N atoms of the solvents, the vibrational frequencies of the related structures increase. This is explainable. When Li+ ions associate with the O or N atoms of the O=C, S=O or CN (in PAN, see the following section) structures, their binding energies are expected to be lowered to some degree, which will bring a decrease in the vibrational frequencies of the groups. On the other hand, the Li+ ion association results in an even greater degree of reduction in the effective mass of the associates (=m Li 0/(mLi +0), where 0=C=O , S=O or NC, mLi <<0). Therefore, an even greater increase in frequency is expected (f /, is directly proportional to the binding energy or the strength of the chemical bonding). Considering these (at least) two opposite but coexisting facts in the structure, a net increase of vibrational frequency is reasonable. Summarizing the above Raman and IR results on LiClO4 dissolved in different solvents (plasticizers), a common feature can be seen. That is, the bands with the most striking spectral changes are all related to the C=O group (in EC and DMF) or the S=O group (in DMSO). Therefore, as has been reported repeatedly, the changes of their Raman shifts indicate unambiguously the interaction between LiClO4 and the solvents, or to be exact, between the Li+ ions and the C=O or S=O groups of them. Of course, due to the electrical negativity of the oxygen atom on the C-O or C-N bond in the structure of the solvents, the possibility cannot be not excluded that some of the Li+ ions may interact with the oxygen atoms on the EC rings. However, the interaction mainly occurs on the carbonyl or the sulfoxide group. B. Symmetry change of the ClO 4- and the formation of associates According to the group theory, ClO4 has four vibrational modes: 1 (A1) (at ca. 930 cm-1), 2 (E1) (at ca. 457 cm-1), 3 (F2) (at ca. 1016 cm-1) and 4 (F2) (at 624 cm -1). All these four modes are Raman active and only 3 and 4 are IR-active for a free ClO4 anion. However, when the structure of the ClO 4 anions are perturbed somehow, some of its vibrational modes may split and some of the inactive modes may be activated. In the present experiments, splittings of the 2 and 4 modes ( new components appear at 636 and 470 cm-1, respectively) are indeed observed in both the Raman and IR spectra and the activation of the 1 mode in the IR spectra (Fig.4). The splitting and activation of these vinbrational modes indicate that the

increase of the concentration of LiClO4 has resulted in perturbation to the ClO 4 anion itself and formation of ion associates, such as solvent-separated ion pairs, Li+-solvent-ClO4 (corresponding to the appearance of a new components at ca. 940 cm-1 for the 1 mode), contact ion pairs, [Li+ClO 4] (corresponding to the ca. 948 cm-1 component), and multiple ion aggregates, {Li +ClO 4}n (n2) (corresponding to the component at 955 cm-1) in the Raman spectrum. Increase of these aggregates will lead to decrease of the concentration of the charge carriers and increase of solution viscosity, which reduces the mobility of the charge carriers. Both of these effects will decrease the ionic conductivity of the electrolytes. C. Coordination number of Li+ ion in EC and the concentration of free ions in EC/LiClO4 solution. In order to determine concentrations of the free and associated solvent molecules from Raman spectra of different salt concentrations, the splitting and the intensity change of the ring bending band of EC (at 715 cm1

) was detected quantitatively. By checking the spectra of EC with different LiClO4 contents, the ring

stretching band of EC (at 1230 cm-1 ) was taken as the internal standard. By band-fitting, it is easy to deconvolute the ring-bending band and calculate the integrated intensity of each component relative to that of the standard band. Table 1 shows the concentration-dependent relative intensities of I715/I1230, I 725/I1230, and (I 715 + I725)/I1230. The intensity of the band at 715 cm-1 decreases while the intensity of the band at 725 cm-1 increases as the concentration of the salt increases. This fact supports the assignment of the two bands to two different species, i.e., bulk (or free) EC and associated EC. The total intensity, (I715 + I725)/I1230, decreases slowly in a large range of salt content. If Cf and C b are defined as the concentrations of the free and associated EC molecules in the electrolyte, respectively, and If and Ib are the relative intensities of the bands related to the free (I715/I1230) and the associated EC (I725/I1230) species, the following relationships can be obtained: I f = C fJ f I b = C bJb (1a) (1b)

where Jf and Jb are coefficients related to the molecular species and their scattering conditions. Thus, after Deng and Irish7, 8 , the total relative intensity of the band is I t = If + Ib = C fJf + C bJb = (1 - J b/J f)If + CJ b (2)

where C = Cf + C b is a constant ( for EC, C = 11.36 mol/Kg ).

