You are on page 1of 35

Advances in Colloid and Interface Science

. 97 2002 135
Electrorheological suspensions
Tian Hao
U
Rutgers, The State Uniersity of New Jersey, Department of Ceramic and Materials Engineering,
607 Taylor Road, Piscataway, NJ 08854-8065, USA
Abstract
.
The objective of this article is to give a review of electrorheological ER suspensions
whose rheological properties can abruptly change under an external electric field. Attention
is given to the physical backgrounds behind ER phenomena reported recently. The criteria
on how to design a high performance ER fluid and mechanisms explaining how an ER
suspension displays the ER effect are focused upon. We begin with a brief historic
introduction, ER materials, followed by positive ER effect, negative ER effect and photo-ER
effect discussions. The physical parameters that can substantially affect the ER effect are
discussed thereafter, and physical processes occurring in ER suspensions under an electric
field are reviewed. The mechanisms of the ER effect proposed before are summarized. A
future outlook on the ER material development and ER fluid applications is given. 2002
Elsevier Science B.V. All rights reserved.
Keywords: Colloidal suspension; Electrorheological effect; Electrorheological materials; Dielectric;
Conductivity
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2. Historical origin of the electrorheological response . . . . . . . . . . . . . . . . . . . . . . . 3
3. The electrorheological materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.1. Liquid continuous phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
U
Tel.: q1-732-445-6760; fax: q1-732-445-3258.
. E-mail address: haot@rci.rutgers.edu T. Hao .
0001-8686r02r$ - see front matter 2002 Elsevier Science B.V. All rights reserved.
. PII: S 0 0 0 1 - 8 6 8 6 0 1 0 0 0 4 5 - 8
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 2
3.2. Dispersed phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
3.2.1. Solid particulates heterogeneous ER materials . . . . . . . . . . . . . . . 8
3.2.2. Liquid materials homogeneous ER materials . . . . . . . . . . . . . . . 10
3.3. Additives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
4. The positive, negative ER and photo-ER effects. . . . . . . . . . . . . . . . . . . . . . . . . 11
5. Critical parameters on the ER effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
5.1. The electric field strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
5.2. Frequency of the electric field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
5.3. Particle conductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
5.4. Particle dielectric property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
5.5. Particle volume fraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
5.6. Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
5.7. Water content . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
5.8. Liquid medium. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
6. Physics of the ER materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
6.1. Forces relevant to the ER effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
6.2. Phase transition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
6.3. Percolation transition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
6.4. Conductive mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
6.5. Polarization process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
7. The mechanism of the ER effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
7.1. Fibrillation model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
.
7.2. Electric double layer EDL model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
7.3. Waterrsurfactant bridge mechanism. . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
7.4. Polarization model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
7.5. Conduction model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
7.6. Dielectric loss model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
8. Application of the ER fluid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
9. Summary and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
Acknowledgements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
1. Introduction
.
An electrorheological ER suspension is made from an insulating liquid medium
embodying either a semi-conductive particulate material or a semi-conductive
.
liquid material usually a liquid crystal material . The rheological properties
.
viscosity, yield stress, shear modulus, etc. of an ER suspension could reversibly
change by several orders of magnitude under an external electric field with the
strength of several kilovolts per millimeter. Since its mechanical properties can be
.
easily controlled within a wide range almost from a pure liquid to a solid , the ER
fluid could be used as an electric and mechanical interface in various industrial
areas. For example, it could be used in the automotive industrial for clutch, brake
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 3
and damping systems. It also could be used in the robotic arm joints and hands. It
could also be used for military purposes. The potential applications have stimu-
lated a great deal of interests both in academic and industrial areas since the ER
w x
effect was first described by Winslow 1 in 1949. There is a large body of literature
on the mechanism of the ER effect and the design of industrially applicable ER
devices.
Several comprehensive reviews on ER fluids and ER mechanisms have been
w x
published 28 , summarizing the ER research achievements obtained before 1996.
Many new findings have been made recently, and some could even totally change
previous concepts on the ER mechanism. A systematic review, combining the
previous achievements and recent findings, would thus be helpful in further
advancing ER research and applications.
The purpose of this article is to give a complete review of the ER area from
material science standpoints to physical mechanism investigations. To achieve our
goal, a brief historic review on earlier studies of a liquid or a dielectric suspension
under an external electric field is presented. The materials that are already used
for making ER fluids are then described. ER related effects including positive,
negative and photo-ER effects, are introduced thereafter. A summary on critical
physical parameters that play a key role in the ER response is presented, and
physical processes that occur in ER fluids are addressed. The ER mechanisms,
proposed on the basis of experimental facts on each period, are discussed chrono-
.
logically, and the comparisons between each model or theory are also provided.
