You are on page 1of 32

www.aspbs.

com/enn
Encyclopedia of
Nanoscience and
Nanotechnology
Polymer Electrolyte Nanocomposites
Mikrajuddin Abdullah
1
, Wuled Lenggoro, Kikuo Okuyama
Hiroshima University, Hiroshima, Japan
CONTENTS
1. Introduction
2. Conductivity Enhancement
in Polymer Electrolytes
3. Development of Polymer
Electrolyte Nanocomposites
4. Preparation Methods
5. Important Parameters
6. Charge Transport Characterizations
7. Spectroscopic Characterizations
8. Microscopic Analysis
9. Thermal Characterizations
10. Density Method
11. Electrical Properties
12. Mechanical Properties
13. Thermal Properties
14. Luminescent Composites
15. Conclusion
Glossary
References
1. INTRODUCTION
Rechargeable cells are key components in mobile technolo-
gies, such as portable consumer electronics and electric vehi-
cles [1]. A search for batteries that provide high energy
density and multiple rechargeability has been a subject of
considerable attentions. Even though battery technology
developed one hundred years ago, progress and improve-
ments in technology have been slow, particularly when com-
pared to the growth of computer technology [2].
A Li-based battery provides a high density and exibility
of design. Todays lithium battery has a high specic energy
(>130 W h kg
1
), a high energy density (>300 W h L
1
),
1
Permanent address: Department of Physics, Bandung Institute of
Technology, Jalan Ganeca 10 Bandung 40132, Indonesia.
high cell voltage (3.5 V), as well as a long cycle life
(5001000) charge/discharge. Worldwide production of such
devices exceeded 200 million in 1997 and it will be approx-
imately three times that number during 2001 [1]. Since
lithium produces an explosion reaction with water-based
electrolytes, a search for nonaqueous electrolytes is critically
important to the production of the next-generation lithium
battery, using electrolytes in an solid phase in an effort to
develop more environmentally friendly materials. Polymer
electrolytes are potential candidates for replacing the con-
ventional aqueous electrolytes in lithium batteries. Polymers
containing esters, ethers, or mixtures thereof which have the
ability to dissolve salts are the base materials for polymer
electrolytes. Polymer electrolytes are generally prepared by
mixing high molecular weight polymers (HMWPs) with a
salt solution. The polymer serves as solid solvent, thus per-
mitting the salt to dissociate into anions and cations. Since
the mass of a cation is much smaller than that of an anion,
the electrical conductivity is dominated by cation transfer.
Lithium salts are usually used for this purpose since they
are the most electropositive of materials (3.04 V relative
to the standard hydrogen electrodes) as well as the lightest
metal (atomic mass 6.94 g/mol, and density 0.53 g/cm
3
) and
thus facilitate the design of storage systems with high energy
density (Watt hour/kg) [1]. Table 1 shows a comparison of
the electrochemical properties of several metals.
Until presently, however, no polymer electrolyte-based
lithium batteries are commercially available in the market.
Therefore, worldwide research is being focused on the devel-
opment of high power and high energy density polymer elec-
trolytes with a major attention to safety, performance, and
reliability. A battery contains two electrodes: positive and
negative (both sources of chemical reactions), separated by
an electrolyte that contains dissociated salts through which
ion carriers ow. Once these electrodes are connected to
external circuits, chemical reaction appears at both elec-
trodes to result in a deliverance of electrons to the external
circuits. The properties of a battery thus strongly depend on
the electrolyte, anode, and cathode. With the use of poly-
mer electrolytes in lithium batteries, high specic energy and
specic power, safe operation, exibility in packaging, and
low cost in fabrication as well as low internal voltage drop
at relatively large current withdraw is expected [3].
ISBN: 1-58883-064-0/$35.00
Copyright 2004 by American Scientic Publishers
All rights of reproduction in any form reserved.
Encyclopedia of Nanoscience and Nanotechnology
Edited by H. S. Nalwa
Volume 8: Pages (731762)
732 Polymer Electrolyte Nanocomposites
Table 1. Electrochemical properties of several metals that have poten-
tial applications for use in batteries.
Atomic weight Valence Specic charge Electrode
Metal [g/mol] charge [A h kg
1
] potential [V]
Li 6.94 1 3862 3.05
Na 22.99 1 1166 2.71
Mg 24.31 2 2205 2.38
Zn 65.38 2 820 0.76
Cd 112.41 2 477 0.40
Pd 207.20 2 250 0.13
Realization of commercial polymer electrolyte batter-
ies is actively investigated in many companies worldwide.
A major effort to develop advanced polymer batteries for
electric vehicles began in the early 1990s by 3M and Hydro-
Quebec [4]. The battery contains a lithium metal anode, a
polyethylene oxide (PEO)-based polymer electrolyte, and a
vanadium oxide (VO
x
) cathode. The reversibility of lithium
intercalation and deintercalation in the VO
x
is quite good
but the average discharge of the cell is low. PolyPlus Bat-
tery company in the United States is developing poly-
mer electrolyte-based lithium battery which would operate
at room temperature with specic energy as high as 500
W h kg
1
[3]. In a prototype cell, using cathode made of
lithium intercalated disulde polymer, a specic energy as
high as 100 W h kg
1
and charge and discharge cycles almost
reproducible for over 350 cycles were observed at 90
o
C [5].
Moltec company reported a specic density of 180 W h kg
1
for an AA-sized battery based on organosulfur cathode [6].
Ultrane Battery company reported a room-temperature
solid polymer battery based on intercalation type electrode
with a specic energy 125 W h kg
1
and charge/discharge
cycling time of 500 [3]. This performance is still below the
consumer expectation threshold. In 1995, Turrentine and
Kurani in the United States did a survey on demand for
alternative fuel cell for vehicles and found that consumers
agreed to buy electric vehicles which would run for at least
200 km per battery [7].
2. CONDUCTIVITY ENHANCEMENT
IN POLYMER ELECTROLYTES
It is believed that in the polymer electrolytes, the cations
are coiled by polymer segment leaving the anions to occupy
separate positions [8]. Battery performance is limited by the
speed of cation diffusion. The transport of cations takes
place if there is a relaxation of the polymer segments so
that cations are released from a segment and then occupy
another segment. Segmental relaxation requires the pres-
ence of free volume in the polymer matrix, a condition that
can be attained if the polymer is in an amorphous state.
Unfortunately, most HMWPs crystallize at ambient temper-
atures. Ions are transported with difculty in a crystalline
matrix since no chain relaxation occurs and, as a result, the
conductivity of polymer electrolytes in this phase (at ambi-
ent temperature) is depressed. The transport of ions in this
state is dominated by the jumping of cations to the nearest
location, which depends on the blocking potential (activa-
tion energy). This is similar to the jump of charge carriers in
crystalline solids. The characteristic time for jumping is pro-
portional to the exponential of the blocking potential. This
results in a conductivity of the order of 10
8
S/cm, a value
that is far below the desired value of about 10
4
S/cm [9].
When it enters the amorphous state, that is, at temperatures
above the melting point, a high conductivity appears. For
a commonly used polymer, that is, polyethylene oxide, the
melting temperature is 65
o
C. This is, of course, impractical
since the operating temperature for most electronic devices
is room temperature. In addition, at temperatures above the
melting point, the polymer becomes soft, causing the solid-
state properties to degrade. Initiated by the work of Wright
and Armand [1012], several kinds of polymer electrolytes
have been intensively investigated around the world. Table 2
displays examples of polymer electrolytes and their mea-
sured conductivities at 20
o
C [13].
Improvements in the electrical conductivity of poly-
mer electrolytes at ambient temperature is therefore of
critical importance for technological applications. Several
approaches have been explored to realize this aim. Some
frequently used methods will be explained briey here.
2.1. Preparing Low Degree
of Crystallinity Polymers
By considering that the presence of amorphous state is
strictly important for improving the conductivity, the main
strategy is to enhance the amorphous state at low tem-
peratures. The rst approach is to prepare low degree of
crystallinity polymer from initial. It includes cross-linking
of two polymers [13, 14], synthesis of new polymer, cross-
linking high molecular weight polymer through -irradiation
[15, 16], addition of plasticizers in polymer electrolytes, addi-
tion of llers, and bending of two polymers [17, 18]. Another
strategy is to prepare an amorphous polymer so as to obtain
a polymer that is composed of four to ve monomeric units.
For this system, the chains must be sufciently long to effec-
tively complex cations but too short to crystallize at low tem-
peratures. Thus the matrix would still be in the amorphous
state even at low temperatures. The polymer host serves as
a solvent and does not include any organic liquids.
2.2. Addition of Side Chains
An alternative way to decrease the crystallinity of polymer
matrix is to introduce side chain to the polymer main chain.
Theoretically, chain ends and branch can be thought of as
impurities, which depress the melting point of the polymer.
Simple mathematical formulation can be used to explain the
melting point lowering by the presence of chain ends and
branch. If A
i
is the mole fraction of impurities (chain ends,
side chains, and branch), then the melting point of polymer,
T
n
, decreases according to [19]
1
T
n

1
T
t
n
=
F
AH
u
ln(1 A
i
) (1)
where T
t
n
= melting point of polymer containing only poly-
mer chain with innite chain length, F = gas constant,
and AH
u
= enthalpy of fusion per mole of repeat unit.
Chung and Sohn showed that the XRD intensity of polymer
Polymer Electrolyte Nanocomposites 733
Table 2. Examples of polymer electrolytes with their corresponding electrical conductivities at 20
o
C.
Example polymer Conductivity
Polymer host Repeat unit electrolyte (S/cm) at 20
o
C
Poly(ethylene oxide), PEO (PEO)
8
:LiClO
4
10
8
CH
2
CH
2
O
n

Poly(oxymethylene), POM POM:LiClO
4
10
8
CH
2
O
n

Poly(propylene oxide), PPO (PPO)
8
LiClO
4
10
8
(CH
3
)CH
2
CH
2
O
n

Poly(oxymethylene- (POO)
25
:LiCF
3
SO
3
310
5
oligo-ethylene), POO
(CH
2
O)(CH
2
CH
2
O)
n

Poly(dimethyl siloxane), DMS DMS:LiClO
4
10
4
(CH
3
)
2
SiO
n

Unsaturated ethylene UP:LiClO
4
10
5
Oxide segmented, UP (EO:Li

= 32 : 1)
HC=CH(CH
2
)
4
O(CH
2
CH
2
O)
n
(CH
2
)
4 x

Poly[(2-methoxy)ethyl (PMEGE)
8
:LiClO
4
10
5
glycidyl ether], PMEGE
CH
2
CHO
CH
2
(OCH
2
CH
2
)
2
OCH
3
n

Poly[(methoxy) poly(ethylene PMG


22
:LiCF
3
SO
3
310
5
glycol)] methacrylate, PMG
n
(EO:Ll

= 18:1)
CH
2
C
(CH
2
CH
2
O)
x
CH
3
C
O O
CH
3
n

(PEO-PPO-PEO)-SC (PEO-PPO-PEO)-SC: 1310


5
SC = siloxane crosslinked LiClO
4
(4:1 molar)
CH
3
CH
3
CH
3
CH
3
O O

PEO(CH
2
)
3
SiOSi(CH
2
)
3
PEO

PEO(CH
2
)
3
SiOSi(CH
2
)
3
PEO
PEO grafted polysiloxane, PGPS PGPS:LiClO
4
10
4
SiO
CH
2
CH
2
PEO
CH
3
n

Poly[bis-2-(2-methoxyethoxy) (MEEP)
4
:LiBF
4
210
5
ethoxy]phosphazene,MEEP (MEEP)
4
:LiN(CF
3
SO
2
)
4
510
5
(MEEP)
4
:LiC(CF
3
SO
2
)
4
10
4 P=N
OCH
2
CH
2
OCH
2
CH
2
OCH
3
OCH
2
CH
2
OCH
2
CH
2
OCH
3
n

decreases with the increase in the length of chain of comb-


shaped polymer [20].
Despite depressing the melting point of polymer, the pres-
ence of side chain also promotes the solvating of a salt
as reported by Ikeda and co-workers [21, 22]. The side
chain has shorter relaxation time compared to the main
chain. The coupling of the side chain with the ion carrier,
therefore, results in an increase in the conductivity. Watan-
abe et al. designed comb-shaped polyether host with short
polyether side chain [23]. However, the mechanical prop-
erties decreased even as the conductivity increased. High
conductivity with good mechanical properties was obtained
by designing a polymer of high molecular weight with tri-
oxyethylene side chain as reported also by Ikeda et al. [24].
With 18 mol.% of side chain, the conductivity was measured
to be 1.5 10
4
S/cm at 40
o
C and raised to 1.4 10
3
S/cm
at 80
o
C.
Composite of polymer with room-temperature molten slat
is also an interesting approach to improve the conductivity of
polymer electrolytes. Watanabe et al. reported the composite
consisting of chloroaluminate molten salt that possesses a
conductivity of 2 10
3
S/cm at 303 K [25, 26]. However,
the disadvantage of chloroaluminate is its hygroscopic prop-
erties such that it is impractical in application. The use of
non-chloroaluminate molten salt, therefore, is required to
avoid the hygroscopic problem. Tsuda et al. reported a con-
ductivity of 2.3 10
2
S/cm in composite of polymer and
room-temperature molten uorohydrogenates [27].
2.3. Addition of Plasticizers
Another approach to improve the conductivity is by addi-
tion of additional material into the host polymer. This
approach appears to be the simplest since a pre-produced
polymer can be used to make the polymer electrolytes. Pre-
viously, low molecular weight polymers were usually used to
reduce the operation temperature of polymer electrolytes.
The low molecular weight polymers which were added to
the matrix of HMWP to reduce the crystallinity at low
temperatures are frequently known as liquid plasticizers.
734 Polymer Electrolyte Nanocomposites
Feullade and Perche demonstrated the idea of plasticizing
the polymer with an aprotic solution containing alkali metal
salt in which the organic solution of the alkali metal salt
remained trapped within the matrix of solid polymer matrix
[28]. Such mixing results in formation of gels with ionic con-
ductivity close to the liquid electrolytes. Less evaporating
solvents such as ethylene carbonate (EC), propylene car-
bonate (PC), dimethyl formamide (DMF), diethyl phthalate
(DEP), diethyl carbonate (DEC), methyl ethyl carbonate
(MEC), dimethyl carbonate (DMC), -butyrolactone (BL),
glycol sulde (GS), and alkyl phthalates have been com-
monly investigated as plasticizers for the gel electrolytes.
Figure 1 shows the effect of plasticizer content tetraglyme
(tetraethylene glycol dimethyl ether) on the glass temper-
ature of a system of PEO-co-PPO (3:1):LiClO
4
[29]. The
decrease in the glass temperature can be simply explained
using a Fox equation:
1
T
g
=
W
1
T
g1

W
2
T
g2
(2)
where W
1
and W
2
denote the weight fractions of compo-
nent 1 and component 2, respectively, and T
g1
and T
g2
are
their corresponding glass transitions. This equation tells that
the glass temperature of the composite locates between the
glass temperature of the components. This relation is also
applicable for copolymer where T
g1
and T
g2
denote the glass
temperature of polymers forming the copolymer. Reduc-
tion in the glass temperature means the enhancement in the
amorphous state at low temperature, and therefore improves
the conductivity at low temperatures. Figure 2 shows the
enhancement of conductivity by the addition of plasticizer
tetraglyme on the system of PEO-co-PPO (3:1):LiClO
4
,
measured at 25
o
C [29]. The decrease in the glass transi-
tion results in the improvement in the fraction of amor-
phous state at room temperature, therefore improving the
conductivity.
However, an improvement in conductivity is adversely
accompanied by a degradation in solid-state conguration
and a loss of compatibility with the lithium electrode [9],
0 25 50 75 100
180
200
220
wt.% tetraglyme
T
g
[
K
]
Figure 1. Effect of plasticizer weight fraction on the glass temperature
of a PEO-co-PPO (3:1):LiClO
4
using plasticizer tetraglyme (tetraethy-
lene glycol dimethyl ether). Data points were derived from [29], D. R.
MacFarlene et al., Electrochim. Acta 40, 2131 (1995).
0
10
5
10
4
10
3
10
2
25 50 100 75
wt.% tetraglyme


[
S
/
c
m
]
Figure 2. Effect of plasticizer weight fraction on the conductivity at
25
o
C of a PEO-co-PPO (3:1):LiClO
4
using plasticizer tetraglyme
(tetraethylene glycol dimethyl ether). Data points were extracted from
[29], D. R. MacFarlene et al., Electrochim. Acta 40, 2131 (1995).
particularly when the fraction of plasticizers is too high.
For example, the modulus of elasticity and elastic strength
signicantly decreases by addition of plasticizers. This is
because the plasticizers are usually low molecular weight
polymer having low mechanical strength. Therefore, addi-
tion of plasticizers decreases the mechanical strength of the
host polymer. Figure 3 shows the effect of plasticizer tro-
glyme content on the elastic modulus and tensile strength
of PEO-co-PPO (3:1):LiClO
4
[29]. The use of moderate or
large quantities of plasticizer results in the production of
Elastic
modulus
Tensile
Strength
0 20 40 60
0.0
1.0
2.0
wt.% tetraglyme
M
o
d
u
l
u
s
[
M
p
a
]
Figure 3. Effect of plasticizer weight fraction on the modulus of a PEO-
co-PPO (3:1):LiClO
4
using plasticizer tetraglyme (tetraethylene glycol
dimethyl ether). Data points were extracted from [29], D. R. MacFar-
lene et al., Electrochim. Acta 40, 2131 (1995).
Polymer Electrolyte Nanocomposites 735
gel electrolyte. The presence of some plasticizer may also
give rise to problems caused by its reaction with the lithium
anode. The poor mechanical stability was accounted to be
mainly due to the solubility of the polymer matrix in the
plasticizer [30]. Cross-linking of the polymer with ultraviolet
radiation [31], thermally [32], by photopolymerization [33],
or electron beam radiation polymerization [34] was found
to reduce the solubility of polymer in the solvent and also
helped to trap liquid electrolytes within the polymer matrix.
3. DEVELOPMENT OF POLYMER
ELECTROLYTE NANOCOMPOSITES
Currently, one popular approach to improve the conduc-
tivity involves dispersing ceramic llers (solid plasticizers)
in the polymer matrix, producing what is currently known
as composite polymer electrolytes. This approach was rst
introduced by Weston and Steele [35]. Ceramic ller was
used to reduce the glass transition temperature and crys-
tallinity of the polymer and thus allow the amorphous
polymer to maintain the liquid-like characteristic at the
microscopic level. Ceramic llers that are frequently used
have particle sizes in the range of about several ten nano-
meters up to several micrometers. Fortunately, such ller
materials are commercially available in various sizes at low
prices. Figure 4 shows the effect of ller content on the con-
ductivity of polymer electrolytes PEO:LiClO
4
[9]. Table 3
displays examples of polymer electrolyte nanocomposites
and their conductivities at around room temperature.
The inorganic ller also acts as a support matrix for the
polymer, so that even at high temperature, the composite
remains solid. However, at the microscopic level, it main-
tains a liquid-like structure, which is important for suf-
cient conductivity. The ller particles, due to high surface
area, prevent the recrystallization of polymer when annealed
above the melting point. The acid-base interaction between
the ller surface group and the oxygen of the PEO leads to
a Lewis acid characteristic of the inorganic ller and favors
2.4 2.6 2.8 3.0 3.2 3.4 3.6
8
6
4
2
L
o
g