Equation (2) indicates that the relation between It and If is linear, with (1-Jb/J f) as the slope and (CJb) as the intercept. Fig.5 is a plot of It vs If. It can be seen that the relation between It and If is roughly linear. Therefore, Fig.8 implies that 1-J b/J f = 0.238 CJ b = 1.386 As a result, Jf = 0.160 J b = 0.122 (4a) (4b) (3a) (3b)

The coordination number of the lithium cation, equivalent to the average ligand number of lithium, ns, is defined as n s = C b/C Li + = Ib/(CLi +Jb) (5) Based on Table 1 and eq.5 a coordination number of six is suggested for the lithium ions in EC. This value is larger than the coordination number of LiAsF6 in acetone and in methyl acetate or methyl formamide by Deng and Irish7, 8, and LiClO4 in EC by Hyodo and Okabayashi9, but is still within the range of values commonly found in lithium cation ionic crystals. The most favorable ns value for lithium ions are expected to be four, five and six10, 11 . Moreover, it is worth noticing that the calculated ns value decreases with the increase of the solution concentration. This is due to the assumption that all the LiClO4 has been solvated in EC when CLi + is replaced with CLiClO4. That is, C Li + in eq.5 should be the concentration of free Li+ cations instead of the concentration of LiClO4 in EC, C LiClO4. Due to the ion-pairing effect, CLi + is smaller than C LiClO4 and the difference between them increases with the solution concentration when ion pairing becomes more significant. Only in infinitely dilute solutions where the ion-pairing effect is negligible, can CLiClO4 and C Li + be equal. That is why a solvation number of six is suggested for Li+ in EC though smaller coordination numbers are also measured at concentrated LiClO4/EC solutions. Many authors have reported ion pairing or ion association of LiClO4 in both aqueous and non-aqueous solutions. However, it seems that few researchers have reported a quantitative dependence between the free anion content in the solution and the vibrational spectra of the ClO4 anions12, 13 . Analogous to the above calculation, the concentrations of the free and perturbed ClO 4 anions were evaluated. The solution concentration (mol/Kg) dependences of free and perturbed ClO 4 ions are shown in Fig.6. It can be seen that the concentration of the free anions has a maximum at [LiClO4] = 3 mol/Kg. The solution concentration dependence of the free ion concentration is comparable to the plot of solution concentration versus conductivity. For example, the conductivity maximum appears at ca. 1 M in

PC/LiClO 414 solution and at weight ratio of 5:20:75 in LiClO4:PAN:plasticizer (solvent)5. It is reasonable that the conductivity increases with the concentration of free ions. Another factor that influences the conductivity is the viscosity of the medium which monotonously increases with the increase of the salt concentration. Therefore the conductivity maximum is reached at a lower salt concentration than the maximum of free ion concentration in the same solution electrolyte.

2. Interactions between Polyacrylonitrile and Plasticizers


Analyzing the Raman spectra of pure EC, pure PAN and their mixtures with different molar ratios it is found that there is a striking influence of PAN on the spectra of EC. The spectral band that changes most significantly is the C=O stretching mode of EC at 1762 cm-1 . It changes from a medium peak in pure EC to a weak peak in a mixture of EC/PAN and its Raman shift moves 10 cm-1 towards the high frequency side of the spectra (Fig.7). In addition, its ring stretching mode moves from 1059 to 1072 cm-1, accompanied with decreasing intensity and increasing linewidth. These indicate that there is a rather strong interaction between PAN and EC. On the other hand, the influence of EC on the spectral features of PAN is not obvious. A plausible explanation is that the CN bond is much stronger than the C=O bond; an interaction of the same strength can bring different effects on different bonds. It is possible that this interaction occurs through the dipolar repulsive interaction of the C=O group of EC and the CN group of PAN, resulting in an increase of their polarities. When PAN is added into DMSO, the spectral changes of DMSO become obvious at a very low molar ratio of PAN:DMSO. A new component appears at 1070 cm-1 (corresponding to the monomers of DMSO) while the intensities of the components at 1013 and 1030 cm-1 (corresponding to the linear dimer and the polymers of DMSO, respectively) decrease. It is also observed that the linewidth of the S=O bending mode at 303 cm-1 increases with the increasing content of PAN in the mixture. Obvious spectral changes are also observed in other vibrational modes, such as the CSC stretching mode. These indicate that the influence of PAN on DMSO is by way of the S=O group and dissociates the linear dimers of the DMSO molecules. On the other hand, DMSO also exerts influences on the PAN molecules. This influence is not obvious at low content of PAN. When the molar ratio reaches 0.8, however, the Raman scattering cross section of the CN stretching mode (at 2240 cm-1) becomes small (Fig.8).

3. Interactions between LiClO4 and Polyacrylonitrile

For the convenience of comparison, the Raman spectra of EC/PAN at the mass ratio of 3:2 and of EC/LiClO4 at the mass ratio of 7:3, together with the Raman spectrum of EC/PAN/LiClO4 sample at the mass ratio of 50:24:26 are shown in Fig.9. Although the CN stretching band shows no obvious change in the Raman shift, a weak but obvious shoulder appears on the high frequency side of the CN stretching band. This shoulder cannot be assigned to EC or LiClO4 or to the interaction between EC and LiClO4 or between EC and PAN because no such a peak or shoulder is observed in this region in these systems. Therefore, it can be reasonably believed that the appearance of the weak shoulder at 2270 cm-1 in Fig.9 results from strong interaction between LiClO4 and PAN molecules. The supposition that the new shoulder at about 2270 cm-1 reflects the interaction between Li+ ions and the CN group also agrees with the analysis of the chemical structure of PAN. There are a pair of unbounded electrons on the nitrogen atom in the CN group and the Li+ ion has an empty orbital. Therefore it is possible for Li+ ions to bond with the nitrogen atom of CN group and an aggregate is formed. Because the intensity of the CN group is very strong, the split shoulder at 2270 cm-1 is too easy to be overlapped by the strong 2244 cm-1 band when the number of Li+ ions is not large enough. Therefore in the normally used concentrations of LiClO4 and PAN, no trace of strong interaction between PAN and LiClO4 could be observed by spectroscopy.