The possible applications of the ER fluids in various industrial fields and the
affiliated technological problems are summarized.
2. Historical origin of the electrorheological response
The structure and the rheological property changes of a liquid or a dispersed
system under the application of an external electric field is called an electrorheo-
.
logical ER effect. The liquid or the dispersed system is generally called an
.
electrorheological ER fluid. At the end of the 19th century, the apparent viscosity
of some pure insulating liquids was found to increase with the application of an
w x
electric field 9 . At that time this phenomenon was termed the electroviscous
effect rather than the ER effect. Comprehensive investigations on the electrovis-
w x
cous effect were carried out by Andrade and co-workers 1013 , and two reasons
were proposed to explain the viscosity increase. Re-orientation of polarized
molecules along the direction of the applied electric field and ion aggregation near
the electrode surfaces were believed to be responsible for the viscosity increase.
According to this theory, non-polar liquids should not display an electroviscous
effect.
After the electroviscous effect was reported, many other solution systems were
studied under an electric field. More complicated systems, e.g. electrolyte solutions
and pure liquid systems embodying micro-particles, were investigated. It was found
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 4
that the electroviscous effect of an electrolyte solution is very much stronger than
w x
the effect observed in a pure liquid. Onsager 14 used the ion atmosphere
concept to understand the viscosity increase observed in the electrolyte solutions.
The distortion of the ion atmosphere under an electric field would generate more
drag force in a shear field, thus leading to the apparent viscosity increase. The
apparent viscosity of a pure liquid containing particulate the suspension system
was found to increase to a considerable extent under an electric field. In a dilute
suspension, the apparent viscosity would increase with the particle volume fraction
and the surface charge of the particulate material. An apparent viscosity equation
w x
was obtained by Smoluchowski 15
2
1
m
.
s 1 q 2.5 1 q 1
m 2
/ 5
2 r
m
where represents the apparent viscosity of the suspension, the viscosity of the
m
.
continuous phase pure liquid , the particle volume fraction, the conductivity
of the suspension, r is the radius of the particle, is the zeta potential of the
.
particle, is the dielectric constant of the continuous phase. Eq. 1 was derived
m
.
under the presumptions that: a the suspension is dilute enough usually the
. .
particle volume fraction is less than 10% ; and b there is no overlapping between
.
the electric double layers EDL . This effect is usually called the primary electro-
viscous effect, stemming from the distortion of the electric double layer under an
electric field and a shear field. If the suspension is concentrated, the electric
double layers would overlap each other, and the electric repulsive forces would be
.
obvious. Eq. 1 could not hold for such a system. The increase of viscosity under
an electric field was also observed in concentrated suspensions; however, this effect
was called the secondary electrovisous effect, resulting from the overlapping of the
electric double layer. The literature discusses a tertiary electroviscous effect that
refers to the viscosity increase observed in the suspension containing polymeric
particulate material. The detailed discussion on the electroviscous effect was made
w x
by Conway and Dobry-Duclaux 16 . Generally speaking, the viscosity increase due
to the electroviscous effect is not very large, mostly within a factor of 2. It could
not be compared with the ER effect, which would give a 10 000-fold increase in
rheological properties. However, in some older studies, the electroviscous effect
actually means the ER effect, which causes some confusion.
In early 1939, Winslow began experimental research on the electric-field-induced
viscosity increase in a suspension system made from solid semi-conducting particu-
lates dispersed into a low viscosity and high insulation oil. He first got a patent in
w x w x
1947 17 and then in 1949 reported his results in the J. Appl. Phys. 1 . He found
that a several hundred gram per square centimeter shear force could be obtained
under an electric field strength of 3 kVrmm. The solid particulate materials he
used included starch, flour, gelatin, and limestone. The liquids he used included a
transformer oil, mineral oil, and silicone oil. He observed that under an electric
field a fibrillation structure bridging between two electrodes is formed, and it has a
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 5
high strength, resulting in viscosity of the suspension increasing by several orders
of magnitude. The electric-field-induced effect he observed is much stronger than
the so-called electroviscous effect, and is an entirely new research field. This is the

reason that the ER effect is also called the Winslow effect in some German
works, the ER effect is also called the Oppermann effect, as G. Oppermann was
.