[
S
/
c
m
]
1000/T [1/K]
VTF
Arrhenius
Figure 4. Arrhenius plot of electrical conductivities of: (solid) ceramic-
free PEO:LiClO
4
, (triangle) PEO:LiClO
4
containing 10 wt.% Al
2
O
3
(5.8 nm), and (square) PEO:LiClO
4
containing 10 wt.% TiO
2
(13 nm).
Data points were extracted from [9], F. Croce et al., Nature 394, 456
(1998).
Table 3. Examples of polymer electrolyte nanocomposites with their
corresponding electrical conductivities at around room temperature.
Conductivity
at around
Polymer electrolytes Fillers rt (S/cm) Ref.
PEO:LiBF
4
nano-sized-Al
2
O
3
10
4
micro-sized-Al
2
O
3
10
5
[36]
PEO:LiCIO
4
SiC 10
7
[37]
EO-co-PO:LiCF
3
SO
3
Li
1.3
Al
0.3
Ti
1.7
(PO4)
3
210
4
(40
o
C) [38]
Brached-poly(ethylene silica (12 nm size) 10
7
[39]
imine):H
3
PO
4
PEO:LiClO
4
o-Al
2
O
3
10
5
[40]
PEO:LiClO
4
AlCl
3
10
5
[40]
PEO:LiClO
4
NNPAAM >10
5
[40]
PEO-PEG:LiI Al
2
O
3
10
6
[41]
PEO-PMMA:EC:LiI Al
2
O
3
10
8
[41]
PEG:LiCF
3
O
4
SiO
2
10
5
[42]
PEG:LiCF
3
O
4
C
12
H
25
OSO
3
Li 5 10
5
[43]
coated-SiO
2
EC:PC:PAN:LiAsF
6
porous zeolite 10
3
[44]
PEO:LiClO
4
AlN, BaTiO
3
, Bi
2
O
3
10
7
10
6
B
4
C, BN, CaSiO
3
CeO
2
, Fe
2
O
3
, MoS
2
,
PbTiO
3
, Si
3
N
4
,
PEO:LiClO
4
carbon black [46]
PEO:Li[(SO
2
CF
3
)
2
N] -LiAlO
2
4 10
6
[47]
PEO:PMMA:EC:LiI MgO 10
7
[48]
PEO:AgSCN Al
2
O
3
8.8 10
4
[49]
PEO:AgSCN Fe
2
O
3
1.1 10
5
[50]
PEO:AgSCN SO
2
3 10
6
[51]
PEO:NaClO
4
Na
2
SiO
3
2 10
6
[52]
PEO:LiCF
3
SO
3
mineral clay 10
3
[53]
PEO:NH
4
I PbS 0.99 10
6
[54]
PEO:NH
4
I CdS 0.96 10
6
[54]
PEO:NH
4
I Pb
x
Cd
1x
S 0.630.84 10
6
[54]
the formation of complexes with PEO. The ller then acts
as cross-linking center for the PEO, reducing the tension of
the polymer for self-organization and promoting stiffness.
On the other hand, the acid-base interaction between the
polar surface group of the ller and electrolyte ions probably
favors the dissolution of the salt.
Another potential application of polymer electrolyte
nanocomposites is for making solar cells [55]. Dye-sensitized
solar cells have attracted great scientic and technologi-
cal interest as potential alternatives to classical photovoltaic
devices. The cell operation mechanism involves absorp-
tion of visible light by the chemisorbed dye, followed by
the electron injection from the excited synthesizer into the
semiconductor conduction band. The selection of liquid
electrolytes, usually containing organic solvent such as ace-
tonitrile and propylene carbonate, assures the perfect regen-
eration of the dye by direct interaction of the dye oxidized
state and I

/I

3
redox couple and leads to impressively high
solar-to-electrical conversion efciencies (711%) [56, 57].
However, the stability and long-term operation of the cell
are affected by solvent evaporation or leakage. Thus com-
mercial exploitation of these devices needs the replace-
ment of the liquid electrolyte by a solid charge transport
medium, which not only offers hermetic sealing and stabil-
ity but also reduces design restriction and endows the cell
736 Polymer Electrolyte Nanocomposites
with shape choices and exibility. Katsaros et al. investi-
gated solid-state dye-sensitized solar cells using composite
polymer electrolytes using PEO and TiO
2
in the presence
of I

/I

3
redox couple [55]. Initially, dye:Ru(dcbpy)
2
(NCS)
2
(dcbpy is 4,4-dicarboxylic acid-2-2
/
-bipyridine) was attached
on the surface of TiO
2
nanoparticles by immersion of the
TiO
2
thin-lm electrode overnight in ethanolic solution of
the complex, followed by drying. The functionalized TiO
2
nanoparticles, I

/I

3
, and PEO were put in acetonitrile, fol-
lowed by heating and drying to evaporate the solvent. Maxi-
mum incident photon to current efciencies as high as 40%
were obtained at 520 nm, only two times lower that than
obtained using liquid electrolytes [58]. The overall conver-
sion efciency was 0.96%. For all-solid-state devices, such
efciency can be considered to be sufciently high [59].
4. PREPARATION METHODS
Now we will briey explain several methods of preparation
of polymer electrolyte nanocomposites that are commonly
used. Which method should be used, of course, depends on
the materials and the form of sample to be formed. One
method can only produce sample in the form of thick lm,
and another one can produce a sample in the form of lm
of submicrometer thickness.
4.1. Casting Method
This method is frequently used due to its simplicity. It can
produce polymer lm from several micrometers up to sev-
eral millimeters thickness. Generally, this method includes
the following steps:
(a) dispersion of ceramic llers in a salt solution,
(b) addition of a specied amount of polymer to the
mixture,
(c) mixing by means of stirrer or ultrasonic equipment to
disperse the particles homogeneously in the polymer
matrix,
(d) casting the mixture on a substrate,
(e) nally drying in vacuum or in an atmosphere of argon.
All these steps are usually performed in a glove box lled
with argon gas and excluding oxygen and water to levels
below 20 parts per million (ppm), to avoid the possible
occurrence of a dangerous reaction between water and
lithium. The solvent must be water-free and should be com-
mon solvent for both the salt and the polymer. Since the
melting point of several polymers is as high as 65
o
C, the
solvent must also easily evaporate so that drying can be per-
formed at temperatures of around 65
o
C. Organic solvents
such as acetonitrile, cyclopentanone, and propylene carbon-
ate, plus inorganic solvents such as thionyl chloride (SOCl
2
),
are typically used.
Sometimes, the insertion of salt is performed after cast-
ing the lm. For example, Ardel et al. prepared PVDF
2801 (Kynar)-based polymer electrolyte composites accord-
ing to the following steps [60]. First, Kynar was dissolved
into cyclopentanone. Nanoparticles of silica were added and
the mixture was mixed for 24 h at room temperature to get
homogeneous slurry. After complete dissolution, the slurry
was cast on the Teon support and spread with the use of
doctor blade technique. To prevent surface irregularities, the
lm was then covered with a box pierced with holes that
allowed a slow evaporation of the cyclopentanone. After
complete evaporation of the cyclopentanone, the polymer
membrane was soaked in a lithium ion solution for 48 h.
Several fresh lithium solutions for each soaking can be used
to ensure a complete impregnation of lithium ion into the
membrane.
4.2. Spin Coating
The spin-coating method is very similar to the casting
method. Instead of casting the lm on a substrate, in this
method, the mixture is dropped on a substrate and placed
in a spin coater that can be rotated at adjustable rota-
tion speed. The lm thickness can be controlled easily by
adjusting the viscosity (concentration) of the mixture and
the speed of rotation. However, this method is only avail-
able if the viscosity of the mixture is not too high. For a gel
mixture, the spin coater rotation is not enough to spread the
mixture droplet to form thin lm.
4.3. Hot Press
Hot press technique equipment is illustrated in Figure 5.
The equipment consists of: (A) weighing cylinder, (B) heat-
ing chamber, (C) basement, and (T) temperature controller.
Proper amounts of polymer, salt, and ller are mixed in a
mortar for about several minutes. The powder mixture is
then sandwiched between two sheets of Mylar or other mate-
rials, and positioned inside the heating chamber that is con-
trolled at temperatures lightly above the melting point of the
polymer. If PEO is used as polymer matrix, temperature of
80
o
C is suitable [61]. The sample is then pressed overnight
with a pressure that can be controlled by weighing cylinder.
After heating and pressing, the sample is then slowly cooled
to room temperature. The sample is then separated from
the Mylar sheet and placed in a glove box.
T
A
B
C Sample
Figure 5. Illustration of hot press equipment: (A) weighing cylinder,
(B) heater, (C) base, and (T) temperature controller.
Polymer Electrolyte Nanocomposites 737
4.4. In-situ Preparation
In-situ preparation explained here is the preparation of
nanoparticles in the polymer matrix. Mikrajuddin et al.
produced polymer electrolytes of polyethylene glycol with
lithium ion by in-situ production of ZnO nanoparticles
in the polymer matrix [62, 63]. The preparation methods
will be briey described as follows. Zinc acetate dihydrate,
(CH
3
COO)
2
Zn 2H
2
O 0.1 M in 100 mL ethanol 99.5, was
heated with stirring in distillation equipment at tempera-
ture of 80
o
C to produce about 60 mL condensate and
40 mL of hygroscopic solution. Lithium hydroxide monohy-
drate, LiOH H
2
O, of various concentrations was suspended
in 40 mL ethanol and stirred until all the granular material
dissolved. Polyethylene glycol (PEG) (Mn = 2,000,000) was
suspended into each LiOH solution and then stirred with
heating at around 60
o
C until homogeneous gel-type mix-
tures were obtained. The mixture temperatures were then
left to go down several minutes, after which 10 mL of hygro-
scopic CH
3
COO
2
Zn 2H
2
O solution was added into each
mixture. The new mixtures were then homogeneously mixed
and then dried in an oven that was kept at temperature of 40
o
C during three days. The schematic of sample preparations
is displayed in Figure 6.
There are many differences between the present method
and the commonly used ones. In the present approach:
(a) Nanoparticles are grown in-situ in polymer matrix.
(b) The size of dispersed particles is controllable. (c) Ion
carriers are inserted in-situ in the polymer matrix during
the growing process. (d) Finally, since the grown nano-
particles are luminescent, we obtain a new class of polymer
electrolytes, namely luminescent polymer electrolytes with
nanoparticles as luminescence centers. Based on the TEM
picture, we found that the size of ZnO nanoparticles was
5 nm.
ZnAc
2
2H
2
O
Ethanol Ethanol
LiOHH
2
O
PEG
Dried at
40C,
3 days
Mixed
at 60C
Distilled
at 80C
Cooled
at 0C
Mixed
several
minutes
Mixed
around
10 min
Left
cooling
unused
ZnO nanoparticles
are formed in the
polymer matrix
Characterizations
Condensate
(60%)
Hygroscopic
solution (40%)
Figure 6. Diagram of in-situ preparation of PEG:Li containing nano-
particles of ZnO. Adapted with permission from [62], Mikrajuddin
et al., J. Electrochem. Soc. 149, H107 (2002). 2002, Elsevier.
Chandra et al. produced polymer electrolytes PEO:NH
4
I
containing nanometer-sized semiconductor particles PbS,
CdS, Pb
x
Cd
1x
S [54]. Methanolic solution of PEO and
NH
4
I was rst stirred roughly at 40
o
C for 810 h, which
resulted in viscous solution of the ion conducting complexes
of PEO/NH
4
I. To this solution, a solution Pb(CH
3
COO)
2
,
Cd(CH
3
COO)
2
, or Pd(CH
3
COO)
2
Cd(CH
3
COO)
2
in a
desired fraction was added. The stirring was continued until
the viscosity was back to the value it was before adding the
acetate compounds. Subsequently, H
2
S was bubbled through
it giving PbS, CdS, or Pb
x
Cd
1x
S. The nal viscous solution
was poured in a petri dish for obtaining solution-cast lm.
Then the lm was dried in vacuum.
5. IMPORTANT PARAMETERS
To bring polymer electrolytes as well as polymer electrolyte
composites, these materials should provide enough values of
several properties as follows.
5.1. Electrical Conductivity
Conductivity denes the density of current that can be trans-
ported in the material by applying a certain electric eld. If
electric eld 1 is applied in the material, the current density
will be proportional to the applied electric eld, where the
proportional constant is the conductivity, or,
J = u1 (3)
with J the current density (A/m
2
) and u (S/m or S/cm) the
electrical conductivity. It is clear that high conductivity mate-
rial will produce high current density upon applying a certain
electric eld. The value of conductivity is determined by the
density of mobile ions (ion carriers) in the material (n), the
scattering time of the ion (t), the ion charge (q), as well as
the mass of ion carrier (n), according to a relation
u =
nq
2
t
n
(4)
This equation gives the reason why most polymer elec-
trolytes use lithium ions as ion carriers. The mass of lithium
is the smallest among all metals, so it produces the highest
conductivity.
For industrial application, the conductivity of polymer
electrolytes must be as high as 10
2
S/cm. However, until
presently, this conductivity can only be achieved at high tem-
peratures in which the polymer is present in the soft phase,
or even liquid phase. The conductivity at room temperature
of most reported polymers is still below 10
4
S/cm.
5.2. Transference Number
Since the electrochemical process in lithium batteries
involves the intercalation and de-intercalation of lithium
cations throughout host compound lattice, solid polymer
electrolytes with cation transference number (i

) approach-
ing unity are desirable for avoiding a concentration gradi-
ent during repeated charge-discharge cycles. The reported
i

value for dried polymer electrolytes range from 0.06 to


0.2 [64]. For a gel polymer system, i

value of 0.40.5 has


738 Polymer Electrolyte Nanocomposites
been found for poly(bis-methoxy ethoxy)phosphazene [65],
and 0.56 in a system of UV-cured gel polymer electrolytes
based on polyethylene glycol diacrylate/polyvinylidene uo-
ride [66].
Transference number of a particle is dened as the ratio
of the conductivity due to it and the total conductivity.
Assume the total conductivity u is due to ionic, u
ion
, and to
electronic, u
e
, then
u = u
ion
u
e
(5)
The ionic and electronic transference numbers are then
i
i
=
u
ion
u
(6)
and
i
e
=
u
e
u
(7)
For pure ionic, i
i
= 1, and for pure electronic, i
e
= 1. For
polymer electrolyte composites, a general condition satised
is 0 - i
i
, i
e
- 1.
5.3. Crystallinity
Crystallinity plays an important role in determining the
conductivity of polymer electrolytes. At crystalline phase,
the transport of ion carriers is very difcult so that the
conductivity is very low. At amorphous phase, there is a
segmental motion of polymer chain that also assists the
displacement of ions. As a result, the transport of ions is rel-
atively easy. Thus, high conductivity will result. One major
route to improve the conductivity of polymer electrolytes
is by increasing the fraction of amorphous states. Addition
of ceramic llers, addition of plasticizer, and production of
branch polymer are efforts to improve the amorphous state
in the polymer.
5.4. Mechanical Strength
One objective of the use of polymer electrolyte is to make a
battery or fuel cell with a strength comparable to that of liq-
uid electrolytes. Therefore, it is expected that the improve-
ment of conductivity is not accompanied by a decrease in
the mechanical strength. It is why the addition of ceramic
ller has received more attention, since the conductivity and
the mechanical strength can be improved simultaneously.
In contrast, the use of liquid plasticizer, although it can
enhance the conductivity much higher than the addition of
ceramic ller, involves such a degradation in the mechanical
strength as to make this approach less interesting.
5.5. Storage Time
Battery or fuel cell made from polymer electrolytes should
have to operate several weeks or several months. Thus, the
properties of polymer electrolytes should not change too
much during this time. For example, the conductivity should
not depend so much on the storage time. Ideally, the prop-
erties should be time independent. However, in reality, the
properties tend to degrade with storage time.
6. CHARGE TRANSPORT
CHARACTERIZATIONS
Electrical conductivity is the critical parameter for polymer
electrolyte composites. One target of the present research in
this eld is to produce polymer electrolyte nanocomposites
that exhibit a high electrical conductivity, especially at room
temperature. Conductivity at around 10
2
10
3
S/cm is
required to bring this material into industry. The electrical
conductivity relates to the value of current that can be pro-
duced by the battery. The potential produced by the battery
depends on the reaction of the battery with the electrode.
Even though the electrode reaction can produce high elec-
trical potential, the use of low conductive electrolytes can
produce only small amount of electric current. And since
the power can be calculated simply by the relation Power =
Voltage Current, the use of low conductive materials will
produce a low specic energy battery.
6.1. d.c. Conductivity
Ideally, the d.c. conductivity should be measured in order
to be sure that the values pertain to long-range ion move-
ment instead of dielectric losses such as would be associ-
ated with limited or localized rattling of ions within cages.
However, the difculty in making a d.c. measurement is
in nding an electrode material that is compatible with
the electrolyte composites. For example, if stainless steel
electrodes are attached to an electrolyte composite, as dis-
played in Figure 7a, and small voltage is applied across
the electrodes, Li

ions migrate preferentially toward the


cathode, but pile up without being discharged at the stain-
less/electrolyte interface. A Li

ion decient layer forms at


the electrolyte/stainless steel interface.
The cell therefore behaves like a capacitor. There is an
accumulation of ions at interface region of electrode and
composite. A large instantaneous current 1
t
presents when
the cell is switched on, whose magnitude is related to the
applied voltage and the resistance of the electrolytes but
then falls exponentially with time, as illustrated in Figure 7b.
The characteristic time of current decreasing is relatively fast
so that it is difcult to make an accurate measurement.
Therefore, the a.c. method is commonly used in the
present to make measurement over a wide range of fre-
quencies. The d.c. value can be extracted from the a.c.
data. Many a.c. measurements are performed with blocking
I
I
o
time
+
+
CPE
(a) (b)
Figure 7. (a) Polymer electrolyte composites sandwiched between two
blocking electrodes. (b) The decay of current when a constant d.c. volt-
age is applied between two electrodes.
Polymer Electrolyte Nanocomposites 739
electrode such that no discharge or reaction occurs at the
electrode/electrolyte interface. Because the current will ow
back and forth, no ions pile up on electrode surface, espe-
cially when using a high a.c. frequency. This is why the a.c.
resistance (impedance) tends to decrease with increase in
the frequency. The electrodes that are commonly used are
platinum, stainless steel, gold, and indium tin oxide (ITO)
glass. The complex impedance method is widely used to
determine the resistance of the sample. The principle of the
method is based on measurements of cell impedance, which
are taken over a wide range of frequency and then analyzed
in the complex impedance plane which is useful for deter-
mining the appropriate equivalent circuits for a system and
for estimating the values of the circuit parameters.
Impedance is nothing but the a.c. resistance of the cell.
The value in general contains the real and the imaginary
part. An electrochemical cell, in general, exhibits resistive,
capacitive, as well as inductive properties. The resistive
property contributes to the real part of the impedance, while
the capacitive and the inductive properties contribute to
the imaginary part of the impedance. Therefore, an elec-
trochemical cell can be considered as a network of resis-
tor, capacitor, as well as conductor. Which arrangement for
which cell is usually determined after performing a measure-
ment, by analyzing the form of impedance curve. A capac-
itor that presents as an open circuit in a d.c. network and
an inductor that appears as a straight conductor wire in a
d.c. circuit, both appear as imaginary resistors in the a.c.
circuit. Until presently, the inductive properties of the elec-
trochemical cell are ignored so that the polymer electrolyte
composite is considered only as a network of resistor and
capacitor.
The complex impedance can be written in a general form
as
Z(w) = Z
/
(w) iZ
//
(w) (8)
where w is the frequency, Z
/
(w) is the real part of
impedance, contributed by resistive part, Z
//
(w) is the imag-
inary part of impedance, contributed by capacitive part, and
i =

1, the imaginary number.