4. Competition between plasticizer and polymer on associating with Li + ions


A. Comparison of Li+-association in liquid and gel electrolytes with different plasticizers (solvents). Fig.10 shows the Raman spectra of the O=C-N deformation mode of DMF with PAN dissolved in the solution of DMF/LiClO4. Comparing Figs.2A and 10, it can be seen that the Raman spectrum of this mode in DMF/LiClO4/PAN system is very similar to that in DMF/LiClO4. At each comparable molar ratio, the spectra show no obvious changes before and after PAN is added to the solution. This indicates that the addition of PAN does not influence significantly the interaction between DMF and Li+ ions. Similar phenomena have also been observed in the system of DMSO/LiClO4/PAN systems. However, the spectral difference of the ring deformation mode of propylene carbonate (PC) is very obvious before and after PAN is dissolved in the solution of PC/LiClO4 (Fig.11, for the detail of the evolution process of the ring deformation mode of PC containing different contents of LiClO4, see Ref.14). Firstly, the addition of PAN into the solution makes the Li+-solvent (or plasticizer) interaction less observable. In the system of PC/LiClO 414, the 722-cm-1 (split from ring deformation mode of PC at 712 cm-1) component

becomes apparent at PC/LiClO4 = 10:1. However, this component is still undetectable at PC/LiClO 4 = 7:1 when the content of PAN is very high (i.e., from PC:LiClO4:PAN=7:1:3 to 7:1:5) in PC/LiClO 4 /PAN. Secondly, Fig.11 indicates that the splitting of the ring deformation mode is strongly dependent on the content of PAN in PC/LiClO4 solution. If the PAN content is very high, the Li+-solvent interaction becomes apparent only at very high PC:LiClO 4 ratios. In addition, the Li+-solvent interaction becomes apparent at a lower PC:LiClO4 ratio if the PAN content is very low. Similar to the system of EC/LiClO4/PAN, the interaction between Li+ ion and PAN can be seen clearly in the system PC/LiClO4/PAN (Fig.12). When the molar fraction of LiClO4 is much smaller than that of PAN, there is almost no difference in the shape of the nitrile stretching band before and after PAN is plasticized with PC. When the molar fraction of PAN in the electrolyte is further decreased, the shape of the 2244 cm-1 band begins to deviate from its originally symmetric shape. The Raman spectra of DMF/LiClO4/PAN electrolytes are also measured at molar ratios from 10:1:3 to 10:3:3 and 10:2:4. However, changes in the molar ratio of DMF:LiClO4:PAN over this range do not influence the shape or the vibrational frequency of the nitrile stretching band (at 2244 cm-1). This means that no strong interaction is detectable between the Li+ ions and the nitrile group. These results agree with those observed in the DMSO/LiClO4/PAN system. Summarizing the above results and discussion, it seems that the polymer host and the plasticizer interfere with each other in their association with the Li+ ions. On one hand, the addition of PAN into PC/LiClO 4 solution suppresses the Li+-PC interaction, and the observation of the Li+-PAN interaction is related to the PAN content in the polymer electrolyte as well as to the plasticizer. On the other hand, PAN only weakly interferes with the Li+-DMF interaction and no Li+-PAN interaction is observed over a wide range of molar ratio of DMF (or DMSO):LiClO4:PAN. A competition exists between the association of plasticizer and PAN with the Li + ions. This assumption is supported by the following experimental facts. B. Interaction between Li+ ion and PAN in plasticizer-removed electrolytes. The three-component polymer electrolyte (plasticizer, polymer, and salt) system is much more complicated than a two-component one (such as polyethylene oxide (PEO)-salt solid electrolyte). If the plasticizer is removed from the solid (gel) electrolyte, the plasticizer-free PAN-based electrolyte (PAN-salt) will be similar in some respects to the PEO-salt solid electrolyte. By studying the properties of such simple PAN-salt systems, some information can be obtained to understand the ion(s) transport mechanism in practical PAN-based gel electrolytes.

Gels composed of LiClO4, PAN and plasticizer (PC, for example) were cast on to a glass plate and heated at 120 to 140C for 6 hours to remove most of the plasticizer. Then the sample was transferred to a vacuum chamber (0.1 Pa) and stored there for a week to eliminate the residue of the plasticizer. Finally the plasticizer-removed PAN-LiClO4 film was peeled off from the glass plate and ready for the measurements. The Raman and IR spectroscopic measurements showed that there was no detectable trace of plasticizer in the samples. For brevity, the plasticizer-removed gel electrolytes and plasticizer/PAN are denoted as LiClO4/PAN and PAN, respectively) in the following discussion. Fig.13 shows the Raman (A) and IR (B) spectra of the CN stretching mode of PAN and LiClO4/PAN at various molar ratios. It can be seen from Fig.13A that the Li+-PAN interaction is evident at a LiClO4/PAN molar ratio as low as 1:7, whereas in Fig.13B the splitting of the CN stretching mode of PAN is still observable at a molar ratio of LiClO4:PAN = 1:10. The common spectral features of Fig.13 can be summarized as follows. Firstly, the relative intensity related to the Li+-perturbed CN stretching band is proportional to the LiClO4:PAN molar ratio over a large range. This is similar to the Li+-solvent interaction previously discussed. Secondly, the interaction between Li+ ion and PAN occurs and can be detected at very low LiClO4:PAN molar ratios. These characteristics have never been observed or reported in plasticizer/LiClO4/PAN systems before. Since different plasticizers were employed in synthesizing the gel electrolytes and similar results were obtained in LiClO4/PAN, it can be reasonably assumed that the Li+-PAN interaction observed in the LiClO4/PAN is independent of the plasticizer originally employed. Therefore, it can be concluded that the lack of Li+-PAN interaction in the plasticizer/LiClO4/PAN system at low LiClO4:PAN molar ratios is due to competition for the Li+ ions between the plasticizer and the PAN host in the gel electrolytes. According to the above results, the competitive capability of these three polar materials to form ion associates with the Li+ ions is: DMF>PC>PAN. Because the competition of PAN is much weaker than that of DMF, it is difficult to observe the Li+-PAN associates when PAN and DMF coexist. Most of the Li+ ions have been associated with DMF. However, in the PC/LiClO 4/PAN electrolyte, such Li+-PAN associates are observed much more easily. When the plasticizer is removed, no plasticizer would compete with the polymer host for Li+ ions. It is reasonable that the Li+-PAN interaction can be easily observed in this case even when the molar fraction of LiClO4 is very low (<1:10 in LiClO4:PAN).

5. The state of ClO4 ions in the plasticizer-removed polymer electrolytes.

The Raman spectra of the 1 mode of ClO 4 anion at different LiClO4/PAN molar ratios are shown in Fig.14. It indicates that two Lorentzian components, at 933 and 954 cm-1, respectively, are sufficient to deconvolute the Raman envelope of the 1 mode quite well when the molar fraction of LiClO4 is high. The 933 cm-1 band is from the unperturbed ClO 4 anions. Based on the assignment in Section 1B, it seems plausible to attribute the 954-component to the multiple ion aggregate in LiClO4/PAN (i.e.,{Li+ClO 4}n). However, such an attribution contradicts the observation of the ion aggregates in LiClO4-solutions. Usually, the multiple ion aggregates follow the appearance of the solvent-separated ion pairs and the contact ion pairs in a liquid perchlorate salt solution. An explanation is that the perturbation of PAN on the ClO 4 anion is even weaker than that of the ordinary solvents of LiClO4 and the 954 cm-1 component comes from recrystallized LiClO4. This may also be the reason why PAN is not as competitive as the plasticizer (for example, PC and DMF) on associating with the Li+ ions. The activated IR spectra of the 1 mode agree well with the Raman spectra (not shown).

6. Coordination number of Li+ ion in plasticizer-removed PAN based electrolytes


Fig.15 shows the XPS spectra of N 1s of the samples. In pure PAN, N 1s spectrum has a single peak at ca. 397.6 eV, corresponding to the single nitrogen atom of the nitrile group in each repeat unit of PAN. This band moves to ca. 398.4 eV when LiClO4 is added into the polymer. It has been verified repeatedly that in the PAN-based electrolytes, PAN-plasticizer-LiClO4, only the polar groups of the plasticizer or the Li+ ions can interact directly with the nitrile groups11,13,14. Therefore, in the present plasticizer-removed PAN-LiClO4 system, it is reasonable to attribute the chemical shift of this band to the association between the Li+ ion and the N atom on the frame of PAN chain. When the Li+ ion associates with the N atoms, the density of the electronic nebula around the N atom becomes lower and the strength of the nitrile bonding becomes weaker. Hence, the electronegativity of the N atom becomes larger. As a result, the binding energy of the nitrile becomes higher. Moreover, it is also worth noticing that the binding energies of N 1s in PAN-3 (LiClO4:PAN=1:3) and PAN-4 (LiClO4:PAN=1:4) are quite similar, though the XPS spectra of N 1s are significantly different before and after adding LiClO4 into PAN. It seems that the spectral changes of the N 1s has reached its saturation at the molar ratio of PAN:LiClO4=4:1 or larger than 4:1. XPS spectra of Li 1s, Cl 2p and O 1s agree with the results from N 1s spectra. These results combine to suggest that the coordination number of Li+ ion in PAN should not be larger than four. Therefore, a coordination number of four is suggested for LiClO4 in plasticizer-removed polymer electrolytes.