As an illustration, Figure 8 shows examples of simple RC
circuits and the corresponding plot of impedance (Nyquist
plot). For a serial arrangement of a resistor and a capacitor,
as displayed in Figure 8b left, the impedance can be written
as
Z = F
i
wC
(9)
or
Z
/
= F (10a)
Z
//
=
1
wC
(10b)
It is clear that the real part of impedance is constant, inde-
pendent of the frequency, while the imaginary part depends
on the frequency. For very small frequency, the imaginary
part is very large and this value decreases inversely with
frequency. For frequency approaches to innity, the imag-
inary part of impedance closes to zero and the impedance
R
R
R
Z []
Z []
Z []
Z []
Z

]
Z

]
Z

]
Z

]
R

m
C
1
R = 1

m
CR = 1
R
C
1
C
2
C
R
R C
R
(a)
(b)
(c)
(d)
Figure 8. Examples of simple RC circuits and the corresponding
impedance (Nyquist) plots.
value at this very high frequency equals to resistance. Thus
the Nyquist plot for this arrangement appears as a verti-
cal straight line, starting from a lower frequency value at
the upper part downwards when the frequency increases, as
shown in Figure 8b right. The intersection of this line with
horizontal axis (the real value of impedance) corresponds to
the resistance.
For a parallel arrangement of resistor F and capacitance
C, as appears in Figure 8c left, the real and imaginary parts
of the impedance are given by
Z
/
=
F
1 (wFC)
2
(11a)
and
Z
//
= F
wFC
1 (wFC)
2
(11b)
and the corresponding Nyquist plot appears in Figure 8c
right. The Nyquist plot appears as an arc. The intersection
of this arc with the vertical axis at a low frequency (right arc)
corresponds to the resistance. The frequency at the peak of
the arc, w
n
, satises the relation
w
n
FC = 1 (12)
From the intersection point at the low frequency region and
the position of the arc peak, the resistance and the capaci-
tance of the system can be determined.
740 Polymer Electrolyte Nanocomposites
More complex arrangements, have a more complex
expression for the impedance. For example, a combination
of serial and parallel circuit as appears in Figure 8d left has
the impedance as
Z =
_
1
F
iwC
1
_
1

1
iwC
2
(13)
with the corresponding Nyquist plot appearing in Figure 8d
right. It contains a vertical line that intersects the horizontal
axis at Z
/
= F, and an arc with the peak satises w
n
FC
1
=
1. Again, from these two values, one can determine F and
C
1
. The value of C
2
is determined by measuring the vertical
component of impedance at a certain frequency, say w

. If
the vertical component of impedance at this point is Z
//

,
the value of C
2
satises
Z
//

=
1
w

C
2
(14)
Sometimes, the form of curve is not as simple as that
described here. However, in principle, we can nd some
circuit arrangement such that the theoretical Nyquist plot
is in agreement with the measured data. Some com-
puter software is commercially available for extracting the
equivalent circuit for measured data. As an example, the
impedance measurement of a system of PEO:LiCF
3
SO
3
con-
taining Li
1.4
Al
0.4
Ge
1.6
(PO
4
)
3
llers is displayed in Figure 9a
[67]. The corresponding a.c. circuit that can produce this
impedance data appears in Figure 9b, with F
b
= bulk resis-
tance, F
pb
= phase boundary resistance, F
int
= interfacial
resistance, C
pb
= phase boundary capacity, C
int
= interfacial
capacity, Z
d
= diffusive impedance. The corresponding
parameter values that can properly t the measured data are
0 0.5 1.0 1.5 2.0 2.5 3.0
0
0.2
0.4
0.6
Z [k]
Z

[
k

]
C
int
R
int
R
pb
R
b
Z
d
C
pb
Figure 9. (Top) Impedance plot PEO:LiCF
3
SO
3
containing
Li
1.4
Al
0.4
Ge
1.6
(PO
4
)
3
obtained from experiment. Data points were
extracted from [67], C. J. Leo et al., Solid State Ionics 148, 159 (2002).
(Bottom) The suggested RC circuit for data in (a). See text for the
explanation of symbols.
F
b
= 593 H, F
pb
= 1637 H, C
pb
= 31 nF, and C
int
= 1.9 F
[67].
From the measured resistance of polymer electrolytes,
the electrical conductivity can be calculated using a simple
equation
u =
1
F
e

.
(15)
where F
e
= resistance of polymer electrolyte, = material
thickness (electrode spacing), and . = material cross sec-
tion.
The common procedure for measuring the temperature
dependence of conductivity is as follows.
(a) Heat the sample at a required temperature. Some-
times, it needs a half hour or more to equilibrate the
sample temperature.
(b) Measure the impedance at all range of frequency.
Sometimes, it can take from tens of mHz up to sev-
eral MHz. The computerized measurement is usually
performed since a great number of data should be
collected for each setting temperature.
(c) Change the setting temperature, and again collect the
impedance data in all frequency regions.
(d) Analyze the collected data and nd the equivalent
circuit.
(e) Determine the resistance of the electrolytes based on
the impedance plot and the equivalent circuit at each
setting temperature.
(f) Calculate the conductivity at each setting tempera-
ture.
6.2. a.c. Conductivity
Despite the d.c. conductivity, the a.c. conductivity sometimes
gives important information such as the dielectric proper-
ties of the composites. The frequency dependence of a.c.
conductivity in polymer electrolytes can be written as [38]
u
ac
(w) = u
Jc
.w
n
(16)
where u
Jc
= d.c. conductivity, . and n are the material
parameters, 0 - n - 1, and w is an angular frequency.
The curve might consist of three regions, a spike at
low frequency, followed by a plateau at medium frequency,
and another spike at high frequency. The high frequency
part corresponds to bulk relaxation phenomena, while the
plateau region is connected to d.c. part of conductivity. The
lower spike is connected to electrode/electrolyte phenom-
ena. Fitting the curve with Eq. (16), one can determine the
parameters . and n, and from those parameters, the hop-
ping frequency [68],
w

=
_
u
Jc
.
_
1,n
(17)
By tting the experimental data of u
ac
w, one can
determine the u
Jc
and w

at each temperature. Using


this approach, Siekierski et al. [69] found in a system of
Polymer Electrolyte Nanocomposites 741
PEO
3
:LiClO
4
o-Al
2
O
3
, that both u
Jc
and w

satisfy the
Arrhenius expression
C = C
t
exp
_

1
kT
_
(18)
where C is either u
Jc
or w

, and C
t
is the corresponding
prefactor.
Furthermore, the temperature-dependent dielectric con-
stant can also be obtained from the u
ac
data. The real part
of the dielectric constant, :
/
, can be expressed as
:
/
=
u
//
ac
w:
t
(19)
where u
//
ac
is the imaginary part of the a.c. conducttivity,
and :
t
is the permitivity of vacuum. The complex dielectric
constant can be written as :(w) = :
/
(w) i:
//
(w), and the
imaginary part can be obtained from the real part using a
KramerKronig relation
:
//
(w) =
2w
=
F
_
~
0
:
/
(s)
s
2
w
2
Js (20)
where F denotes the principal part of the integral [70]. On
the other hand, if the imaginary part has been known, the
real part can be determined using a relation
:
/
(w) =
2
=
F
_
~
0
s:
//
(s)
s
2
w
2
Js (21)
6.3. Diffusion Coefcient
Electrical conductivity can also be determined by measuring
the diffusion coefcient. From the temperature-dependent
diffusion coefcient, the temperature dependence of elec-
trical conductivity can be determined using NerstEinstein
equation
u =
ne
2
D
kT
(22)
where n = charge carrier concentration e = electron charge,
and D = diffusion coefcient.
Salt diffusion coefcient can be obtained by galvanostati-
cally polarizing a symmetric cell containing no-blocking elec-
trode for a short period of time. For example, assume a cell
containing Li-based polymer electrolytes and lithium elec-
trodes at both sides. When the current is turned off, the
induced concentration prole is allowed to relax. At long
time after the current interrupt, the following equation is
applicable [71]:
ln A4 =
=
2
D
s
1
2
i .
1
(23)
where A4 = measured cell potential, D
s
= salt diffusion
coefcient, 1 = electrolyte thickness, i = time, and .
1
= a
constant.
It appears that D
s
is proportional to the slope of curve ln
A4 with respect to i. The dependence of salt diffusion con-
stant on the salt concentration is displayed in Figure 10 [71]
for system of PEO:NaCF
3
SO
3
at 83
o
C. The D
s
decreases
0.0
8.0
7.8
7.6
7.4
7.2
7.0
0.5 1.0 1.5 2.0 2.5
L
o
g
D
s
[
c
m
2
/
s
]
Salt concentration [mol/L]
Figure 10. Effect of salt concentration on the diffusion coefcient for
PEO:NaCF
3
SO
3
system at 83
o
C. Data points were extracted from [71],
Y. Ma et al., J. Electrochem. Soc. 142, 1859 (1995).
as the salt concentration increases from about 8 10
8
cms
for dilute solution.
Diffusion coefcient can also be determined from the
Nyquist plot as discussed by Strauss et al. [72]. The medium
frequency arc is attributed to the solid/electrolyte interface.
At lower frequencies, the impedance is affected by concen-
tration gradient (diffusion) and ionic aggregates. The dif-
fusion impedance of symmetric electrolyte with no-blocking
electrode, such as Li/CPE/Li, can be written as
D =
FT1
n
2
1
2
C
b
Z
DC
(24)
where F = the gas constant, n = ratio of EO/cations, 1 =
Faraday number, C
b
= bulk concentration of cation, T =
temperature, and 1 = electrolyte thickness.
Lorimer also introduced another formula for calculating
the diffusion constant, that is [73],
D =
w
n
1
2
2.54
(25)
where w
n
= the frequency at the maxima of low frequency
arc, and 1 = electrolyte thickness. The values predicted by
Eq. (25), however, are around 610 times as large as that
predicted by Eq. (14). The error can be contributed by the
shift of w
n
due to the formation of ion pairs [72].
6.4. Transference Number
Transference number can be calculated by analyzing the arc
impedance spectrum of symmetrical cell with no blocking
electrode. The transference number can be calculated by
comparing the width of the skew low frequency semicircle,
Z
J
, with the value of the bulk resistance, that is [74],
i

=
1
1 Z
J
,F
b
(26)
Transference number can also be determined by measure-
ment of the electrochemical potential of the cell as illus-
trated in Figure 11 [75]. Suppose the polymer composite is
742 Polymer Electrolyte Nanocomposites
CPE
Electrode 1 Electrode 2

1

2
Figure 11. A simple experiment for determining the transference num-
ber to electrodes with differential chemical potentials.
sandwiched between two electrodes with different chemical
potential
1
and
2
. The electrochemical potential across
this cell is given by
1 =
1
[z[1
_

2

1
i
i
J =
1
[z[1
i
i
(
2

1
) (27)
where [z[ = absolute value of the valence of the mobile ion
in the electrolyte; and 1 = Faraday number.
For pure ionic composite, i
i
= 1 so that
1
pure
= (
2

1
)[z[
1
1
1
(28)
Thus
1 = i
i
1
pure
(29)
By measuring 1 and calculating 1
pure
, we can obtain i
i
.
Another method based on a combination of d.c. polariza-
tion and a.c. impedance has been introduced by Evans et al.
This method involves measuring the resistance and current
across a symmetrical Li/electrolyte/Li cell polarized by a d.c.
voltage [76]. The i

is given by
i

=
1
S
(I 1
t
F
t
)
1
t
(I 1
S
F
S
)
(30)
where I = d.c. voltage applied to the cell, F
t
= initial resis-
tance of the passivating layer, F
S
= steady-state resistance
of the passivating layer, 1
t
= initial current, and 1
S
= steady-
state current.
The d.c. polarization potential usually used is several tens
of millivolts. This equation is applicable for ideal, dilute
solutions. However, Doyle and Newman state that although
this equation is not strictly applicable in concentrated elec-
trolytes, the ratio of steady-state to initial current provides
useful information on the contribution by organic additives
to the ionic conductivity of polymer electrolytes [77]. The
simplication of Eq. (30) was also used, that is, i

= 1
SS
/1
O
.
However, signicant errors resulted from neglect of kinetic
resistances at the electrode/electrolyte interface [78]. Trans-
ference numbers of some composites appear in Table 4.
7. SPECTROSCOPIC
CHARACTERIZATIONS
7.1. NMR Spectroscopy
A moving ion would substantially modify the interaction of
electromagnetic waves with matter. Investigating this inter-
action gives a better understanding of ion dynamics on
Table 4. Transference number of several composites.
Transference
Composites number Temperature Ref.
PEO:LiCF
3
SO
3
-LiAlO
2
(4 m) 0.29 90
o
C [79]
PEO:LiBF
4
-LiAlO
2
(4 m) 0.26 90
o
C [79]
PEO:LiClO
4
TiO
2
(13 nm) 0.50.6 90
o
C [79]
PEO:LiClO
4
Al
2
O
3
(6 nm) 0.310.33 90
o
C [80]
(PEO)
30
LiClO
4
0.180.19 100
o
C [80]
(PEO)
8
LiClO
4
0.190.20 90
o
C [80]
(PEO)
8
LiClO
4
SiO
2
0.220.23 100
o
C [80]
a microscopic scale. An example of method for studying
the ion dynamics is nuclear magnetic resonance (NMR)
spectroscopy. This method probes the spin of ion using
an electromagnetic wave in radio frequency. In amorphous
single-phase polymer electrolytes, there is usually found a
straight relationship between polymer segmental motion and
ionic mobility by observing a strong correlation between the
onset of NMR line-narrowing and the glass transition [81].
NMR has contributed signicantly to the understanding of
the physical properties of the composite polymer electrolytes
mainly because it offers the possibility to selectively study
the ionic and polymer chain dynamics. For example, mea-
surement of the temperature dependence of
7
Li lineshapes
and spin-lattice relaxation allows the determination of the
activation energy and the correlation time of the cation
motion. Gang et al. described the
7
Li line-narrowing in the
composite of PEO:LiBF
4
-LiAlO
2
(1030 wt%) in the
temperature range of 270270 K [82]. Dai et al. reported
wide line and high resolution solid-state
7
Li NMR [83].
In material, each spin interacts with other spins, giving
rise to spin-spin interaction or relaxation time T
2
. Further-
more, a new thermal equilibrium distribution of the spin,
which has to be mediated through lattice, is forced by the
magnetic eld. The characterization time required for the
excess energy to be given to the lattice or for attainment
of new thermal equilibrium is expressed in terms of spin-
lattice or thermal relaxation time T
1
. Under simultaneous
application of static and radio frequency magnetic eld in
perpendicular direction,
H
z
= H
t
(31a)
H
x
= 2H
1
cos wi (31b)
The interaction of this magnetic eld with the nuclear spin
results in the Bloch susceptibility [84]

/
=
1
2

t
w
t
T
2
T
2
(w
t
w)
1 T
3
2
(w
t
w)
2

2
H
2
1
T
1
T
2
(32a)

//
=
1
2

t
w
t
T
2
1
1 T
2
2
(w
t
w)
2

2
H
2
1
T
1
T
2
(32b)
where w
t
= H
t
, and is the gyromagnetic ratio of the
spin. For low RF eld H
1
,
2
H
2
1
T
1
T
2
_ 1, then Eqs. (32a)
and (32b) give the familiar absorption curve with half-width
Aw = w
t
w = 1,T
2
(33)
However, the exact lineshape and linewidth can be deter-
mined using Van Vleck method of moment. This method
Polymer Electrolyte Nanocomposites 743
allows the connection of the absorption line with the
motional behavior of the nuclei. The second moment, M
2
is
given by
M
2
=
_
~
~
(w w
t
)
2
] (w) Jw
=
3
4

2
1(1 1)
_ (1 3 cos 0
i]
)
2
r
6
i]
(34)
where 1 is the nuclear spin [85].
Nuclear spin and NMR frequencies of some nuclei at
H
t
= 10, 000 Gauss are given in Table 5 [86]. Based on this
data, we can select the magnetic eld H
t
at around 10,000
Gauss (1 Tesla) and frequency w
t
of around 16.574 MHz, so
that the absorption of the target spin can be observed. For
example, in order to detect the absorption of lithium ion in
the lithium-based polymer electrolyte, the magnetic eld H
t
can be set at around 16,574 Gauss.
If the spins are in motion, such as in uid or intramolec-
ular motion of liquid-like lattice, the value of M
2
is small
because of the small time-average local eld component (the
average of cos
2
0
i]
= 1,3). The linewidth decreases as
(w w
t
) M
1,2
2
(35)
If t
c
is the uctuation time, M
1,2
2
t
c
_1 corresponds to rigid
lattice and motional narrowing corresponds to M
1,2
2
t
c
1.
The region with 1,t
c
approximate to M
1,2
2
represents, the
region of the onset of motional narrowing. Motional nar-
rowing is used to study the ionic motion. Considerable
line-narrowing takes place above a certain temperature,
indicating diffusional motion.
A rough estimation of the activation energy for motion
can be obtained by using a modied Bloembergen, Purcell,
and Pound expression [87]:

c
=
1
t
c
=
oAw
tan
_
=
2
_
Aw
2
E
2
.
2
E
2
__ (36)
where
c
= jump frequency, Aw = measured linewidth at
temperature T , . = unnarrowed linewidth of the rigid lat-
tice, E = fully narrowed linewidth, and o is a constant.
Table 5. Nuclear spin and resonance frequency at 10,000
Gauss of several nuclei.
Magnetic resonance
frequency (MHz)
Nuclei Nuclear spin at 10,000 Gauss
1
H 1/2 42,577
6
Li 1 6,265
7
Li 3/2 16,574
17
O 5/2 5,72
19
F 1/2 40,055
23
Na 3/2 11,262
39
K 3/2 1,987
41
K 3/2 1,092
63
Cu 3/2 11,285
65
Cu 3/2 12,090
107
Ag 1/2 1,220
109
Ag 1/2 1,981
Fitting
c
,o with Arrhenius expression,

c
o
=

t
o
exp
_

A1
kT
_
(37)
one obtains the value of A1 based on the Arrhenius plot of

c
,o.
Another expression for determining the activation energy
is given by Hendrickson and Bray [88]:
ln
_
1
Aw D

1
.
_
=
A1
kT
ln
_
1
E

1
.
_
(38)
where D is a temperature-independent constant of a line-
broadening term.
The relation between spin-lattice relaxation time T
1
and
the uctuation time t
c
can be written as [89]
1
T
1
= S
_
t
c
1 w
2
t
t
2
c

4t
c
1 4w
2
t
t
2
c
_
(39)
where S is a constant. For an approximate case, in which
relaxation from a pair of nuclei of xed internuclear spacing
r only is considered, one obtains [90]
S =
_
2
5r
6
_