7. 7Li NMR Spectra of the practically used PAN-based electrolyte


A spectrum of PAN-EC-PC-DMSO-LiClO4 at room temperature is shown in Fig.16. It can be seen that the spectrum is composed of three components: a narrow and strong Gaussian peak, a broad Lorentzian peak and a small peak with a chemical shift. Chemical shift represents the change of chemical environment around the subject ion. Usually the linewidth of an NMR peak reflects the ion mobility in the system. The faster the ions move, the narrower its related peak will be. Fig.16 indicates that there are three kinds of Li+ ions corresponding to the three components in the NMR spectrum. The narrowest one is caused by the fast-moving ions in the gel state; the broad peak corresponds to the slow-moving ions in solid PAN and the chemical shifts comes from the ions associated with the plasticizer. Because gel is the major phase in PAN-based electrolytes, fast ion transport in the gel state is the major contribution to the high conductivity of the PAN-based electrolytes. The temperature-dependent linewidth of the 7Li spectra shows a motion narrowing temperature at ca.200 K (Fig.17). Below 200 K the spectra are very broad, indicating a slow mobility of most of the Li+ ions. As the temperature increases, a narrow peak appears and becomes the dominant component around room temperature. Using HB equation, log(1/W) = Ea /(2.303RT) + log(1/B). an activation energy of 0.28 eV is obtained at ca. 25C, slightly smaller than the value from the VTF equation. A plausible explanation is that triplet ion aggregates may also contribute to the ionic conductivities as well as the free ions. In a HB equation, the B value represents the extent to which the dipole-induced line broadening can be eliminated at infinitely high temperatures. If the process is a long-range one, the dipolar interactions are completely averaged out and the linewidth would be narrowed to its lifetime values (B<10-12 KHz). Otherwise, the B values related to the dipolar interaction becomes very large (B>10-4 KHz) if the motion is short-ranged. The values measured from our sample at 25C are about 10-7 KHz. It can be concluded that the Li+ ion motion in the electrolyte is mainly a long-range process.

8. Ion Transport Mechanisms in PAN Based Electrolytes


The conductivity of an electrolyte is determined by the concentration of the charge carriers and their mobilities. The extent to which ion-pairs are dissociated to form charge carrying ions is determined by the dielectric constant of the plasticized polymer with a plasticizer. Based on the above knowledge about the microstructure of a PAN-based electrolyte, it is possible to depict the total structural geometry in PAN-based electrolytes. When the concentration of LiClO4 is very low

in a LiClO4-solvent (plasticizer) system, most of the LiClO4 is dissociated. The ionic conductivity of the solution increases with increasing LiClO4 content (Scheme 1a). However, with increasing LiClO4 concentration, ion pairs appear and the association becomes dominant, leading to a decrease of concentration of free ionic carriers and therefore, a decrease of the ionic conductivity (Scheme 1b through d). The number of the EC molecules around one Li+ ion can be as large as six in the EC/LiClO4 solution. In a (plasticizer-removed) PAN/LiClO4 system, the Li+ ions are associated with the PAN chain by way of the comb-like CN group. One Li+ ion can associate with as many as four CN groups. Such association is shown in Scheme 2a. ClO4 anion, due to its very weak interaction with PAN, is separated from the Li+ ions and exists as a lone ionic carrier. When the content of LiClO4 is further increased, the surplus Li+ ions (in respect to the coordination number of four) recrystallized with ClO4 anions (Scheme 2b). As the Li+ ions are held in the polymer matrix by means of the coordination bond with the nitrogen and the chain of the polymer is rigid at room temperature (without plasticizer, the glass transition temperature of PAN is as high as 169C), the ionic conductivity of PAN/LiClO4 system would be very low and the contribution of the anion to the total ionic conductivity is dominant, i.e. the transference number t>>t +). When LiClO4 and PAN are dissolved in a plasticizer the coordination bonds of Li+ ion to PAN can be decoupled due to the strong competitive capability of the plasticizer for the Li+ ions. In a PC- or ECplasticized polymer electrolyte, the decoupling behavior becomes obvious with increasing molar ratio of the plasticizer as shown in Scheme 3. The decoupling behavior becomes more obvious due to the stronger competitive capability of the plasticizer in DMF- or DMSO-plasticized polymer electrolyte than that of PC or EC (Scheme 3d and e). In such an electrolyte, no Li+-coordinated PAN molecules can be detected by the Raman or IR spectroscopy as long as the LiClO4/PAN mixture can be dissolved by the plasticizer. Therefore, the Li +-coordinated PAN molecules must be very scarce, if exist, in such a system. Besides the decoupling effect of the plasticizer to the Li+-PAN coordinate species, repulsions also exist between the polymer and the plasticizer through their dipoles, which has been shown in Scheme 3 as well. It has been verified repeatedly that the amorphous region in the PEO-salt electrolytes is responsible for the high mobility of the Li+ ions. XRD patterns by Huang et al 5 demonstrated that this is also the case in the PAN-based electrolytes. 7Li NMR studies on the PAN-based electrolytes indicates that there are at least three states of Li+ ions in the plasticized PAN-based electrolytes: i. Li+ ions in the gel state, which constitutes the most important part of the polymer electrolytes; ii. Li+ ions moving along the PAN chain; and iii. Li+ ions associated with the plasticizer molecules. Moreover, analysis of the temperature-dependent linewidth of the 7Li

NMR spectra demonstrates that the movement of most of the Li+ ions is between a long-range and a shortrange one. As to the long-range movement, it is quite possible that the Li+ ions will move along the side chain of PAN, jumping from one site to the next. According to the Raman, IR and XPS spectra this site is most probably provided by the CN group of PAN. Based on polymer physics, the length of a segmental chain of PAN is a dozen of PAN basic units (-CH2-CH-CN-), a few nanometers long. As the polymer chain moves in the form of segmental chain movement, it is easy to understand that the Li+ ion movement with the polymer is a long-range one in the solid state. If the Li+ ions move in the liquid medium, its movement will be strongly affected by the Brown movement of the solvent molecules and will be a short-range one. NMR spectrum (Fig.16) has shown that the dominant part of the PAN-based electrolytes is the one with a gel structure (corresponding to the strongest Lorentzian peak). Gel is a state between solid and liquid, the movement of the Li+ ions in the gel is between a short-range one and a long-range one. As a result, the Li+ ions moving in the gel contribute the most to the ionic conductivity of the polymer electrolyte. There are also Li+ ions moving along the segmental chain of PAN. The transport mechanism of these ions are quite similar to those in the PEO-salt electrolyte. By coordinating with the side chain of the polymer, the movement of the Li+ ions is quite similar to that of the segmental chain in a PEO-salt electrolyte. Therefore, their movement is a long-range one. However the Li+ ion movement along the segmental chain cannot be the major contribution to the ionic conductivity in a PAN-based electrolyte at room temperature because the movement of the segmental chain of PAN is slow at room temperature in respect to the ions in the gel. This may be quite different from the PEO-salt electrolyte, where the segmental chain movement is the most important to the conductivity. Of course, when the temperature is elevated, the movement of the segmental chain becomes faster. So the total ionic conductivity of the polymer electrolyte tends to be similar to that of the liquid electrolyte. This hypothesis agrees with the experimental results of Olsen and Koksbang15. We also notice that the transference number of the anion16 ( 0.64) is usually larger than that of the cation ( 0.36) in the PAN-based electrolyte at room temperature. The movement of ClO4 anion in the gel is similar to that in a liquid medium; instead of moving along the segmental chain by jumping from one site to the next, these anions move outside the tunnel provided by the PAN-chain due to the weak interaction between the anion and the polymer. On one hand, the weak solvation effect with the plasticizer allows the anions to move faster than the cations. On the other hand, the large ionic size of the anion blocks its fast movement more severely in the gel than in the liquid due to the much larger viscosity of the gel than of the

liquid. These two effects coexist and combine to make the movement of the anions move slower in a plasticized polymer electrolyte than in a liquid electrolyte of the same solvent. Considering the above transport mechanisms of the ions (anion and cation), the functions of the plasticizer in the PAN-based electrolytes may be summarized as the following: i. decreasing the glass transition temperature of the polymer and dissolving the crystallites of the polymer. This will lead to the increase of the mobility of the segmental chain of the polymer and help the transport of the charge carriers depending on the movement of the segmental chain (ions moving in the solid state polymer); ii. dissolving the electrolytic salt so as to provide charge carriers for the polymer electrolyte; iii. interacting with the polymer molecules by way of dipolar interactions and therefore, increasing the polarities of the polymer and the plasticizer itself which, in turn, will help the dissociation of the electrolyte in the polymer electrolyte; iv. destructing the coordination bonds of the Li+ ions with PAN and making more Li+ ions move in the gel instead in the solid state. This will help to increase the mobility of the Li+ ions in the polymer electrolytes. Based on the above functions of the plasticizer in a polymer electrolyte, some suggestions may be proposed for selecting the plasticizer of the polymer electrolyte: i. compatible with the polymer and the electrodes; ii. thermodynamically stable; iii. low viscosity; iv. high dielectric constant; v. low melting point and high boiling point; vi. inexotic and easily obtainable. The above six points are conventional requirements to the plasticizer for polymer electrolytes. However, according to our research results, one more requirement must be added to the plasticizer: vii. the plasticizer should have a strong competitive capability to decouple the coordination bond of PAN-Li+, but this capability should not be too strong in order to avoid any strong Li+-plasticizer-ClO4 coordination. This requirement is different from the above six requirements. Dielectric constant of a plasticizer has no direct relationship with its competitive capability. For example, Li+-PAN associates are easier to be formed in EC (or PC)/PAN/LiClO4 system than in DMF or DMSO/PAN/LiClO4 system though PC and EC

have very high dielectric constant compared with DMF or DMSO. On the other hand, although DMF has a stronger competitive ability to decouple the Li+-PAN coordination bond, it has a lower dissolving capability for LiClO4 and Li+-DMF associates are much easier to be formed in DMF/PAN/LiClO4 than in EC (or PC)/PAN/LiClO 4 system. Usually a single-component plasticizer cannot satisfactorily meet all the above requirements. Therefore, a multi-component plasticizer is necessary to prepare polymer electrolytes with high ionic conductivity, electrochemical capability with the electrode, thermodynamic stability and enough mechanical strength, etc..