2
1(1 1) (40)
The spin-spin relaxation time T
2
is given by [87]
1
T
2
= S
_
5
2
t
c
1 w
2
t
t
2
c

t
c
1 4w
2
t
t
2
c

3
2
t
c
_
(41)
To study dynamics, the relaxation time is measured as a
function of temperature so that t
c
can be calculated. Using
an Arrhenius relation
t
c
= t
t
exp(A1,kT ) (42)
the activation energy can be calculated.
Many NMR experiments have been performed on poly-
mer electrolyte nanocomposites [9195]. Figure 12 shows
the Arrhenius plot of spin-lattice relaxation time for
PEO
8
:LiClO
4
polymer electrolytes and the composite pre-
pared with 5.3 wt.% -Al
2
O
3
detected at a Larmor fre-
quency of w
t
= 155.4 MHz [96]. The attempt frequency 1,t
t
can be interpreted as a vibrational frequency, of order of
optical phonon frequency (10
12
10
12
Hz). It can be calcu-
lated (e.g., using Mathematica software) that the function
1,T
1
nds a maximum value at w
t
t
c
= 0.613. The maximum
value of 1,T
1
appears at temperature of T = 330 K for both
samples. Using a Larmor frequency of w
t
= 155.4 MHz,
one obtains t
c
= 0.613/(1.554 10
8
) = 3.94 10
9
s. For
composite of PEO:LiClO
4
carbon black, Franco et al. esti-
mated the relation time of about 4.4 10
9
s based on the
1
H resonance measurement [97]. Using Eq. (42) and assum-
ing 1/t
t
10
12
, we have A1 = 0.24 eV for both samples.
The electrical conductivity can be determined from
Eq. (42) and different form of NernstEinstein equation,
that is,
u =
AJ
2
q
2
6t
c
kT
(43)
744 Polymer Electrolyte Nanocomposites
2.5
10
1
10
0
3.5 4.5 5.5
1000/T [K
1
]
S
p
i
n
-
l
a
t
t
i
c
e
r
e
l
a
x
a
t
i
o
n
r
a
t
e
(
1
/
T
1
)
[
s

1
]
Figure 12. Arrhenius plot of spin-lattice relaxation time (1/T
1
) of
7
Li NMR spectra for (open circle) PEO:LiClO
4
and (solid circle)
PEO:LiClO
4
containing carbon black. Data points were extracted from
[96], A. C. Bloise et al., Electrochim. Acta 46, 1571 (2001).
where A = lithium concentration per unit volume, J =
average ionic jump distance, and q = ionic charge.
For example, the value of A in PEO
8
:LiClO
4
was deter-
mined from the molecular weights of PEO and LiClO
4
and
the density of the electrolytes (j 1.3 g,cm
3
) [96], to result
in A 1.7 10
21
cm
3
. Considering the average Li-Li dis-
tance of 4 [96], the conductivity at 300 K is about 7.1
10
4
S/cm.
Temperature-dependent linewidth of composite of
PO
1.5
LiI 6 v% Al
2
O
3
is displayed in Figure 13 [83]. It
can be seen that there is a drastic change in the line-width
when temperature is changed from 30
o
C to 50
o
C. It is
assumed that the onset of the motional narrowing is at
temperature of 40
o
C. At 20
o
C the sample exhibits a
rigid limit lineshape with baseline due to a distribution
of
7
Li nuclear quadrupole satellite transition [98]. As the
temperature is raised, partial line-narrowing results.
Forsyth et al. reported the effect of ller content on
the
7
Li linewidth for a composite of copolymer trihydrox-
ypoly(ethylene oxide-co-propylene oxide) with an EO/PO
20 0 20 40 60 80 100 120
1
2
3
4
5
6
7
7
L
i
N
M
R
l
i
n
e
w
i
d
t
h
[
k
H
z
]
Temperature [C]
Figure 13. Temperature dependence of
7
Li NMR linewidth of PEO
1.5
LiI containing 6 vol.% Al
2
O
3
. Data points were extracted from [83],
Y. Dai et al., Electrochim. Acta 43, 1557 (1998).
ratio of 3:1 containg LiClO
4
and TiO
2
nanoparticles [99].
The linewidth increases with increase in the ller content
and reaches a plateau after about 10 wt.%, as observed in
Figure 14. Although the linewidth relates to the mobility of
lithium ion, in which the line-narrowing represents mobile
ion, an interpretation that the increase of linewidth relates
to decrease in the mobility of lithium ions will be contradic-
tory with the reported result. The reported result conrmed
that addition of ller at lower content increases the lithium
ion mobility and therefore increases the conductivity. A pos-
sible interpretation is the lithium ion environment changes
with the addition of ller and the linewidth changes are
more likely to reect the changing environment rather than
the changing mobility. A possible interpretation is chemical
shift dispersion, that is, lithium ion occupying many differ-
ent state environments (but magnetically degenerate) such
that a distribution of chemical shift is obtained.
The broad signal of resonance is attributed to crystalline
state, while the narrow component is attributed to amor-
phous phase. This difference contribution can be used to
predict the crystallinity of polymer in composite. For exam-
ple, Singh et al. observed the
1
H resonance of a system
of PEG
46
:LiClO
4
nanoparticles of Mn
0.03
Zn
0.97
Al
2
O
4
[95].
They tted the broad part of the signal with a Gaussian func-
tion and the narrow component with the Lorentzian func-
tion. The crystallinity of the polymer equals to the fraction
of the area of narrow and broad components of the signal.
For pure PEG they found a crystalline of about 83% using
this method.
7.2. Raman Scattering
Raman spectroscopy is important to probe molecules with
anisotropic polarizability. The vibrating atoms are not able
to follow the incident radiation frequency if it is much higher
than the phonon frequency. However, the electron cloud
of the vibrating atom can interact with the frequency of
the incident radiation. These oscillating dipoles can absorb
energy from the radiation eld and re-emit radiation of the
same frequency. This radiation is detected as scattered light
0 4 8 12 16
200
300
400
500
600
wt.% TiO
2
7
L
i
L
i
n
e
w
i
d
t
h
[
H
z
]
Figure 14. Effect of ller weight fraction on the
7
Li NMR linewidth of
copolymer of trihydroxypoly(EO-co-PO) with (EO/PO = 3/1) contain-
ing LiClO
4
and TiO
2
nanoparticles. Data points were extracted from
[99], M. Forsyth et al., Solid State Ionics 147, 203 (2002).
Polymer Electrolyte Nanocomposites 745
and is known as Rayleigh scattering. Moreover, the oscillat-
ing dipoles see the force eld of the vibrating atomic nuclei
also as the nuclei oscillate around the equilibrium position;
the deformability of the electron cloud varies with the oscil-
lation frequency of the nuclei.
The Raman scattering is described in a simple term here.
If a time-dependent electric eld, 1(i), is applied to a
molecule, it produces an induced dipole moment, (i),
(i) = o(i)1(i) (44)
where o(i) is the polarizability of molecule. If the inci-
dent frequency and the polarizability of the molecule change
between o
min
and o
max
of frequency w
int
as a result of its
rotation and vibration, we can write [100]
(i) =
_
o
1
2
Aocos w
int
i
_
1
t
cos wi (45)
with
A: = o
max
o
min
(46)
Therefore, we obtain,
(i) = o1
t
cos wi
1
2
Ao1
t
|cos(w w
ini
)i cos(w w
int
)i] (47)
The induced dipole moment has then components as fol-
lows: unshifted frequency, w, known as Rayleigh line; lower
frequency, w w
int
, known as Stokes line; and higher fre-
quency, w w
int
, known as anti-Stokes line.
Raman spectra can be used to determine the concen-
tration of free ions in the electrolytes. For example, the
Raman spectra of LiCF
3
SO
3
have been tabulated as appear
in Table 6 [101]. A composite containing SO
3
, the Raman
spectrum of the
S
(SO
3
) spectral region of the triate anion
of poly(ethylene oxide) dimetyl ether (400) complexed with
LiCF
3
SO
3
, along with the three-component curves t are
shown in Figure 15. It was explained that the compo-
nent observed at 1032 cm
1
corresponds to free anions not
interacting directly with lithium cations. The component of
1042 cm
1
has been attributed to contact pair and the com-
ponent of 1052 cm
1
has been attributed to Li
2
CF
3
SO
3
triple
ions [102].
Because of the multicomponent nature of the spectrum,
it can be concluded that the ion-ion interaction is present in
Table 6. LiCF
3
SO
3
vibrational assignments.
Band Wavenumber (cm
1
) Assignment

s
(SO
3
) 1033 free ion
1043 monodenate ion pairs, LiX,
also LiX

2
and LiX
2
3
1053 Li
2
X

aggregate
1062 LiX
2
3

as
(SO
3
) 1272 free ion
1257, 1302 monodenate ion pairs, LiX,
also LiX

2
, LiX
2
3
1270, 1308 Li
2
X

aggregate
1288 Li
3
X
2
aggregate
I
n
t
e
n
s
i
t
y
[
a
.
u
.
]
1042 cm
1
1052 cm
1
1060 1050 1040
Raman shift [cm
1
]
1030 1020 1010
1032 cm
1
O:M = 110:1
Figure 15. Raman spectra of LiCF
3
SO
3
. Reprinted with permission
from [102], A. Ferry et al., Electrochim. Acta. 43, 1471 (1998). 1998,
Elsevier Science.
the system even down to the concentration of O:M= 563:1.
The relative amount of anions not interacting directly with
lithium ions, that is, spectroscopically free, increases with
increasing concentration from approximately 22% at O:M
563:1 to 40% at O:M = 110 and then falls off slightly at
higher concentration. The 1042 cm
1
band initially decreases
in relative intensity with increasing concentration and then
levels off at 54% in the upper concentration range.
The fraction of free ion obtained from Raman spectra in
a system of PPO:NaCF
3
SO
3
is displayed in Figure 16 [103],
for a system of hydroxyl end capped PPO with NaCF
3
SO
3
.
7.3. FTIR Spectroscopy
Atoms in solid vibrate at frequency approximately 10
12
10
13
Hz. Vibration mode involving pairs of groups of bonded
atoms can be excited to higher energy by absorption of radi-
ation at appropriate frequency. In the infrared (IR) technol-
ogy, the frequency at the incident radiation is varied and the
200 250 300 350
0.0
0.2
0.4
0.6
0.8
1.0
F
r
a
c
t
i
o
n
o
f
a
n
i
o
n
s
Temperature [K]
(OH-PPO400)
16
:NaCF
3
SO
3
(OH-PPO4000)
16
:NaCF
3
SO
3
Figure 16. Effect of temperature on the fraction of anions obtained
from the Raman spectrometry. Data points were extracted from [103],
H. Ericson et al., Electrochim. Acta 43, 1401 (1998).
746 Polymer Electrolyte Nanocomposites
quantity of radiation absorption or transmission by sample
is observed. In the Raman technique, on the other hand,
the sample is illuminated with monochromatic light, usually
generated by a laser. In order for a particular mode to be
IR active, the associated dipole moment must vary during
the vibrational cycle. Therefore, centrosymmetric vibrational
modes are IR inactive. The principal selection rule for a
vibrational mode to be Raman active is that the nuclear
motions involved must produce a change in polarizability.
In-situ FTIR spectroscopy can also be used to determine
the conductivity of ion in composites. For example, a system
of poly(methyl methacrylate-co-alkali metal methacrylate)-
ethylene carbonate:LiClO
4
is displayed in Figure 17a. A typ-
ical spectrum of carbonyl stretches region to appear in four
peaks at 1730 cm
1
, 1751 cm
1
, 1775 cm
1
, and 1806 cm
1
[104]. The evolution of carbonyl peak was observed with
time under the application of d.c. eld such that the ion
transport in the composite is shown in Figure 17b [104].
The change in the intensity of carbonyl group in the ethy-
lene carbonate represents the different chemical environ-
ments within the composites. The intensities of the peak at
1730 cm
1
, 1775 cm
1
, and 1806 cm
1
were found to be fairly
0 4 8 12
0
10
20
30
40
50
I
n
t
e
n
s
i
t
y
[
a
u
]
Time (min)
1729 cm
1
1750 cm
1
1775 cm
1
1805 cm
1
(b)
2100
EC/LiClO
4
= 3/1
EC/LiClO
4
= 4/1
EC/LiClO
4
= 5/1
2000 1900 1800
Wavenumbers (cm
1
)
I
n
t
e
n
s
i
t
y
1700 1600
(a)
Figure 17. (a) IR spectra of poly(methyl methacrylate-co-alkali metal
methacrylate)-ethylene carbonate:LiClO
4
and (b) the decay of the
observed peaks when the applied d.c. potential was switched off. Data
points we extracted from [104], C. H. Kim et al., Electrochim. Acta 43,
1421 (1998).
constant with time under the electric eld. It indicated that
these three peaks might not be correlated with movement
of charge carriers. However, the intensity of the 1750 cm
1
peak decreased with the time, eventually reaching a limit
value. The decrease in the intensity of the 1750 cm
1
peak
related to the change in the concentration of the plasticizer
solvating the salt. The change in the peak intensity corre-
sponding to various chemical environments is due to the
migration of the charge carrier. The concentration of the
plasticizer to solvate the salt was found to decrease, which
implies that the movement of cations in the electrolytes is
strongly correlated with the motion of the plasticizers.
The change in the peak intensity of the plasticizer sol-
vating salt can be analyzed solving the following transport
equation [104]:
oC
oi
= D
o
2
C
ox
2
1
oC
ox
(48)
where C = concentration of charge carriers, represented by
the intensity of the 1750 cm
1
peak, D = diffusion coef-
cient, = mobility of the charge carriers, and 1 = electric
eld.
The solution of Eq. (48) at a xed position, x = 0, can be
expressed as
C(x = 0, i) = .Eexp
_

(1)
2
4D
i
_
(49)
where . and E are constants. From the slope of curve inten-
sity of 1750 cm
1
peak, one can determine the mobility of
the ion carriers. Measuring the mobility at various ion con-
centrations, we can determine the mobility at various ion
concentrations as appears in Figure 18 [104].
The ion movement can also be analyzed based on the
change in the intensity of 624 cm
1
, arising from ClO

4
stretching band [105]. Kim et al. showed the dependence of
624 cm
1
intensity on time at 25
o
C and found that the mobil-
ity calculated from that data is similar to that obtained from
the 1750 cm
1
absorption of ethylene carbonate [104]. The
change of anionic mobility with the ion content of ionomers
was found to be nearly the same as that of cation mobility.
2 10 12
5.5
5.0
4.5

[
c
m
2
/
s
V
]
Ion content [mol%]
4 6 8
Figure 18. The effect of ion content on the mobility of poly(methyl
methacrylate-co-alkali metal methacrylate)-ethylene carbonate:LiClO
4
determined from the decay curve of 1729 cm
1
peak. Data points we
extracted from [104], C. H. Kim et al., Electrochim. Acta 43, 1421 (1998).
Polymer Electrolyte Nanocomposites 747
For the infrared spectra of PEG:LiClO
4
:AlBr
3
1 mass%,
the
4
(ClO

4
) spectra can be separated into two contribu-
tions with maxima at 623 and 633 cm
1
[106]. The 623 cm
1
band is attributed to free anions and the 633 cm
1
peak is
related to bound or contact (ClO

4
) [107]. The fraction of
free anions can be calculated as the fraction of area under
623 cm
1
mode to the total area of
4
(ClO