Conclusions
By analyzing the Raman, IR, fluorescence, X-ray photoelectron and the 7Li NMR spectra of the components of PAN based electrolytes, we can conclude: I. Strong interactions between plasticizer and the Li+ ions mainly occur on the O atoms of its C=O or S=O bonds. The increase of the LiClO4 concentration not only affects the structure of the plasticizer but also perturbing the symmetry of the ClO4 anion itself, forming ion associates, such as solvent separated ion pairs, contact ion pairs and even multiple ion aggregates. This will result in a decrease of the concentration of the free charge carriers. A coordination number of 6 is proposed for the Li+ ions in EC/LiClO4 solutions. The quantitative relationship between the solution concentration and the concentration of free ions is roughly discussed. II. Evidence of strong interactions between plasticizers and PAN is experimentally found. Interactions of PAN with different plasticizers show different characteristics. The interaction between EC and PAN mainly occurs by way of dipolar repulsions, while the self-associated DMSO linear dimers must be first decoupled before they can interact with PAN as monomers in a DMSO/PAN system. III. For the first time, interactions between the Li+ ions and the PAN were evidently observed in polymer electrolytes with EC and PC as plasticizers. These interactions occur through the Li+ ion and the nitrogen atom of the CN bonds in PAN molecules. However, similar interactions failed to be observed in polymer electrolytes plasticized with DMF or DMSO. This is ascribed to the strong competition between the plasticizer and the polymer on associating with the Li+ ions, evidenced by comparing the Raman and IR spectra of the polymer electrolytes before and after the plasticizer is removed. XPS studies indicate that the coordination number of the Li+ ion in plasticizer-removed PAN-based electrolyte is four.

IV. Combining the results from the Raman and IR spectra with the 7Li NMR results, three existence states are identified for the Li+ ions in plasticized PAN-based electrolytes: the narrow Gaussian peak is ascribed to the fast moving Li+ ions in the gel state, the most important phase for the ionic conductivity of the electrolytes. The broad peak is attributed to the slow moving Li+ ions in the solid phase of PAN, which contributes little to the ionic conductivity of the polymer electrolyte due to their slow mobility. Finally, the chemical shift observed is assigned to the association between the Li+ ion and the plasticizer, a popular phenomenon observed in the plasticized polymer electrolytes. V. Considering the strong competition effects between the plasticizer and the polymer and the differences of the charge carriers moving in the gel and the solid phases, the function of the plasticizer for the mobility of the ions in the electrolytes is emphasized. It is suggested that on choosing a plasticizer for the PAN based electrolytes, an important factor that should be considered is the competitive capability of the plasticizer with the polymer on associating with the Li+ ions. In order that, more Li+ ions can move in the gel state instead of moving in the solid phase of PAN or as plasticizer-separated ion pairs, which will drop the ionic conductivity.

Acknowledgments
This work was supported by the National Science Foundation of China (Contract No.59672027).

References
1. P. V. Wright, Brit. Polym. J. 7(1975) 319. 2. P. V. Wright, J. Polym. Sci. Polym. Phys. Ed. 14(1976) 955. 3. M. Armand, J. M. Chabagno, and M. J. Duclot, 2nd International Conference on Solid Electrolytes, St Andrews paper 6.5, 1978. 4. K. M. Abraham and M. Alamgir, J. Electrochem. Soc., 137(1990)1657 5. Biying Huang, Thesis for Ph.D, Institute of Physics, Chinese Academy of Sciences (1995). 6. G.Pistoria, Bull. Electrochem., 7(1991)524. 7. Z. Deng and D. E. Irish, J. Chem. Soc. Faraday Trans. 88 (1992) 2891. 8. Z. Deng and D. E. Irish, Can. J. Chem. 69 (1991) 1766. 9. S. Hyodo and K. Okabayashi, Electrochim. Acta, 34(1989) 1551

10. U.Olsher, R.M.Izatt, J.S.Bradshaw and N.K.Dalley, Chem.Rev., 91(1991)137. 11. J.E.Enderby, Chem.Soc.Rev., 24(1995)159. 12. S.Schantz, J.Chem.Phys., 94(1991)6296. 13. J.Stevens and P.Jacobbson, Can.J.Chem., 69(1991)1980. 14. D.Battisti, G.A.Nazri, B.Klassen, and R.Arona, J.Phys.Chem., 97(1993)5826. 15. Ib.I.Olsen and R.Koksbang, J.Electrochem.Soc., 143 (1996)570. 16. M.Ue and S.Mori, J.Electrochem. Soc., 142 (1995)2577.

Captions to the figures and schemes in the text


Fig.1 Raman spectra of the ring-bending mode of LiClO4/EC at different molar ratios of LiClO4:EC: a. 0 (pure EC); b. 0.008; c. 0.04; d. 0.09; e. 0.21 and f. 0.35. Fig.2A The evolution of the C=ON deformation mode of DMF (at 658 cm-1) and the v4 mode of the ClO 4 anion (at 624 cm-1) in the Raman spectrum with different molar ratios of LiClO4:DMF: a. 0 (pure DMF); b. 0.04; c. 0.10; d. 0.18; and e. 0.30. Fig.2B The IR spectra of the O=CN deformation mode of DMF and the v4 mode of the ClO 4 anion at different molar ratios of LiClO4:DMF: a. 0 (pure DMF); b. 0.04; c. 0.10; d. 0.18; and e. 0.30 LiClO4 (legend: solid line, the experimental data; circle, the profile of the band fitting; short dashed line, the band-fitted component). Fig.3 The Raman spectra of the C-H symmetric stretching (at 2912 cm-1) and asymmetric stretching (at 2996 cm-1) modes of DMSO at various molar ratios of LiClO4:DMSO: a. pure DMSO; b. 0.05; c. 0.1; d. 0.2; e. 0.3; f. 0.4 and g. 0.6.