4
) envelope.
Compared to composite in the absence of AlBr
3
, there is a
dramatic increase in the fraction of free anions when AlBr
3
is dispersed in the electrolytes.
7.4. X-ray Photoelectron Spectroscopy
X-ray photoelectron spectroscopy (XPS) is a powerful tech-
nique for studying the surface of solids. The data obtained
using this technique are mainly used to extract the infor-
mation regarding the bonding energies of various core-level
electrons from different elements of solid materials. These
values are then interpreted as the bonding between the ele-
ment under consideration with their neighbor. Information
about the local structure and interaction of an element with
its neighbor can be extracted from the XPS data [108].
In the development of battery, the interaction between the
electrolyte and the electrode determines the performance
of the battery. Understanding this interaction is important
to optimize the preparation parameters for realizing high-
performance batteries. Therefore, surface studies, for exam-
ple using XPS, may be of great signicance to understand
the interaction of polymer electrolytes with electrode.
In principle, this technique measures the kinetic energy of
electrons that are emitted from matter as a consequence of
bombardment with ionizing radiation or high-energy parti-
cles. If the process results in an ionization of electrons from
the bombarded material, the kinetic energy of the electron
will satisfy
1
k
= | 1
b
(50)
where | = energy of incident radiation, and 1
b
= binding
energy of electron.
For a given atom, a range of 1
b
values is possible corre-
sponding to ionization of electrons from different inner and
outer valence shells. Measurement of the value of 1
k
, and
therefore 1
b
, provides a means of identication of atoms.
In the XPS method, the ionizing radiation is usually MgK
o
(1254 eV) or AlK
o
(1487 eV) monochromatic radiation.
Using XPS method, Vosshage and Chowdary investigated
the interaction of salt with the polymer chain in the systems
of PEO
n
LiCF
3
SO
3
and PEO
n
Cu(CF
3
SO
3
)
2
[108]. There is
evidence of the interaction of the carbon from CH
2
-CH
2
-
O- polymer chains with the cation of the salt through the
ether oxygen of PEO and anion of the salt. A complexation
between the action of the salt and the oxygen of the PEO
is identied using this method. It is also observed that for
the PEO
n
Cu(CF
3
SO
3
)
2
system, the change in the nature of
the complex occurs at low salt concentrations. However, for
PEO
n
LiCF
3
SO
3
system, such a change is identied at higher
salt concentration.
Liu et al. prepared polymer electrolyte nanocomposites
using SiO
2
ller and 2-[methoxy(polyethylenoxy)-propyl]
trimethoxy silane coated SiO
2
[109]. The XPS method has
been utilized to conrm the success of coating. During treat-
ment with 2-[methoxy(polyethylenoxy)-propyl] trimethoxy
silane, the OH- groups of the surface of the SiO
2
react with
partially hydrolized silane. For the untreated SiO
2
, the XPS
data can be tted by band centering at 284.6 eV, which is
often characteristic of adventitious CH
2
species in the mea-
surement chamber. On the other hand, the coated SiO
2
can
be tted with two bands, centered at 284.6 eV and 286.2 eV.
The former corresponds to the adventitious CH
2
species
while the latter corresponds to C-O-C group that appears as
a result of the functionalization reaction.
7.5. X-ray Diffraction
X-rays are electromagnetic radiation with a wavelength
around 1 (1010 m). The X-rays which are used in almost
all diffraction experiments are produced by accelerating an
electron beam through 30000 V and permitting it to strike
a metal target, such as copper. The incident electron has
sufcient energy to ionize some of the copper 1s (K shell)
electrons. An electron in an outer orbital (2 or 3) imme-
diately drops down to occupy the vacant 1s level and the
energy released in the transition appears as X-radiation. The
transition energies have xed values and so create the spec-
tra of characteristic X-radiation. For copper, the 2 1s
transition, called K
o
, has a wavelength of 1.5418 and the
3 1s transition, K
p
, has a wavelength of 1.3922 .
If the average crystalline size is below a critical limit
(200 nm diameter), a broadening of diffraction X-ray
beam occurs [110]. The commonly accepted formula for par-
ticle size broadening is the Scheerer formula
J =
0.9\
Ecos 0
E
(51)
where J = crystalline radius, \ = X-ray wavelength, 0
E
=
Bragg angle, E = the line-broadening, given by Warren for-
mula
E
2
= E
2
M
E
2
S
(52)
where E
M
= measured peak width in radiation at half peak
height, E
S
= width of a peak of a standard material mixed
with the sample, whose particle size is considerably greater
than 200 nm and which has a diffraction peak near the rel-
evant peak of the sample. For relatively large particle, the
width of peak (unbroadened peak) is very small, so that
E
M
E
S
. Therefore, it is often possible to approximate E
with E
M
.
Figure 19 shows the XRD spectra of polyethylene glycol-
based polymer electrolyte nanocomposites containing ZnO
nanoparticles [62]. Using 110) with a position at around
20
E
= 56.7
o
, and the width of the peak of around A(20
E
) =
1.25
o
, we obtain 0
E
= 0.5 rad and A0
E
= 0.022 rad. Thus the
predicted nanoparticle size is around J = 7.2 nm.
The degree of crystallinity of polymer can be determined
also by wide-angle X-ray scattering (WAXS). Figure 20 illus-
trates the XRD spectra of polymer, where the relatively
sharp peaks are due to scattering from the crystalline regions
and the broad underlying hump is due to scattering from
noncrystalline region. The degree of crystallinity can be pre-
dicted based on the measurement of the area under the
sharp peak (.
c
) and the broad hump (.
a
) and using a
748 Polymer Electrolyte Nanocomposites
50 60 70
0
40
80
120
2 []
55 65
I
n
t
e
n
s
i
t
y
[
a
u
]
Figure 19. XRD pattern of PEG containing Li ions and ZnO nano-
particles. Data were extracted from [62], Mikrajuddin et al., J. Elec-
trochem. Soc. 149, H107 (2002).
simple equation
x
c
=
.
c
.
c
.
a
(53)
Shin et al. observed the effect of ller content on the XRD
pattern on PEO in a system of PEO
10
LiCF
3
SO
3
contain-
ing TinO
2n1
. As observed in Figure 21, the PEO peak
intensities decrease by the increase in the volume fraction
of ller content [111]. It indicates that the crystallinity of
the sample decreases by increasing the volume fraction of
ller content. Similar results have also been reported by
Leo et al. in a system of PEO:LiCF
3
SO
3
containing ller
of Li
1.4
(Al
0.4
Ge
1.6
)(PO
4
)
3
[112]. With addition of ller, the
intensity of the crystalline peaks has decreased and a notice-
able broadening of the area under the peak was observed.
It is a clear indication of the reduction of the crystallinity of
the polymer.
8. MICROSCOPIC ANALYSIS
8.1. Scanning Electron Microscopy
The morphology of the sample surface can be observed
using scanning electron microscopy (SEM). The surface
smoothness and the presence of holes in the scale down to
2
I
n
t
e
n
s
i
t
y
[
a
u
]
A
a
A
c
Figure 20. Typical form of XRD pattern of high molecular weight poly-
mer.
10
(a)
(b)
(c)
(d)
20 30
2 (degree)
I
n
t
e
n
s
i
t
y
(
A
.
U
.
)
40 50
Figure 21. XRD patterns of PEO
10
LiCF
3
SO
3
polymer electrolytes with
(a) 0, (b) 5, (c) 10, and (d) 15 wt.% Ti
n
O
2n1
. (, crystalline of PEO).
Reprinted with permission from [111], J. H. Shin et al., Mater. Sci. Eng.
B 95, 148 (2002). 2002, Elsevier Ltd.
several tens of nanometers can be viewed using the advanced
eld emission SEM. Wen et al. compared the SEM pic-
ture of PEO:LiClO
4
and PEO:LiClO
4
containing alumina
whisker. For PEO:LiClO
4
, great amounts of microcracks
were observed on the surface, as observed in Figure 22a
[113]. The size of islands are several micrometers, com-
pared to the PEO spherulite [114]. Addition of 10 wt.%
(a) (b)
(c)
Figure 22. SEM photograph of (a) PEO:LiClO
4
, (b) PEO:LiClO
4
con-
taining 10 wt.% whisker, and (c) PEO:LiClO
4
containing 20 wt.%
whisker. Reprinted with permission from [113], Z. Wen et al., Solid
State Ionics 148, 185 (2002). 2002, Elsevier Ltd.
Polymer Electrolyte Nanocomposites 749
whisker in the polymer electrolytes inhibited the formation
of microcracks inside the composite polymer electrolytes
(Fig. 22b). Further addition of the amount of the whisker,
for example 20 wt.%, is more effective in avoiding the for-
mation of microcracks (Fig. 22c).
Golodnitsky et al. observed the SEM picture of ller-free
PEO
n
:LiI and PEO
n
:LiI containing alumina [115]. PEO
n
:LiI
containing high concentration salt is made up of units whose
area is hundreds of square microns. Addition of alumina
causes a minor reduction in the grain size. In a PVdF gel
polymer electrolyte containing CuO ller, Wang and Gu
found that the surface presented a multitude of PVdF grains
and pores of average size of about 24 m in diameter [116].
They assumed the CuO nanoparticles distributed uniformly
in the polymer matrix.
8.2. Transmission Electron Microscopy
In the case of polymer electrolyte nanocomposites, transmis-
sion electron microscopy (TEM) is usually used for deter-
mining the dispersion of nanoparticles in the polymer matrix
and the size of nanoparticles. Capiglia et al. showed that
larger ller (submicron size) is quite well dispersed in the
polymer matrix while the smaller ller (tens of nanometers
size) is not well distributed in the polymer matrix [61].
Instead, the small ller is condensed in large blocks of size
up to 1 m. Based on the TEM photograph, Mikrajuddin
et al. showed that PEG based polymer electrolyte nano-
composites made by in-situ growth of ZnO nanoparticles in
the polymer matrix and in-situ insertion of lithium ion dur-
ing nanoparticles growth have particle size of about 6 nm
[63]. This result is consistent with the calculation of the par-
ticle size using size-dependent bandgap equation [117119].
Chandra et al. also observed the size of nanoparticles syn-
thesized in-situ in polymer electrolyte nanocomposites using
TEM and found that small content nanoparticles (about
1 wt.%) have smaller size [54].
8.3. Atomic Force Microscopy
There are not many reports on the atomic force microscopy
(AFM) investigation of polymer electrolytes. Instead of
investigating the polymer electrolytes themselves, Granvalet-
Manchini et al. investigated the change in the surface of
lithium electrode when making contact with polymer elec-
trolytes [120]. After about three days contact with polymer
electrolytes, self-assembled polymer layer is developed on
the surface of lithium electrode.
8.4. Optical Microscopy
Optical microscopy characterization of polymer electrolyte
nanocomposites was not reported too much. To date, Kim
et al. reported the optical micrograph of a system of PEO
16
:
LiClO
4
containing various ller content, taken under crossed
polarizers [121]. They observed well-dened spherulitic mor-
phologies. The spherulites were observed in thin lms
deposited on the glass substrate whose typical thickness was
about 20 m. They proposed that the size and the morphol-
ogy of the spherulites can be related to the melting point or
glass temperature of the composites. This was based on the
fact that the spherulites have a lamellar structure for almost
all polymers and the increase in the lamella thickness results
in the increase in the melting temperature [122].
9. THERMAL CHARACTERIZATIONS
9.1. Thermogravimetry
Thermogravimetry (TG) is a technique for measuring the
change in the weight of a substance as a function of tem-
perature or time. The result usually appears as a continuous
chart record, as displayed in Figure 23. The sample, usually
a few milligrams in weight, is heated at a constant rate, typ-
ically 120
o
C/min. It has constant weight until it begins to
decompose at temperature T
i
. Decomposition usually takes
place over a range of temperature T
i
to T
]
and second con-
stant plateau is then observed above T
]
, which corresponds
to the weight of residue.
A TGA curve of composite made by PEO:LiBF
4
contain-
ing 2-[methoxy(polyethylenoxy)-propyl]trimethoxy silane-
coated SiO
2
nanoparticles is displayed in Figure 24 [109].
At temperatures below 180200
o
C, the sample experiences
only a small weight loss due to removal of residual water
on the surface of SiO
2
. At increasing temperatures, the OH
groups on the SiO
2
surfaces begin to decompose to give
rise to slight increase in the weight loss. In addition to the
decomposition of the OH groups, another weight loss take
place at about 350
o
C that can be attributed to the decompo-
sition of silane molecules that are bonded on the surface of
SiO
2
. Liu et al. suggested that the adsorption of containing
2-[methoxy(polyethylenoxy)-propyl]trimethoxy silane on the
surface of SiO
2
is likely to give a sub-monolayer coverage
[109].
9.2. Differential Thermal Analysis
Differential thermal analysis (DTA) is a technique in which
the temperature of a sample is compared with that of inert
reference material during a programmed change of temper-
ature. The temperature of sample and reference should be
the same until some thermal event such as melting, decom-
position, or change in crystal structure occurs in the sample.
W
T
i
W
i
W
f
T
f
Temperature
W
e
i
g
h
t
Figure 23. Typical form of TGA curve.
750 Polymer Electrolyte Nanocomposites
0 100 200 300 400 500
Temperature [C]
96
97
98
99
100
W
e
i
g
h
t
p
e
r
c
e
n
t
a
g
e
[
%
]
OH-decomposition
silane decomposition
Figure 24. TGA curve of PEO:LiBF
4
containing 2-[methoxy (polyethyl-
enoxy)-propyl]trimethoxy silane coated SiO
2
nanoparticles. The curve
was replotted from [109], Y. Liu et al., J. Power Sources 109, 507 (2002).
If the sample temperature lags behind the reference temper-
ature, the process is called endothermic. On the contrary,
if the sample temperature leads the reference temperature,
the process is known as exothermic. The sample size is usu-
ally a few milligrams and heating and cooling rate is usually
150
o
C/min. The difference in the sample and reference
temperature will appear as Figure 25.
If calorimetric data is required, it is better to use differen-
tial scanning calorimetry (DSC). DSC is very similar to DTA.
A sample and an inert reference are also used in DSC sys-
tem but the cell is designed differently. The sample and the
reference are maintained at the same temperature during
the heating program and extra heat input to the sample (or
to the reference if sample undergoes an exothermic change)
is required to maintain this balance. Enthalpy changes are
therefore measured directly.
Examples of DSC curves for various PEO:LiCF
3
SO
3

clay composites are displayed in Figure 26 [123]. From this
gure, we can extract several parameters like the glass tem-
perature, the melting point, and enthalpy. These values were
found to depend on the ller content, as summarized in
Table 7. A pure PEO has one rst-order endothermic tran-
sition at around 70
o
C, corresponding to the melting of the
Temperature
Heating
Cooling
ENDO
EXO
Polymorphic
change Melting
Solidification Polymorphic
change

T
Figure 25. Typical form of DTA curve.
90 90 120 150 180 60 60
Temperature (C)
(PEO)
8
LiCF
3
SO
3
/clay (wt %)
E
n
d
o
t
h
e
r
m
a
l
30
(a) undoped PEO
(b) 1/0
(c) 97/3
(d) 91/9
(e) 70/30
(f) 50/50
30 0
Figure 26. DSC curve of PEO:LiCF
3
SO
3
containing clay. Reprinted
with permission from [123], H.-W. Chen and F.-C. Chang, Polymer 42,
9763 (2001). 2001, Elsevier Ltd.
PEO crystalline phase. When salt is added, a second minor
endothermic transition was observed at around 140150
o
C,
due to the melting of crystalline complex phase formed by
PEO and LiCF
3
SO
3
[124126]. The melting temperature of
the crystalline PEO phase depends on the ller content. The
T
n
initially shifted to higher temperature when the ller con-
tent increased, and reached the highest value at 9 wt.% clay,
and then decreased with further increase in the clay content.
Table 7. The parameters of polymer electrolyte nanocomposites
extracted from the DSC curve.
Samples T
g
(K) T
n
(K) AH (J/g) A
c
T
nc
(K)
PEO 70.50 165.5
PEO:LiCF
3
SO
3
46 69.84 59.9 36.2 152.0
PEO:LiCF
3
SO
3
50 71.58 68.9 41.6 149.7
clay 3 wt%
PEO:LiCF
3
SO
3
43 77.4 77.8 47.0 156.9
clay 9 wt%
PEO:LiCF
3
SO
3
46 54.93 40.1 24.2 142.6
clay 30 wt%
PEO:LiCF
3
SO
3
48 45.99 29.1 17.6 132.4
clay 50 wt%
Adapted with permission from [123], H.-W. Chen and F.-C. Chang, Polymer 42,
9763 (2001). 2001, Elsevier.
Polymer Electrolyte Nanocomposites 751
The PEO crystallinity (expressed by the area covered by
transition curve) is also dependent on the clay content. The
PEO crystallinity initially increases with increasing clay con-
tent up to 9 wt.% and then decreases with further increase
in the clay content.
Melt-crystallized polymers are never completely crystal-
lized. This is because there are an enormous number of
chain entanglements in the melt and it is impossible for the
organization to form 100% crystalline polymer during crys-
tallization. The degree of crystallinity is therefore of great
technological importance. The degree of crystallinity can be
deduced from the DTA data, in which the melting enthalpy
can be obtained. The crystallinity can be calculated using a
simple equation [127]:
A
c
=
AH
n
AH
n, c
]
PEO
(54)
where AH
n
= melting enthalpy measured, AH
n, c
= melting
enthalpy of 100% crystalline (for PEO, AH
n, c
= 196.4 J/g),
and ]
PEO
= weight fraction of PEO in polymer electrolytes.
10. DENSITY METHOD
Another method for determining the crystallinity is based
on the knowledge of density of crystalline and amorphous
phases as well as the density of the polymer specimen.
The crystallization of polymer from melt is accompanied by
the reduction in the volume due to an increase in density.
The crystals have a higher density than the molten or non-
crystalline polymer since the last two contain also free vol-
ume. Based on this difference, the density method can be
utilized to determine the degree of crystallinity.
If I is volume of polymer specimen, and I
a
and I
c
are
volumes of amorphous and crystalline regions, respectively,
we have I = I
a
I
c
. The relation of polymer specimen
mass and the amorphous and the crystalline region masses
is given by
jI = j
a
I
a
j
c
I
c
(55)
where j = density of polymer specimen, j
a
= density of
amorphous region, and j
c
= density of crystalline region.
By dening the crystallinity as x
c
= j
c
I
c
,jI, we obtain
x
c
=
j
c
j
_
j j
a
j
c
j
a
_
(56)
The density of polymer specimen can be determined by
simply measuring the volume and weighing the mass. The
density of crystalline region can be calculated from the
knowledge of the crystal structure. The density of amor-
phous phase can be determined by measuring the density
of almost completely amorphous polymer, such as polymer
obtained by rapid cooling from polymer melt.
Equation (56), however, is valid if polymer specimen con-
tains no holes which are often present in molded samples.
In addition, since packing of the molecules in amorphous
region is random, it is likely that the density of amorphous
phase depends on the thermal treatment of the specimen.
11. ELECTRICAL PROPERTIES
11.1. Effect of Filler Content on Conductivity
The role of the ceramic ller is to inuence the recrystalliza-
tion kinetics of the polymer matrix chains, thereby ultimately
promoting localized amorphous regions and thus enhancing
the transport of cations [9]. To produce a high fraction of
amorphous state in the composite, the as-prepared compos-
ite is rst heated above the melting point so that all parts
of the polymer are converted to the amorphous state. Dur-
ing the cooling process, the matrix part around particles
remains in the amorphous state, even when the tempera-
ture drops below the melting point. Therefore, a high elec-
trical conductivity would be expected to appear at ambient
temperatures. Indeed, enhancement in conductivity of up to
about three orders of magnitude at low temperatures and
about one order of magnitude at high temperatures has been
reported for the system of poly(ethylene oxide)-LiClO
4
con-
taining ceramic llers [9]. In addition, composites contain-
ing ceramic llers in the nanoscale particle size exhibit both
excellent mechanical stability (promoted by the network of
the llers into the polymer bulk) and high ionic conductivity
(promoted by the high surface area of the dispersed ller).
The volume fraction of ller particles affects the conduc-
tivity of a composite. The symbols in Figure 27 display the
effect of ller loading on the electrical conductivity of a
poly(ethylene oxide)/NaI containing ller of O-Al
2
O
3
[128].
The conductivity increases with an increase in the volume
fraction of llers, reaches a maximum at a certain value of
ller particles, and then decreases toward zero for further
increases in the volume fraction of llers. This observation
0.0
10
9
10
8
10
7
10
6
10
5
10
4
0.1 0.2 0.3 0.4 0.5
Volume fraction of fillers


[
S
/
c
m
]
Figure 27. Effect of ller volume fraction on the electrical conduc-
tivity at 25
o
C for composites of (open circle) PEO:LiClO
4
PAAM
and (solid circle) PEO:NaI O-Al
2
O
3
. Date points were extracted from
Y. Liu et al., J. Power Sources 109, 507 (2002). Curves were obtained
from theoretical calculation to t the data using (i/F = 0.6) for tting
the open circle data and (i/F = 1.16) for tting the solid circle data.
Curves were replotted from [128], J. Przyluski et al., Electrochim. Acta
40, 2102 (1995).
752 Polymer Electrolyte Nanocomposites
can be explained as follows. By increasing the fraction of
llers, the total amount of amorphous state around the llers
increases since the surface area increases, thus increasing
the conductivity. If the ller content is so high, some of the
ller agglomerates (making contact) so that the surface area
is reduced, thus reducing the fraction of amorphous state
around the ller, thereby reducing the conductivity. At a
specied amount of ller fraction, the continuous network of
amorphous state disappears so that the transport of cations
is blocked. The conductivity of composites can be approxi-
mated with the conductivity of ller particles.
The effective medium approximation was used to explain
the conductivity enhancement in polymer electrolyte com-
posites [128]. Dispersed ceramics create an amorphous layer
around the particles, which have a high conductivity. The
composite is considered to be a two-phase system: parti-
cles and amorphous layer as one phase and the rest of the
polymer as the other phase. The conductivity of the second
phase (polymer matrix) is equal to the conductivity of the
polymer electrolytes free of dispersed particles. The con-
ductivity of the particleamorphous layer can be calculated
using MaxwellGarnett equation [129],
u
c
= u
1
2u
1
2Y(u
2
u
1
)
2u
1
u
2
Y(u
2
u
1
)
(57)
where u
1
= conductivity of the interface layer, u
2
=
conductivity of dispersed grain, and Y = volume fraction of
ller in the composite, according to equation [128]
Y =
1
(1 i,F)
3
(58)
where i = thickness of the conducting layer, and F = radius
of ller.
The system is analogous to a system containing conduct-
ing particles dispersed in a polymer matrix in which the
conducting particles correspond to the particleamorphous
layer. The improved conductivity of grain and polymer
medium is calculated by the Nakamura [130] or Nan and
Smith [131, 132] equation,
u
a
c
= 2u
c
I
c
3 I
c
(59a)
u
e
a
= 2u
e
1 I
c
2 I
c
(59b)
where I
c
= I
2
,Y, and I
2
= volume fraction of ller in a
bulk electrolyte.
The dependence of composite conductivity on the load
fraction of dispersed particles and particle size can be cal-
culated using the effective medium theory [133]:
_
I
2
Y
_
u
a
c
u
n
u
n

c
(u
a
c
u
n
)

_
1
I
2
Y
_
u
a
e
u
n
u
n

c
(u
a
e
u
n
)
=0
(60)
where u
n
is the conductivity of the composite and
c
is the
continuous percolation threshold for the composite grain.
F
c
can be taken to be 0.28 on the basis of general perco-
lation theory [131, 132]. Lines in Figure 27 are the theo-
retical prediction for a system of PEO:LiClO
4
PAAM and
PEO:NaI O-Al
2
O
3
[128]. The data were collected at 25
o
C.
The theoretical curves were calculated using i,F = 0.6 for
solid line and i,F = 1.16 for dashed line.
However, this theory cannot explain the enhancement
of conductivity at high temperatures compared to polymer
electrolytes which are free of ller as shown in Figure 4. An
alternative approach has been used to explain conductivity
enhancement due to ller dispersion [134]. The presence of
a ller induces the formation of an amorphous layer around
the ller particles as well as the presence of an ionic layer
around the particles. Typically, an amorphous layer around
the ller particles is effectively created if the polymer con-
taining a ller is heated above the melting point and then
cooled down. The presence of a ceramic ller inhibits the
recrystallization of a polymer layer around the particles so
that layer remains in the amorphous state after the poly-
mer reaches room temperature. That means that, without
heating above the melting point, the conductivity of polymer
electrolytes containing a ller is close to that of ller-free
polymer electrolytes. However, by assuming the presence
of an ionic layer around the ller which has a conductivity
higher than that when the ionic layer is absent, the conduc-
tivity is then enhanced even though all parts of the polymer
are in the crystalline phase or in the amorphous phase. This
assumption explains the enhancement in conductivity at high
temperatures even when all parts of the polymer are in the
amorphous state.
This approach is similar to the observation of conductiv-
ity enhancement in solid-state ionic composites, into which
ceramic llers are dispersed [135]. The ionic transport in
solid-state ionics is very similar to polymer electrolytes. Poly-
mer electrolytes in the crystalline phase can be thought of
as analogous to solid-state ionics having a low conductiv-
ity, and polymer electrolytes in the amorphous phase can
be analogous to solid-state ionics having a high conductivity.
Therefore, since the ionic layer is assumed to be present in
solid-state ionic materials when the llers are dispersed, a
similar layer would be expected to appear in polymer elec-
trolytes containing a ceramic ller. The conductivity of the
ionic layer can be roughly approximated with u
l
= u
~
Au,
in which u
~
is the conductivity of the medium around
the particles in the absence of an ionic layer, and Au is
the excess conductivity due to the presence of a charged
layer. Thus, if the polymer around the ller is in the crys-
talline state, the conductivity of the layer around the parti-
cles would be higher than the conductivity of this crystalline
state. If the polymer medium around the particles is in the
amorphous state, the conductivity of the ionic layer would
be higher than the conductivity of the amorphous polymer.
Mikrajuddin et al. calculated the loading effect on the elec-
trical conductivity of composites by assuming a simple cubic
packing of ller particles and obtained a bell-shaped conduc-
tivity variation [134], similar to that obtained experimentally.
Most of llers used for making polymer electrolyte com-
posites are restricted to nonconductive materials. Recent
reports of the addition of small amount of conducting mate-
rials such as carbon (1.5 wt.%) with surface area of 60 m
2
/g
showed an enhancement of conductivity and interfacial sta-
bility compared to carbon-free composites [136].
Polymer Electrolyte Nanocomposites 753
11.2. Effect of Salt Concentration
on Conductivity
Another potential approach is to increase the concentra-
tion of salt ions in the matrix. However, this approach is
unpredictable. Figure 28 shows the effect of ion content
on the electrical conductivity of a system of PEG:LiClO
4

o-Al
2
O
3
at 25
o
C [137]. By increasing the concentration of
salt ions, the electrical conductivity initially increases, fol-
lowed by a decrease after passing a specied salt concen-
tration. Explanations have been proposed to explain this
observation. Doeff et al. explained this behavior in terms
of a trade-off between an increasing number of charge car-
riers and ion aggregation and increased viscosity due to
ionic cross-linking, which lowers the conductivity as the salt
concentration passes a critical value [138]. Dupon et al.
explained this dependence as being due to strong ion pair-
ing, which effectively traps the mobile cations, and therefore
signicantly reduces the ionic conductivity relative to non-
ion-paired complexes on a similar structure [139].
Marcinek [137] found that up to LiClO
4
concentra-
tion equal to 0.25 mol/kg, the conductivity of both sys-
tems PEG:LiClO
4
and PEG:LiClO
4
o-Al
2
O
3
system of
PEG:LiClO
4
o-Al
2
O
3
do not differ from each other by
more than 3040%. However, for high salt concentrations,
the conductivity of electrolytes containing o-Al
2
O
3
are much
higher. The difference rises up to 3040 times for a system
with the highest salt concentration.
11.3. Effect of Temperature on Conductivity
Since the transport of ions in the amorphous state is assisted
by chain segment relaxation, the temperature dependence
for ion transport in this state follows the temperature
dependence of chain relaxation. A VogellTammanFulcher
(VTF) dependence of electrical conductivity then results,
that is,
u(T ) =
.

T
exp
_

E
T T
t
_
(61)
0.001 0.01 0.1 1
10
6
10
6
10
6
Concentration [mol/kg]


[
S
/
c
m
]
Figure 28. Effect of salt concentration on the conductivity at 25
o
C
of PEG:LiClO
4
o-Al
2
O
3
. Data points were extracted from [137],
M. Marcinek et al., Solid State Ionics 136137, 1175 (2000).
in which ., E, and T
t
are constants and T is the tempera-
ture. The value of the constant parameter depends on the
type of composites as well as the conditions used in prepar-
ing the polymer. The value of T
t
depends on the glass tran-
sition of the polymer. A generally accepted relation is T
t
=
T
g
50, in which T
g
denotes the glass transition of the poly-
mer [140]. The glass transition of most polymers is related to
the melting point that satises T
g
/T
n
= 0.50.8 [19]. Param-
eter E can be expressed as a linear function of T
t
[134].
The value of VTF parameters was found to depend on
the concentration of salt. For example, in a system of
PPG4000:AgCF
3
SO
3
, the value of VTF parameters at var-
ious salt concentrations was reported by Eliasson et al.
[141]. The VTF parameters also depend on the salt type, as
reported by Florzanczyk et al. [142].
On the other hand, since the transport of ions in the
crystalline state is dominated by ion carriers jumping to the
nearby locations, the temperature dependence of conductiv-
ity then follows the Arrhenius law,
u(T ) = .exp
_

1
a
kT
_
(62)
in which 1
a
is the activation energy. The value of 1
a
/q,
in which q is the charge of the ion carrier, represents the
blocking potential that must be overcome for an ion carrier
to jump to the nearest locations. For polymer electrolytes
only (free of ceramic llers), a VTF behavior of conduc-
tivity is observed at temperatures above the melting point;
Arrhenius behavior prevails at low temperatures. A drastic
depression of electrical conductivity occurs at temperatures
where the VTF behavior changes to Arrhenius behavior [9].
Since the dispersion of ller particles will generate amor-
phous states even at low temperatures, VTF behavior is
observed even at low temperatures for conductivity, as can
also be seen in Figure 2. Continuing the VTF behavior up
to ambient temperature overcomes a drastic depression of
conductivity so that high conductivity prevails at low tem-
peratures.
11.4. Effect of Particle Surface
on Conductivity
As observed in Figure 4, at high temperatures where all
parts of the polymer are in an amorphous state, the con-
ductivity of a composite is higher than that of polymer
electrolytes which are free of llers. This indicates that
the amorphous layer around ller particles should not be
the only contributor to conductivity enhancement. It has
been proposed that the surface properties of the dispersed
ller also play a signicant role in the electrical conduc-
tivity of polymer electrolyte nanocomposites. Prior to use,
the llers are usually cleaned by drying at certain temper-
atures in vacuum or washing with a solution followed by
drying in vacuum to remove contaminants on the surface.
Matsuo and Kuwano observed an enhancement in conduc-
tivity by a factor of ve to ten when a lithium dodecylsul-
fate surfactant was added to the surface of an SiO
2
in a
system of PEG-LiCF
3
SO
3
-SiO
2
[43]. Prior to the synthesis
of composites, the silica nanoparticles were dispersed in a
754 Polymer Electrolyte Nanocomposites
solution of the surfactant to produce a self-assembly of sur-
factant molecules on the particle surface. Figure 29 shows
the effect of surfactant content (lithium dodecylsulfate, LDS
(C
12
H
25
OSO
3
Li)) on the electrical conductivity of a system
of poly(ethylene oxide)-LiCF
3
SO
3
-SiO
2
. The ionic conduc-
tivity rapidly increases with increasing LDS content and then
decreases after reaching a critical value at a certain weight
fraction of LDS. The weight fraction of LDS to produce the
conductivity peaks is the amount necessary for producing a
monolayer on the surface of SiO
2
. Further increases in LDS
content led to a reduction in conductivity.
Functionalization of SiO
2
in a poly(ethylene glycol)-based
polymer electrolyte using 2-[methoxy(polyethylenoxy)-
propyl] trimethoxy silane was reported to enhance electrical
conductivity as well as the tensile strength of compos-
ites (the tensile strength was increased by about 1 Mpa
compared to that using pristine SiO
2
) [109]. A small
improvement in conductivity compared to pristine SiO
2
was observed for small loading levels of surfactant. It was
assumed that silane moieties attached on the SiO
2
surface
effectively prevent coagulation of the SiO
2
particles during
blending because of steric repulsive forces. It then increases
the dispersibility of SiO
2
powder in the PEO matrix [109].
The relaxation of chain segments is also improved by adding
treated SiO
2
. In poly(ethylene oxide)-based electrolytes, Li

ions are coordinated to oxygen atoms in the polymer chains.


Movement of the dissociated Li

ions can be constrained


by multiple oxygen atoms coordinated to the same Li

central ion. Upon the addition of functionalized SiO


2
, the
oxygen atoms from the short polyether units on the surface
of SiO
2
compete with the oxygen atoms from the polymer
chains for complexion with Li

. The result is a more relaxed


coordination between oxygen atoms and Li

ions, which
in turn, facilitates the transport of Li

ions through the


polymer.
Croce et al. observed a difference in conductivity when
acidic, neutral, and basic Al
2
O
3
were used as llers [143].
A specic interaction between the surface group of ceramic
particles and both the PEO segment and lithium salt ions
0 10 20 30 40 50
5.5
5.0
4.5
L
o
g


[
S
/
c
m
]
LDS/(SiO
2
+LDS) wt%
T = 40 C
Figure 29. Effect of surfactant weight fraction on the electrical con-
ductivity of PEG:LiCF
3
SO
3
containing LDS coated SiO
2
nanoparticles.
Data points were extracted from [43], Y. Matsuo and J. Kuwano, Solid
State Ionics 79, 295 (1995).
was assumed to exist, which can be attributed to Lewis acid-
base interactions. This interaction may act as cross-linking
centers for the PEO segments and for the X

anions to
lower the PEO reorganization tendency and thus promoting
the structure modication of the polymer chains. The effect
promotes Li

conducting pathways at the ceramic surface.


Second, Lewis acid-base interactions between the electrolyte
ionic species lower ionic coupling. The expected result is
the promotion of salt dissociation via a type of ion-ceramic
complex formation. The surface of the ller also affects the
transference number of ion in the composites. Table 8 shows
the effect of acid-base properties of the ller on the transfer-
ence number in a system of P(EO)
20
:LiCF
3
SO
3
containing
Al
2
O
3
[143].
Capiglia et al. investigated the properties of composite
electrolytes consisting of SiO
2
powder with different surface
features in PEO:LiClO
4
system. They removed the hydroxyl
groups on SiO
2
by calcinations at 900
o
C in order to reduce
the hygroscopic properties of the particles. The reduction
of hydroxyl group on the particle surface was found to
increase the ionic conductivity [144]. Walls et al. prepared
series of SiO
2
material with hydrophilic and hydrophobic
surface groups and used them as ller in PEGDM system.
They showed clearly a cause and effect relationship between
the observed properties and the surface chemistry [145].
Fan et al. attached different surface group to the silica and
used the functionalized particles in PEGDM-based compos-
ite electrolyte. They found that the silica surface chemistry
affected mostly the rheological properties, but not the ionic
conductivity [146].
11.5. Effect of Particle Size on Conductivity
The use of nanometer- and micrometer-sized Al
2
O
3
in a sys-
tem of PEO:LiBF
4
improved the electrolyte mechanical and
electrochemical performance with decreasing size of ceramic
ller. Nanometer-sized particles provide an order increase in
conductivity compared to micrometer-sized ller [36]. Chan-
dra et al. observed a decrease in the conductivity at 333 K
of two orders of magnitude when particle size was increased
from 10 to 100 m. However, only a small decrease was
observed at 373 K when the particle size was increased in
the same range [147]. Wieczorek et al. reported a decrease
in conductivity when the particle size is increased from 2 to
7 m [148]. Figure 30 shows the Arrhenius plot of electrical
conductivity of a system of PEO-based polymer electrolyte
nanocomposites at different ller sizes [148].
The dependence of conductivity on the ller size can be
described qualitatively using a space-charge layer concept
Table 8. Lithium transference number at 95
o
C.
Samples i

P(EO)
20
LiCF
3
SO
3
0.46
P(EO)
20
LiCF
3
SO
3
10%Al
2
O
3
(basic) 0.48
P(EO)
20
LiCF
3
SO
3
10%Al
2
O
3
(neutral) 0.54
P(EO)
20
LiCF
3
SO
3
10%Al
2
O
3
(acidic) 0.63
Adapted with permission from [143], F. Croce et al.,
Electrochim. Acta 46, 2457 (2001). 2001, Elsevier.
Polymer Electrolyte Nanocomposites 755
2.83 3 3.19 3.41
3
4
5
6
7
8
1000/T [K
1
]
L
o
g


[
S
/
c
m
]
Figure 30. Arrhenius plot of (square) PEO:LiBF
4
, (diamond) PEO:
LiBF
4
containing 10 wt.% Al
2
O
3
(micrometer size), and (circle)
PEO:LiBF
4
containing 10 wt.% Al
2
O
3
(nanometer size). Data points
were extracted from [148], W. Wieczorek et al., Electrcochim. Acta 40,
2251 (1995).
that is commonly used to explain similar effect in a com-
posite of solid-state ionics. The origin of space-charge layer
is ascribed to the difference in the free energy formation
of the individual vacancies in case of Schottky disorder and
of the vacancies and the interstices in the case of Frenkel
disorder. The space-charge layer for a two-phase compos-
ite of solid-state ionics has been introduced by Maier, who
found the enhanced conductivity of the composite material
as [149]
u = (1 )u
t
(3

2)p
1
_

r
_
_
::
t
FT
I
n

:
_
A
:t
(63)
where u
t
= conductivity of ller-free polymer electrolytes,
= volume fraction of ller, p
1
= a constant to account for
the topology of ller, r = ller particle radius, : = dielectric
constant, :
t
= permittivity of vacuum, I
n
= bulk molar vol-
ume of the matrix phase,
:
= the mobility of dominant
defect, and A
:t
= the mole fraction of dominant defect at
the interface. This equation can t the data of conductivity
of NaCl containing various sizes of Al
2
O
3
[150].
11.6. Effect of Plasticizers on Conductivity
The plasticizers have a great inuence on the electrical,
chemical, and electrochemical properties of the polymer
electrolyte composites. The addition of ethylene carbonate
results in a more homogeneous and transparent lm. Golod-
nitsky et al. observed a conductivity enhancement by a fac-
tor of two in system of LiI:PEO
9
PMMA
0.5
6 vol.% Al
2
O
3
when ECl is added [151]. Enhancement about seven times
was observed in a system of LiI:PEO
6
PMMA
0.5
6 vol.%
Al
2
O
3
if PEG
0.5
is added as appears in Table 9 [151].
Leo et al. [152] observed the different temperature depen-
dence of conductivity when plasticizer is added into the
polymer composites. Different weight percent of plasticizer
alters the shape of log u against 1/T . The temperature
dependence of conductivity below 50
o
C shows an Arrhenius
Table 9. Effect of EC and PEG plasticizers on the electrical conductiv-
ity at 120
o
C.
Conductivity
Electrolytes Plasticizer Filler (S/cm)
PEO
9
PMMA
0.5
:Lil Al
2
O
3
6 wt.% 0.610
3
PEO
9
PMMA
0.5
:Lil ECl Al
2
O
3
6 wt.% 1.210
3
PEO
6
PMMA
0.6
:Lil Al
2
O
3
6 wt.% 0.410
3
PEO
6
PMMA
0.6
:Lil PEG
0.5
Al
2
O
3
6 wt.% 1.010
3
Adapted with permission from [151], D. Golodnitsky et al., Solid State Ionics
85, 231 (1996). 1996, Elsevier.
type for all plasticizer content. However, above 50
o
C, the
temperature dependence of conductivity seems to change
with the change in the content of plasticizer. For 5 wt.%
plasticizer (ethylene carbonate), there is a temperature T
n
above which the Arrhenius plot of conductivity was observed
and below which a qualitative change in the curve was
observed. The discontinuity in the slope of the curve of
the unplasticized composite disappears. At higher plasticizer
content, a distinct change in the temperature-dependent
conductivity was observed.
11.7. Effect of Storage Time on Conductivity
Shin et al. measured the effect of storage time on the
electrical conductivity of polymer electrolyte composites of
PEO
10
:LiCF
3
SO
3
containing 515 wt.% of Ti
n
O
2n1
stored
at 90
o
C. The conductivity increased with increase in the
storage time until approximately 5 h, when the conductiv-
ity reached a steady-state value. The conductivity of poly-
mer electrolytes containing ceramics reaches the steady-state
value faster than ceramic-free polymer electrolytes [153].
The conductivity of PEO
8
:LiClO
4
containing 10 wt.% SiO
2
is relatively stable up to two weeks storage as displayed in
Figure 31 [154]. Croce et al. suggested that the recrystalliza-
tion may occur with kinetics which appear to be critically
dependent on the annealing condition [9].
0 4 6 8 10 12 14
Time [days]
1.00
1.25
1.50
1.75
2.00
2.25
2.50