Fig.4 The IR spectral evolution of the 1 mode (ca. 930 cm-1) of ClO4 with different LiClO4:EC molar ratios in LiClO4/EC solution: a. 0.008; b. 0.04; c. 0.09; d. 0.21 and e. 0.35. Fig.5 The relative intensity of the ring-bending mode.of free EC (If) vs that of the total (free and bound) EC (It ) in LiClO4/EC solutions. Fig.6 The relationship between the concentrations of free and bound ClO4- anions and the LiClO4/EC solution concentration. Fig.7 The Raman spectra of the C=O stretching mode of EC at 1762 cm-1 with different EC:PAN molar ratios: a. pure EC; b. 4:1; and c. 3:2. Fig.8 The molar ratio dependence of the relative intensity of the nitrile stretching mode to different modes of DMSO. Fig.9 The Raman spectra of EC/PAN/LiClO4 at different mass ratio of EC:PAN:LiClO4: a. 7:0:3; b. pure PAN heat-treated at 110C in air for 10 h; c. 3:2:0; and d. 49:25:26. Fig.10 The Raman spectra of the O=C-N deformation mode of DMF (at 658 cm-1 in pure DMF) and the 4 mode of ClO 4 anion (at 624 cm-1) in DMF/LiClO4/PAN electrolytes with different molar ratios of DMF:LiClO4:PAN (the band with an arrow overhead is from the ClO4 anion) Fig.11 Lorentzian band fitting to the Raman spectra of the ring deformation mode (at 712cm-1 ) of the LiClO4perturbed PC in PC/LiClO 4/PAN electrolyte at various molar ratios of PC:LiClO 4:PAN (note: the band with an arrow overhead is from the 4 mode of ClO4). Fig.12 The Raman spectra in the CN stretching region of PAN in PC/LiClO 4/PAN at various molar ratios of PC:LiClO4:PAN. Fig.13 The Raman (A) and the IR (B) spectra of the CN stretching mode of PAN in plasticizer-free LiClO4/PAN with various LiClO4:PAN molar ratios. Fig.14 The Raman spectra of the 1 mode of the ClO4- at different molar ratios of LiClO4:PAN in plasticizerremoved solid electrolyte. Fig.15 The XPS spectra of N 1s in pure PAN and PAN containing different contents of LiClO4 in plasticizerfree solid electrolyte. Fig.16 A typical 7Li NMR spectrum of PAN-EC-PC-DMSO-LiClO4 at room temperature (Frequency: 34.964 MHz). Fig.17 The evolution process of the line width with the change of temperature of PAN-EC-PC-DMSOLiClO 4 electrolyte.

Scheme 1 An evolution process of the ionic species in a LiClO4-solvent solution. a. at very low concentration, the salt is completely dissolved and there are only solvated Li+ ions; b. when the concentration is raised solvent-separated ion pairs appear in the solution; c. further increase of slt concentration leads to ion pairs and the ionic atmosphere is damaged; d. in concentrated solution, ionic aggregates appear and dominate the solution while the concentration of the free Li+ ions drops dramatically. Scheme 2 Association of the ionic species in a plasticizer-removed LiClO4/PAN system. a. the Li + ions are associated with the CN groups of the PAN molecules. The coordination number could be as large as 4. The ClO4 ions move outside the PAN matrix; b. when the salt content is further increased, the extra ions will turn to form ion pairs, [Li+ClO 4]. Scheme 3 Association of ionic species in a plasticized LiClO 4/PAN system. a. when a plasticizer is added to a LiClO4/PAN system, the Li-PAN associates and the ion pairs are decoupled due to the solvation and strong competition effects of the plasticizer for the Li+ ions (ref. Scheme 2a); a repulsion exists between the polymer and the plasticizer; b and c. When more plasticizer is added, more and more ion associates are decoupled and the content of the solvent (plasticizer)-separated ion pairs is low; a gel polymer electrolyte with high ionic conductivity is obtained. d. if a plsticizer with very strong competitive capability is added to the LiClO4/PAN system, the Li+PAN associates are deccoupled and solvent-separated ion pairs tends to be formed. The appearance of the ion pairs will reduce the concentration of free charge carriers.

Table 1. Relative integrated intensity of the ring-bending mode of EC at 715 cm-1 (If) and 725 cm-1 (Ib) and their sum (It ) at different concentrations of LiClO4 with the corresponding I1230 as the standard. Cont. (%) peak (cm-1) 1 5 10 20 30 37.6 45.4 51.7 716.5 716.4 716.4 716.1 715.7 717.9 718.9 -If 1.6873 1.3996 1.1542 0.5839 0.4764 0.3277 0.2325 0 peak (cm-1) 726.0 727.5 727.2 727.5 727.4 728.5 728.5 729.8 Ib 0.0623 0.3063 0.5215 0.9445 1.0164 1.1384 1.2040 1.3824 It 1.7496 1.7059 1.6757 1.5284 1.4927 1.4662 1.4365 1.3824

You might also like