1
0

5
S
/
c
m
]
2
Figure 31. Effect of storage time on the conductivity of PEO
8
:LiClO
4
containing 10 wt.% SiO
2
. Data points were extracted from [154],
G. Appetecchi et al., Electrochim. Acta 45, 1481 (2000).
756 Polymer Electrolyte Nanocomposites
Appetecchi et al. compared the conductivity of EC-
DMC-PAN:LiPF
6
containing 6 wt.% Al
2
O
3
and EC-DMC-
PAN:LiPF
6
free of Al
2
O
3
. They found that, after 30 days
aging at 75
o
C, there was no signicant change in the con-
ductivity of ceramics containing polymer electrolytes, while
the ceramic-free polymer suffered a depression in conduc-
tivity to become about 10% of the initial value [155].
12. MECHANICAL PROPERTIES
12.1. Effect of Filler Content
Weston and Steele used ceramic ller to improve the
mechanical properties of PEO lms [156]. Addition of
ceramic ller to improve the mechanical properties of poly-
mer is not new. The addition of carbon black ller can
extend the lifetime of tires from 5,000 miles if no car-
bon black is used to a potential 80,000 miles in some cur-
rent tires. Figure 32 shows the effect of ller content on
the elastic modulus of a system of poly(ethylene glycol)
dimethyl ether:LiN(CF
3
SO
2
)
2
containing fumed silica with
a surface group modied to become Si-C
8
H
17
instead of
Si-OH. Increasing the concentration of the ller generates
a stronger network structure with the modulus increasing by
two orders of magnitude [157].
Although addition of ller is effective in hardening the
polymer electrolytes at lower temperature, the effective-
ness becomes lower at temperatures higher than the melting
point of the matrix. Wen et al. used a whisker to improve
the mechanical strength of polymer electrolyte compos-
ites. The work is based on the morphology of whisker
which has a special shape similar to rigid network formed
inside the polymer electrolytes. The mechanical strength
of the polymers was evaluated with their thermal creep
behavior, indicated by the relative resistance change of the
polymer electrolyte lms after being kept for some time
under a certain pressure [113]. Addition of ceramic ller
in a composite of PEO
8
:LiClO
4
remarkably affected their
0 10 15 20 25
10
3
10
4
10
5
10
6
Fumed silica weight %
E
l
a
s
t
i
c
m
o
d
u
l
u
s
[
P
a
]
5
Figure 32. Effect of weight fraction of ller on the elastic modulus of
PEG dimethyl ether:LiN(CF
3
SO
2
)
2
containing C
8
H
17
coated fumed sil-
ica. Data points were extracted from [152], C. J. Leo et al., Solid State
Ionics 148, 159 (2002).
thermal creep properties, in particular at high tempera-
tures. The electrolytes containing whisker ller are much
more mechanically stable than those containing nanosized
-Al
2
O
3
. However, at low temperatures (below the melting
point), the thermal creep properties of both composites are
similar.
Leo et al. observed improvement of mechanical strength
of PEO-based polymer electrolytes by a factor of three when
7.5 wt.% Nasicon glass-ceramic llers is added [152]. This
ller content also increased the glass temperature of about
14
o
C. It is suggested that the addition of ller increases
the stiffness of polymer segment, thereby suppressing the
polymer chain motion [158].
12.2. Effect of Particle Surface
Inorganic llers are actually bonded to the molecular chain
and thereby immobilize the polymer chain. The degree
of adhesion between the polymer matrix and the llers,
the surface area of the ller, and the packing charac-
teristic of the ller are important factors that determine
the mechanical characteristics of the composites [159].
Liu et al. compared the effect of tensile strength of
composites of PEO:LiBF
4
+SiO
2
containing pristine SiO
2
and organic coated SiO
2
[109]. The SiO
2
was coated
using 2-[methoxy(polyethylenoxy)-propyl] trimethoxy silane.
Figure 33 shows the effect of ller loading on the tensile
strength. It can be seen that the trend in the tensile strength
variation is similar for both composites. However, the ten-
sile strength of composite made using coated SiO
2
ller is
always larger than that of material made by pristine SiO
2
.
A double enhancement was observed in all ller loading
region. Functionalized SiO
2
might be considered as a ller
which carries a pendant coupling agent. The coupling agent
has two types of functional groups, one capable of forming
chemical bond with the surface of the ller and the other
capable of entangling with PEO chains [160].
4 8 12 16 20
0
1
2
3
PEO + pristine SiO
2
PEO + funcionalized SiO
2
PEO
wt.% SiO
2
T
e
n
s
i
l
e
s
t
r
e
n
g
t
h
[
M
p
a
]
Figure 33. Effect of ller content on the tensile strength of PEO:LiBF
4
containing pristine SiO
2
(open circle) and functionalized SiO
2
(solid
circle). Data points were extracted from [109], Y. Liu et al., J. Power
Sources 109, 507 (2002), and lines are for eye guide.
Polymer Electrolyte Nanocomposites 757
13. THERMAL PROPERTIES
13.1. Effect of Filler Content
on the Glass Temperature
Franco et al. observed the reduction in the glass transition
temperature in composite of PEO:LiClO
4
containing carbon
black (conducting ller) [97]. T
g
= 220 K was observed for
composite containing 10 wt.% carbon black, 210 K for that
containing 20 wt.% carbon black, and 202 K for that con-
taining 30 wt.% carbon black. Choi and Shin measured the
effect of the ller loading on the glass temperature of a sys-
tem of PEO
16
LiClO
4
SiC. T
g
rst decreases rapidly with
the increase in the ller content and then reaches a plateau
beyond which it further drops as can be seen in Figure 34
[161].
On the contrary, Chung and Sohn observed also an
enhancement of glass temperature as the concentration of
salt increases in a system of polymer comb-shaped polymer
matrix by attaching triethoxyethylene side chain on the main
chain of polyethylene oxide [162]. Wieczorek et al. observed
that the SiO
2
ller which enhanced the conductivity raised
the T
g
[163], while the o-Al
2
O
3
ller which reduced the con-
ductivity above room temperature lowered the T
g
[164]. The
data indicated that the variation of T
g
neither exhibits sys-
tematic results nor meets the expected variation of conduc-
tivity.
Filler content also affects the melting enthalpy of com-
posites. Figure 35 shows the effect of ller loading on the
melting enthalply extracted from DCS curve for system of
PEO
16
:LiClO
4
+SiC, and the normalization relative to the
mass fraction of pure (PEO)
16
:LiClO
4
[161]. Using Eq. (54)
and remembering AH
n, c
is proportional to mass fraction of
pure (PEO)
16
:LiClO
4
, curve (b) reects the crystallinity of
the composite. It is clear that the crystallinity increases with
the loading of the ller at low loading and then decreases
with further increase in the ller content. Choi and Shin
proposed the enhancement of crystallinity with the loading
content of ller is due to the role of the ller which may
act as nucleation centers of crystalline polymer phase. The
0 20 40 60
50
45
40
35
30
25
SiC wt.%
T
g
[

C
]
Figure 34. Effect of ller content on the glass temperature of PEO
16
:
LiClO
4
containing SiC. Data points were extracted from [161], B. Choi
and K. Shin, Solid State Ionics 8688, 303 (1996), and line is for eye
guide.
0 10 20 30 40 50 60 70
SiC wt.%
20
30
40
50
60
70
80

H
m
[
J
/
g
]
Figure 35. Effect of ller content on the melting enthalpy (solid circle)
and crystallinity (open circle) of PEO
16
:LiClO
4
containing SiC. Data
points were extracted from [161], B. Choi and K. Shin, Solid State Ionics
8688, 303 (1996).
nucleation effect was considered to be sufcient to overcome
the hindrance of crystallization due to the enhancement of
the segmental motion of the PEO [161].
13.2. Effect of Salt Concentration
on the Glass Temperature
Marcinek et al. observed the effect of salt concentration
of the system of PEG:LiClO
4
o-Al
2
O
3
and found that
the glass temperature also increases with the salt con-
centration as appears in Figure 36 [137]. For concen-
trations up to 0.25 mol/kg salt concentration, the T
g
is
almost constant. At higher salt concentrations, the plot
of T
g
with respect to the logarithm of salt concentra-
tion likely increases with an exponential trend. They also
compared the variation of glass temperature of system
free of ller and found that at up to 0.26 mol/kg salt
concentration, the T
g
of both samples is almost similar.
10
6
10
4
10
2
10
0
190
200
210
220
230
Salt cocentration [mol/kg]
T
g
[
K
]
Figure 36. Effect of salt concentration on the glass temperature of
PEG:LiClO
4
. containing o-Al
2
O
3
. Data points were extracted from
[137], M. Marcinek et al., Solid State Ionics 136137, 1175 (2000).
758 Polymer Electrolyte Nanocomposites
At higher salt concentration, the T
g
of ller-free samples
is higher that that of sample containing ller. Sun et al.
found that the glass temperature of a system of poly(A,A-
dimethylaminoethyl methacrylate)-tetraglyme:LiClO
4
poly-
mer electrolytes increases strongly with the increasing salt
concentration over the rst 0.5 mol/kg addition of LiClO
4
.
Further addition of salt results in only a minor increase in
the glass temperature [165].
14. LUMINESCENT COMPOSITES
Most llers used in polymer electrolyte nanocomposites are
optically inactive, that is, they do not have any lumi-
nescence properties. It would be interesting to investigate
the properties of composite if the llers loaded in polymer
emit luminescence. This approach would have interesting
applications such as the possibility of obtaining information
regarding the degradation of battery or fuel cells using such
composites, for example based on the luminescence intensity
emitted by the ller. Changes in properties of the polymer
when the battery or fuel cell degrades from the initial per-
formance would be expected. This indicates that the proper-
ties of the medium around the ller would change when the
battery or fuel cell degrades. It would affect the detected
luminescence intensity emitted by the ller change. Based
on this detected luminescence, the current performance of
the battery or fuel cell could be determined.
Zinc oxide (ZnO) is one of the promising materials for
preparing composite polymer electrolytes that are lumines-
cent. ZnO nanoparticles play a role as an agent for reducing
the tendency of the polymer matrix to crystallize and simul-
taneously as luminescence centers. Since the luminescence
spectra of ZnO are dependent on crystalline size (quantum
size effect), the emitted wavelength can be controlled by
adjusting the size of the ZnO nanocrystallites. ZnO particles
of different sizes can be produced easily using, for example,
a colloidal process, so that the color of the composite can
be easily tuned. A blue composite can be prepared using
smaller ZnO nanoparticles, while a yellow composite can
be produced using large ZnO particle size. This composite
has also considerable potential for fabricating devices that
simultaneously produce both electrical current and light,
such as luminescent electrochemical cells, self-powered dis-
plays, etc.
Figure 37 (right) shows an example of a photolumines-
cence (PL) spectrum of composite polymer electrolytes con-
taining ZnO nanoparticles [62]. The sample produced a
highly intense green luminescence. The color of those sam-
ples can clearly be observed by the naked eye even using a
hand-held UV lamp as background illumination in the lab-
oratory. The PL peaks can be shifted to lower or higher
wavelengths by altering the concentration of the precursors.
The luminescence spectrum was produced by an electron
transition from the bottom of conduction band to a state
located near the center of the band.
The corresponding excitation spectra of the samples are
shown in Figure 37 (left). A peak is related to the band-
edge transition. The wavelength peak is smaller than that
observed for bulk ZnO, which shows a peak at 365 nm
(energy about 3.4 eV). The shift of the band-edge peak to
lower wavelengths indicates the presence of a quantum size
300 400 500 600
Wavelength [nm]
Excitation Luminescence
I
n
t
e
n
s
i
t
y
[
-
]
Figure 37. Left: Excitation spectra of PEG:Li containing ZnO nano-
particles. Right: The corresponding luminescence spectra. Data were
extracted from [62], Mikrajuddin et al., J. Electrochem. Soc. 149, H107
(2002).
effect of the optical bandgap. Using the size-dependent opti-
cal bandgap equation, the size of ZnO nanoparticles in the
matrix is estimated to be around 6 nm.
The in-situ growth of nanoparticles in the polymer matrix
ensures the dispersion of nanoparticles in polymer bulk
easily without the need for an intensive mixing process.
The selection of a precursor containing lithium ions per-
mits the insertion of cations during the growth process.
This approach has been used to produce composite poly-
mer electrolytes containing ZnO nanoparticles using lithium
hydroxide as one precursor and a hygroscopic solution of
zinc acetate as another precursor. A high molecular weight
polyethylene glycol was rst mixed with a solution of lithium
hydroxide and then further mixed with a hygroscopic solu-
tion of zinc acetate. Since lithium did not participate in the
formation of zinc oxide, lithium ions were left in the polymer
matrix and were then able to take part as ion carriers.
15. CONCLUSION
Stable and high electrical conductivity polymer electrolytes
are required for fabricating exible and environmentally
friendly fuel cells and batteries. Some important parameters
critical to polymer electrolytes in order to bring this material
into industries have been discussed in this review. Although
high electrical conductivity of polymer electrolytes usually
appears at high temperatures, many efforts to improve the
electrical conductivity at room temperature have been intro-
duced by researchers all around the world. They include
the synthesis of amorphous polymer matrix, addition of sec-
ond phase in the host polymer matrix (plasticizers) like low
molecular weight polymers, and addition of size chains. The
most popular approach now is the addition of ceramic ller
in nanometer scale. This approach becomes interesting since
the addition of nanometer-sized ceramic ller improves not
only the electrical conductivity of polymer electrolytes but
also the mechanical properties. This material, usually named
Polymer Electrolyte Nanocomposites 759
as polymer electrolyte nanocomposite, attracts a lot of atten-
tion because of the simplicity in the processing. Various
kinds of polymer in a lot of molecular weights as well as
various kinds of ceramic ller in a lot of sizes are available
in the market. The process of production of composites is
relatively simple, that is, by just mixing the polymer, salt,
and ller in a common solution, ended with drying. Methods
of preparation of polymer electrolyte nanocomposites were
briey discussed in this review.
Characterizations of the prepared lms are important to
understand the properties as well as to nd the optimum
preparation conditions. Various methods of characterization,
including electrical, spectroscopic, microscopic, and thermal
characterizations, were discussed briey. Detailed explana-
tions of the methods might be found in some analytical
chemistry books.
In the nal part, we discussed some parameters which
affect the properties of polymer electrolyte composites. It
was found that the properties of the composites are sensi-
tive to the preparation condition as well as material used for
making the composites. For example, the electrical conduc-
tivity is greatly dependent on the volume fraction of llers,
surface property, ller size, type of salt, side chain, etc. Also
the mechanical properties are affected by ller content, sur-
face property, and salt concentration.
GLOSSARY
Anion An ionic species having a negative charge.
Battery A device that stores energy and makes it available
in an electrical form.
Blending Physically mixing several types of materials so
thoroughly that they appear to be indistinguishable from
each other in the product.
Casting method A method of forming sheets of composite
materials by pouring uid materials onto a at surface.
Cation An ionic species having a positive charge.
Charging Supply electric current to a battery.
Copolymerization Polymerization with two or more differ-
ent monomers.
Cycle life The number of charge/discharge cycles that are
possible before failure occurs.
Differential scanning calorimetry A materials characteri-
zation laboratory technique by which the temperature of a
sample of the substance in question is raised in increments
while a reference is heated in the same rate. The amount
of heat that is required to heat the sample and the refer-
ence to each temperature is recorded and can be plotted.
From these plots, melting points, phase change tempera-
tures, chemical reaction temperatures, and glass transition
temperature of polymers can be determined.
Discharging Withdrawing charge from a battery.
Doctor blade technique A technique using a at bar used
for regulating the amount of liquid material on the rollers of
a coating machine, or to control the thickness of a coating
after it has been applied to a substrate.
Electrical conductivity A measure of how well a material
accommodates the transport of electric charge.
Energy density The energy obtainable per unit volume.
Energy efciency (energy released on discharge)/(energy
required for charge).
Filler Inorganic particles added to polymer matrix.
Fourier transform spectroscopy 1A measurement tech-
nique whereby spectra are collected based on the response
from a pulse of electromagnetic radiation. Fourier transform
spectroscopy is more sensitive and has a much shorter sam-
pling time than conventional spectroscopic techniques.
Free volume The extrapolated differences in volume
between a glass and the extrapolation of the melt curve is
called the free volume. The free volume is associated with
the space between molecules in a sample.
Glass transition temperature (T
g
) The temperature below
which molecules have very little mobility. On a larger scale,
polymers are rigid and brittle below their glass transition
temperature and elastic above it. T
g
is usually applicable to
amorphous phases and is commonly applicable to glasses
and plastics.
Hot press The forming of a compact material at tempera-
tures sufciently high to cause concurrent sintering.
Melting point The temperature at which it changes state
from solid to liquid. When considered as the temperature
of the reverse change, the temperature is referred to as the
freezing point.
Nuclear magnetic resonance (NMR) A physical phe-
nomenon described independently by Felix Bloch and
Edward Mills Purcell in 1946. It involves the interaction
of atomic nuclei placed in an external magnetic eld with
an applied electromagnetic eld oscillating at a particular
frequency. Magnetic conditions within the material are
measured by monitoring the radiation absorbed and emitted
by the atomic nuclei. NMR is used as a spectroscopy tech-
nique to obtain physical, chemical, and electronic properties
of molecules. It is also the underlying principle of Magnetic
Resonance Imaging. NMR is one of the techniques used to
build quantum computers.
Number average molecular weight A way of determining
the molecular weight of a polymer. Polymer molecules, even
ones of the same type, come in different sizes (chain lengths,
for linear polymers), so we have to take an average of some
kind. The number average molecular weight is the common
average of the molecular weights of the individual polymers.
It is determined by measuring the molecular weight of n
polymer molecules, summing the weights, and dividing by n.
The number average molecular weight of a polymer can be
determined by osmometry, end-group titration, and colliga-
tive properties.
Plasticizers Materials added to polymer matrix to improve
the fraction of amorphous state.
Polymer electrolytes Electrolytes which using polymer as a
media to dissociate ions.
Rechargeable batteries Batteries that can be restored to
full charge by the application of electricity. They come in
many different designs using different chemistry.
Small angle x-ray scattering A laboratory technique in
which photons are elastically scattered from a sample. The
sample can not be too thick. Liquids can be examined. Fea-
tures on the nanometer length scale can be examined.
760 Polymer Electrolyte Nanocomposites
Specic energy The energy obtainable per unit weight.
Weight average molecular weight A way of determining
the molecular weight of a polymer. Polymer molecules, even
if of the same type, come in different sizes (chain lengths,
for linear polymers), so we have to take an average of some
kind. For the weight average molecular weight, this is done
as follows: weigh a number of polymer molecules, add the
squares of these weights, and then divide by the total weight
of the molecules. Intuitively, if the weight average molecu-
lar weight is u, and you pick a random monomer, then the
polymer it belongs to will have a weight of u on average.
The weight average molecular weight can be determined by
light scattering, small angle neutron scattering (SANS), and
by sedimentation velocity.
Wide angle x-ray scattering An X-ray diffraction tech-
nique that is often used to determine the crystalline struc-
ture of polymers.
ACKNOWLEDGMENT
Japan Society for the Promotion of Science (JSPS) Post-
doctoral Fellowship for Mikrajuddin Abdullah is gratefully
acknowledged.
REFERENCES
1. J. O. Besenhard and M. Winter, ChemPhysChem 3, 155 (2002).
2. J.-M. Tarascon and M. Armand, Nature 414, 359 (2001).
3. F. B. Dias, L. Plomp, and J. B. J. Veldhuis, J. Power Sources 88,
169 (2000).
4. M. Gauthier, M. Armand, and L. Krause, Proceedings of the 7th
International Meeting on Lithium Batteries, 1994, abstract p. 177.
5. M. Liu, S. J. Visco, and L. C. De Jonghe, J. Electrochem. Soc. 138,
1891 (1991).
6. H. Oman, Proceedings of the 12th Annual Battery Conference
on Applications and Advances, 1997, p. 31.
7. T. Turrentine and K. Kurani, Report UCD-ITS-RR-95-5, Institute
of Transportation Studies, 1995.
8. P. Lightfoot, M. A. Metha, and P. G. Bruce, Science 262, 883
(1993).
9. F. Croce, G. B. Appetecchi, L. Persi, and B. Scrosati, Nature 394,
456 (1998).
10. D. E. Fenton, J. E. Parker, and P. V. Wright, Polymer 14, 589
(1973).
11. P. V. Wright, Br. Polym. J. 7, 319 (1975).
12. M. B. Armand, J. M. Chabagno, and M. J. Duclot, in Fast Ion
Transport in Solids (M. J. Duclot, P. Vashishta, J. M. Mundy, and
G. K. Shenoi, Eds.), Elsevier, North Holland, Amsterdam, 1979.
13. F. B. Dias, L. Plomp, and J. B. J. Feldhuis, J. Power Sources 88,
169 (2000).
14. J. R. MacCalum, M. J. Smith, and C. A. Vincent, Solid State Ionics
11, 307 (1981).
15. M. Watanabe, S. Nagano, K. Sanui, and N. Ogata, J. Power Sources
35, 327 (1987).
16. Y. Song, X. Peng, Y. Lin, B. Wang, and D. Chen, Solid State Ionics
76, 35 (1995).
17. O. Inganas, Br. Polym. J. 20, 233 (1988).
18. D. W. Kim, J. K. Park, and H. W. Rhee, Solid State Ionics 83, 49
(1996).
19. R. J. Young and P. A. Lovell, Introduction to Polymers, 2nd ed.,
p. 297. Chapman & Hall, London, 1991.
20. J.-S. Chung and H.-J. Sohn, J. Power Sources 112, 671 (2002).
21. A. Nishimoto, M. Watanabe, Y. Ikeda, and S. Kohjiya, Elec-
trochim. Acta 43, 1177 (1998).
22. Y. Ikeda, Y. Wada, Y. Mataba, S. Murakami, and S. Kohjiya, Elec-
trochim. Acta 45, 1167 (2000).
23. K. Motogami, M. Kono, S. Mori, M. Watanabe, and N. Ogata,
Electrochim. Acta 37, 1725 (1992).
24. Y. Ikeda, H. Masuo, S. Syoji, T. Sakashita, Y. Matoba, and
S. Kohjiya, Polym. Int. 43, 269 (1997).
25. M. Watanabe, S. Yamada, K. Sanui, and N. Ogata, J. Chem. Soc.
Chem. Commun. 929 (1993).
26. M. Watanabe, S. Yamada, and N. Ogata, Electrochim. Acta 40,
2285 (1995).
27. Y. Tsuda, T. Nohira, Y. Nakamori, K. Matsumoto, R. Hagiwara,
and Y. Ito, Solid State Ionics 149, 295 (2002).
28. G. Feullade and P. Perche, J. Appl. Electrochem. 5, 63 (1975).
29. D. R. MacFarlene, J. Sun, P. Meakin, P. Fasoulopoulos, J. Hey,
and M. Forsyth, Electrochim. Acta 40, 2131 (1995).
30. S. Chintapalli and R. French, Solid State Ionics 8688, 341 (1996).
31. H. W. Rhee, W. I. Jung, M. K. Song, S. Y. Oh, and J. W. Choi,
Mol. Cryst. Liq. Cryst. Sci. Technol. A 294, 225 (1997).
32. M. Morita et al., J. Electrochem. Soc. 137, 3401 (1990).
33. D. W. Xia, D. Soltz, and J. Smidt, Solid State Ionics 14, 221 (1984).
34. M. S. Michael, M. M. E. Jacob, S. R. S. Prabaharan, and S. Rad-
hakrishna, Solid State Ionics 98, 167 (1997).
35. J. E. Weston and B. C. H. Steele, Solid State Ionics 7, 75 (1982).
36. W. Krawiec, L. G. Scanlon, Jr., J. P. Fellner, R. A. Vaia, S. Vasude-
van, and E. P. Gianellis, J. Power Sources 54, 310 (1995).
37. B. K. Choi and K. H. Shin, Solid State Ionics 8688, 303 (1996).
38. K. Nairn, M. Forsyth, H. Every, M. Greville, and D. R. MacFar-
lane, Solid State Ionics 8688, 589 (1996).
39. G. K. R. Sanadeera, M. A. Careem, S. Skaarup, and K. West, Solid
State Ionics 85, 37 (1996).
40. W. Wieczorek, J. R. Stevens, and Z. Florzanczyc, Solid State Ionics
85, 67 (1996).
41. D. Golodnisky, G. Ardel, and E. Peled, Solid State Ionics 85, 231
(1996).
42. E. Peled, D. Golodnisky, G. Ardel, and V. Eshkenazy, Electrochim.
Acta 40, 2197 (1995).
43. Y. Matsuo and J. Kuwano, Solid State Ionics 79, 295 (1995).
44. M. Slane and M. Solomon, J. Power Sources 55, 7 (1995).
45. Y. W. Kim, W. Lee, and B. K. Choi, Electrochim. Acta 45, 1473
(2000).
46. J. M. Pernaut and A. L. deOlieveira, Synth. Met. 84, 443 (1997).
47. M. C. Borghini, M. Mastragostino, and A. Zanelli, J. Power Sources
68, 52 (1997).
48. G. Ardel, G. Golodnitsky, E. Stauss, and E. Peled, J. Power
Sources, in press.
49. S. S. Sekhon, G. S. Sandhar, S. A. Agnihotri, and S. Chandra, Bull.
Electrochem. 12, 415 (1996).
50. S. S. Sekhon and A. Singh, Bull. Electrochem. 12, 671 (1996).
51. S. S. Sekhon and G. S. Sandhar, Eur. Polym. J. 34, 435 (1998).
52. S. A. Hashmi, A. K. Thakur, and H. M. Upadhyaya, Eur. Polym.
J. 34, 1277 (1998).
53. H. W. Chen and F. C. Chang, Polymer 42, 9763 (2001).
54. A. Chandra, P. K. Singh, and S. Chandra, Solid State Ionics,
154155, 15 (2002).
55. G. Katsaros, T. Stergiopolous, I. M. Arabatzis, K. G. Papago-
dokostaki, and P. Falaras, J. Photochem. Photobiol. A 149, 191
(2002).
56. B. ORegan and M. Gratzel, Nature 353, 737 (1991).
57. M. K. Nazeeruddin, A. Kay, I. Rodicio, R. Humphry-Baker,
E. Muller, P. Liska, N. Vlachopoulos, and M. Gratzel, J. Am.
Chem. Soc. 115, 6382 (1993).
58. M. K. Nazeeruddin, A. Kay, I. Rodicio, R. Humphry-Baker,
E. Muller, P. Liska, N. Vlachopoulos, and M. Gratzel, J. Am.
Chem. Soc. 115, 6382 (1993).
59. A. F. Nogueira, J. R. Durrant, and M.-A. De Paoli, Adv. Mater. 11,
826 (2001).
Polymer Electrolyte Nanocomposites 761
60. G. Ardel, D. Golodnitsky, K. Freedman, E. Peled, G. B. Appetec-
chi, P. Romagnoli, and B. Scrosati, J. Power Sources 110, 152
(2002).
61. C. Capiglia, J. Yang, N. Imanishi, A. Hirano, Y. Takeda, and
O. Yamamoto, J. Power Sources (in press).
62. Mikrajuddin, I. W. Lenggoro, F. G. Shi, and K. Okuyama, J. Elec-
trochem. Soc. 149, H107 (2002).
63. Mikrajuddin, I. W. Lenggoro, F. G. Shi, and K. Okuyama, J. Phys.
Chem. B, 107, 1957 (2003).
64. J. B. Goodenough, Solid State Ionics 69, 184 (1994).
65. K. M. Abraham and M. Alamgir, Chem. Mater. 3, 339 (1991).
66. M. K. Song, J. Y. Cho, B. W. Cho, and H. W. Rhee, J. Power
Sources 110, 209 (2002).
67. C. J. Leo, G. V. S. Rao, and B. V. R. Chowdary, Solid State Ionics
148, 159 (2002).
68. D. P. Almond and A. R. West, Solid State Ionics 23, 27 (1987).
69. M. Siekierski, W. Wieczorek, and J. Przyluski, Electrochim. Acta
43, 1339 (1998).
70. C. Kittel, Introduction to Solid State Physics, 7th ed., p. 310.
John Wiley, New York, 1996.
71. Y. Ma, M. Doyle, T. Fuller, M. M. Doeff, L. C. DeJonghe, and
J. Newman, J. Electrochem. Soc. 142, 1859 (1995).
72. S. Strauss, D. Golodnisky, G. Ardel, and E. Peled, Electrochim.
Acta 43, 1315 (1998).
73. J. W. Lorimer, J. Power Sources 26, 491 (1989).
74. P. R. Sorensen and T. Jacobson, Electrochim. Acta 27, 1671 (1982).
75. K. Kiukkola and C. Wagner, J. Electrochem. Soc. 104, 308 (1957);
104, 379 (1957).
76. J. Evans, C. A. Vincent, and P. G. Bruce, Polymer 28, 2324 (1987).
77. M. Doyle and J. Newman, J. Electrochem. Soc. 142, 3465 (1995).
78. P. G. Bruce, M. T. Hardgrave, and C. A. Vincent, Electrochim. Acta
37, 1517 (1992).
79. F. Croce, L. Persi, F. Ronci, and B. Scrosati, Solid State Ionics 135,
47 (2000).
80. S. Chung, Y. Wang, L. Persi, F. Croce, S. G. Greenbaum,
B. Scrosati, and E. Plichta, J. Power Sources 9798, 644 (2001).
81. S. D. Brown, S. G. Greenbaum, M. G. McLin, M. C. Wintersgill,
and J. J. Fontanella, Solid State Ionics 67, 257 (1994).
82. W. Gang, J. Roos, D. Brinkmann, F. Capuano, F. Croce, and
B. Scrosati, Solid State Ionics 5356, 1102 (1992).
83. Y. Dai, Y. Wang, S. G. Greenbaum, S. A. Bajue, D. Golodnitsky,
G. Ardel, E. Strauss, and E. Peled, Electrochim. Acta 43, 1557
(1998).
84. S. Chandra, Superionic Solids. North-Holland, Amsterdam,
1981.
85. P. A. Allen, in International Review of SciencePhysical Chem-
istry (C. A. McDowell, Ed.), Ser. 1, Vol. 4. Butterworths, London
and Boston, 1972.
86. G. E. Pake, in Advances in Solid State Physics (F. Seitz and
D. Turnbull, Eds.), Vol. 2, Academic Press, New York, 1956.
87. A. Abragam, The Principles of Nuclear Magnetic Resonance.
Oxford Univ. Press, London, 1961.
88. J. R. Hendrickson and P. J. Bray, J. Magn. Reson. 9, 341 (1973).
89. N. Bloembergen, E. M. Purcell, and R. V. Pound, Phys. Rev. 73,
679 (1948).
90. P. Manseld, Progress in NMR Spectroscopy, Vol. 8. Pergamon,
Oxford, 1971.
91. S. Wong, R. A. Vaia, E. P. Giannelis, and D. B. Zax, Solid State
Ionics 8688, 547 (1996).
92. S. Wong and D. B. Zax, Electrochim. Acta 42, 3513 (1997).
93. F. Croce, G. B. Appetecchi, S. Slane, M. Salomon, M. Tavarez,
S. Arumugam, Y. Wang, and S. G. Greenbaum, Solid State Ionics
8688, 307 (1996).
94. C. G. Joo, L. M. Bronstein, R. L. Karlinsey, and J. W. Zwanzider,
Solid State Nucl. Magn. Res. 22, 235 (2002).
95. T. J. Singh, T. Mimani, K. C. Patil, and S. V. Bhat, Solid State
Ionics, 154155, 21 (2002).
96. A. C. Bloise, C. C. Tambelli, R. W. A. Franco, J. P. Donoso, C. J.
Magon, M. F. Souza, A. V. Rosario, and E. C. Pereira, Electrochim.
Acta 46, 1571 (2001).
97. R. W. A. Franco et al., Solid State Ionics 136137, 1181 (2000).
98. S. D. Brown, S. G. Greenbaum, M. G. McLin, M. C. Wintergill,
and J. J. Fontanella, Solid State Ionics 67, 257 (1994).
99. M. Forsyth, D. R. MacFarlene, A. Best, J. Adebahr, P. Jacobson,
and A. J. Hill, Solid State Ionics 147, 203 (2002).
100. P. W. Atkins and R. S. Friedman, Molecular Quantum Mechan-
ics. Oxford Univ. Press, London, 1997.
101. W. Huang, R. French, and R. A. Wheeler, J. Phys. Chem. 98, 100
(1994).
102. A. Ferry, G. Oradd, and P. Jacobsson, Electrochim. Acta. 43, 1471
(1998).
103. H. Ericson, B. Mattsson, L. M. Torell, H. Rinne, and F. Sundholm,
Electrochim. Acta 43, 1401 (1998).
104. C. H. Kim, J. K. Park, S. I. Moon, and M. S. Yoon, Electrochim.
Acta 43, 1421 (1998).
105. A. Beronson and J. Lindgre, Solid State Ionics 60, 37 (1993).
106. A. Borkowska, A. Reda, A. Zalewska, and W. Wieczorek, Elec-
trochim. Acta 46, 1731 (2001).
107. M. Solomon, M. Xu, E. M. Eyring, and S. Petrucci, J. Phys. Chem.
98, 8234 (1994).
108. D. M. Vosshage and B. V. R. Chowdary, Electrochim. Acta 40, 2109
(1995).
109. Y. Liu, J. Y. Lee, and L. Hong, J. Power Sources 109, 507 (2002).
110. A. R. West, Basic Solid State Chemistry. Wiley, Chichester,
1988.
111. J. H. Shin, K. W. Kim, H. J. Ahn, and J. H. Ahn, Mater. Sci. Eng.
B 95, 148 (2002).
112. C. G. Leo, G. V. S. Rao, and B. V. R. Chowdari, Solid State Ionics
148, 159 2002.
113. Z. Wen, M. Wu, T. Itoh, M. Kubo, Z. Lin, and O. Yamamoto, Solid
State Ionics 148, 185 (2002).
114. Z. Y. Wen, T. Itoh, N. Hirata, M. Ikeda, M. Kubo, and
O. Yamamoto, J. Power Sources 90, 20 (2000).
115. D. Golodnitsky, G. Ardel, and E. Peled, Solid State Ionics 47, 141
(2002).
116. B. Wang and L. Gu, Mater. Lett. 57, 361 (2002).
117. L. E. Brus, J. Phys. Chem. 90, 2555 (1986).
118. L. E. Brus, J. Chem. Phys. 79, 5566 (1983).
119. L. E. Brus, J. Chem. Phys. 80, 4403 (1984).
120. M. L. Granvalet-Manchini, T. Hanrath, and D. Teeters, Solid State
Ionics 135, 283 (2000).
121. Y. W. Kim, W. Lee, and B. K. Choi, Electrochim. Acta 45, 1473
(2000).
122. J. H. Magill, in Treatise on Materials Science and Technology
(J. M. Schultz, Ed.), Vol. 10A. Academic Press, New York, 1977.
123. H.-W. Chen and F.-C. Chang, Polymer 42, 9763 (2001).
124. W. Wieczorek, D. Raducha, A. Zalewska, and J. R. Stevens,
J. Phys. Chem. B 102, 8725 (1998).
125. W. Wieczorek, A. Zalewska, D. Raducha, Z. Florzanczyk, and J. R.
Stevens, Macromolecules 29, 143 (1996).
126. W. Wieczorek, A. Zalewska, D. Raducha, Z. Florzanczyk, and J. R.
Stevens, J. Phys. Chem. B 102, 352 (1998).
127. G. Drezen, D. A. Ivanov, B. Nysten, and G. Groeninckx, Polymer
41, 1395 (2000).
128. J. Przyluski, M. Siekierski, and W. Wieczorek, Electrochim. Acta
40, 2102 (1995).
129. J. C. Maxwell, A Treatise in Electricity and Magnetism, Vol. 1,
Chapter XI. Clarendon, London, 1892.
130. M. Nakamura, Phys. Rev. B 29, 3691 (1984).
131. C. W. Nan, Prog. Mater. Sci. 37, 1 (1993).
132. C. W. Nan and D. M. Smith, Mater. Sci. Eng. B 10, 99 (1991).
133. S. Kirkpatrick, Rev. Mod. Phys. 45, 574 (1973).
134. Mikrajuddin, F. G. Shi, and K. Okuyama, J. Electrochem. Soc. 147,
3157 (2000).
762 Polymer Electrolyte Nanocomposites
135. C. C. Liang, J. Electrochem. Soc. 120, 1289 (1973).
136. G. B. Appetecchi and S. Passerini, Electrochim Acta 45, 2139
(2000).
137. M. Marcinek, A. Zalewska, G. Zukowska, and W. Wieczorek, Solid
State Ionics 136137, 1175 (2000).
138. M. M. Doeff, P. Georen, J. Qiao, J. Kerr, and L. C. De Jonghe,
J. Electrochem. Soc. 146, 2024 (1999).
139. R. Dupon, B. L. Papke, M. A. Ratner, D. H. Whitmore, and D. F.
Shriver, J. Am. Chem. Soc. 104, 6247 (1982).
140. P. Jeevanandam and S. Vasudevan, J. Phys. Chem. B 102, 4753
(1998).
141. H. Eliasson, I. Albinsson, and B. E. Mellander, Electrochim. Acta
43, 1459 (1998).
142. Z. Florzanczyk, E. Zygado-Monikowska, E. Rogalska-Joska,
F. Krok, J. R. Dygao, and B. Misztal-Faraj, Solid State Ionics,
152153, 227 (2002).
143. F. Croce, L. Persi, B. Scrosati, F. Serraino-Fiory, E. Plichta, and
M. A. Hendrickson, Electrochim. Acta 46, 2457 (2001).
144. C. Capiglia, P. Mustarelli, E. Quartarome, C. Tomasi, and A. Mag-
istris, Solid State Ionics 118, 73 (1999).
145. H. J. Walls, J. Zhou, J. A. Yerian, P. S. Fedkiw, S. A. Khan, M. K.
Stowe, and G. L Baker, J. Power Sources 89, 156 (2000).
146. J. Fan, S. R. Rafhavan, X. Y. Yu, X. A. Khan, P. S. Fedkiw, J. Hou,
and G. L. Baker, Solid State Ionics 111, 117 (1998).
147. A. Chandra, P. C. Srivastava, and S. Chandra, J. Mater. Sci. 30,
3633 (1995).
148. W. Wieczorek, Z. Florjanczyk, and J. R. Stevens, Electrochim. Acta
40, 2251 (1995).
149. J. Maier, J. Phys. Chem. Solids 46, 309 (1985).
150. A. Kumar and K. Shahi, Solid State Commun. 94, 813 (1995).
151. D. Golodnitsky, G. Ardel, and E. Peled, Solid State Ionics 85, 231
(1996).
152. C. J. Leo, G. V. S. Rao, and B. V. R. Chowdari, Solid State Ionics
148, 159 (2002).
153. J. H. Shin, K. W. Kim, H. J. Ahn, and A. H. Ahn, Mater. Sci. Eng.
B 95, 148 (2002).
154. G. Appetecchi, F. Croce, L. Persi, F. Ronci, and B. Scrosati, Elec-
trochim. Acta 45, 1481 (2000).
155. G. B. Appetecchi, P. Romagnoli, and B. Scrosati, Electrochem.
Commun. 3, 251 (2001).
156. J. W. Weston and B. C. H. Steele, Solid State Ionics 7, 75 (1982).
157. H. J. Wals, J. Zhou, J. A. Yerian, P. S. Fedkiw, S. A. Khan, M. K.
Stowe, and G. L. Baker, J. Power Sources 89, 156 (2000).
158. B. Kumar and L. G. Scanlon, J. Electroceram. 5, 127 (2000).
159. R. B. Seymour, Reinforced Plastics: Properties and Applica-
tions, p. 52. American Society of Metals, Int., Metals Park, OH,
1991.
160. F. M. Gray, Solid Polymer Electrolytes. VCH, Weinheim, 1991.
161. B. Choi and K. Shin, Solid State Ionics 8688, 303 (1996).
162. J.-S. Chung and H.-J. Sohn, J. Power Sources 112, 671 (2002).
163. W. Wieczorek, K. Such, S. H. Chung, and J. R. Stevens, J. Phys.
Chem. 98, 9047 (1994).
164. J. Przyluski, K. Such, H. Wycislik, W. Wieczorek, and Z. Flori-
anczyk, Synth. Metals 35, 241 (1990).
165. J. Sun, D. R. MacFarlane, and M. Forsyth, Electrochim. Acta 40,
2301 (1995).

You might also like