thought to be a pioneer in this field . Winslow is regarded as the pioneer who first
found that the fibrillated chain structure was formed in the ER suspension and
attributed the ER effect to the fibrillated structure formation, which becomes a
typical feature of an ER fluid, differentiating it from a non-ER fluid. Fig. 1 shows
the microstructure change of an ER fluid before and after an electric field is
applied. Before an electric field is applied, the particulates in the suspension are
randomly distributed, while after an electric field is applied the particulates are
orientated along the direction of the applied electric field. The microstructure
change from the disordered state to an ordered state determines the response time
of the ER effect. Besides these milestone findings, Winslow also proposed many
potential ER devices, such as clutches, brakes, and valves. However, his research
did not attract much attention at that time. There were almost no ER-related
w x
publications until 1960. In early 1962, Deneiga 18 investigated how an electric
field would change the rheological properties of a suspension. Extensive studies
w x
were carried out by Klass 19,20 in 1967, and the dielectric tool was first used to
characterize the ER effect, and the polarization of particles was linked to the ER
effect. The distorted electric double layer model was proposed to explain the ER
effect. In the early 1970s, the electric double layer model was further extended by
w x
Uejima 21 and later in 1984 the proton polarization model, based on the electric
w x
double layer model, was proposed by Deinega 22 . Since water is important in the
w x
electric double layer formation, Stangroom 23,24 extended the electric double
layer model to such an extent that water is regarded as necessary material for the
ER effect.
. . Fig. 1. Schematic illustration of the structure change of an ER fluid before a , and after b an
external electric field is applied. The two parallel dark lines stand for two electrodes. The crystallized
. w x structure is of the body-centered-tetragonal b.c.t lattice. Reprinted with permission from Ref. 126 .
Copyright 1999 American Chemical Society.
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 6
ER research became active at the beginning of 1980, and many publications
appeared at that time to address both basic academic interests and potential
industrial applications. ER technology was believed to bring a revolutionary inno-
vation to many industrial areas, especially to the automotive industry. Efforts were
made to elucidate the mechanism of the ER effect and to make a high perfor-
mance ER fluid that meets industrial requirements.
3. The electrorheological materials
Most ER fluids are made of solid particulate material dispersed into an insulat-
ing non-polar liquid. The solid particulate material includes inorganic non-metallic,
organic and polymeric semi-conductive materials. The inorganic materials are
basically the ionic crystalline material, while organic and polymeric semi-conduc-
tive materials generally have a bond conjugated structure, and are electronic
conductive materials. The continuous liquid phase is usually polydimethylsiloxane
.
oil silicone oil , vegetable oil, mineral oil, paraffin and chlorinated hydrocarbon
oils, etc. These oils must have a low conductivity and a high breakdown strength. A
.
good ER fluid should have: a a high yield stress preferably equal to or larger than
.
5 kPa under an electric field of 2 kVrmm; b a low current density passing
2
.
through the ER fluid preferably less than 20 Arcm ; c a wide working
temperature range. A good ER fluid should give a strong ER effect within the
.
temperature range y30 to 120C; d a short response time. The response time of
an ER fluid scales at 10
y3
s. For some specific purposes, an even faster response is
.
required; and e high stability. The ER fluid should be chemically and physically
stable. There should be no particle sedimentation and material degradation prob-
lems. The experimental characteristics of ER fluids are shown in Table 1. The
particle size is between 0.1 and 100 m, and the particle volume fraction is
between 0.05 and 0.50. The applied electric field is between 0.5 and 3 kVrmm.
Before 1985, all ER fluids contained small amounts of water. Such ER fluids
have many shortcomings, such as narrow working temperature range due to the
water evaporation at high temperatures; high current density due to the higher
w x
conductivity of water; and device erosion caused by water, etc. In 1985 Block 25
developed a water-free acenequinone radical polymer ER fluid, which was thought
Table 1
The experimental characteristics of an ER fluid
Liquid phase Particle ER suspension
Relative dielectric constant ;2 210 000
y10 y16 y7 y9 y16
. Conductivity Srm 10 to 10 10 10 10
. Viscosity at no electric field Pa.s 0.0110 0.110
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 7
to be a new class ER fluid termed as the anhydrous ER fluid. Anhydrous ER fluid
is believed to be much more promising from the standpoint of industrial applica-
tions. It thus received greater attention afterwards, and many other kinds of
w x
water-free ER fluids were developed thereafter 26,27 . In the late 1980s, it was
realized that water-free ER fluid had another big problem particle sedimenta-
tion, which could cause the ER fluid to malfunction completely, thus severely
limiting practical applications. Effort was then expended to develop a homoge-
neous ER fluid with no particulate material inside. Low molecular liquid crystalline
.
LC materials were extensively studied. In the early 1960s, the viscosity of LC
materials was found to increase under an electric field, though the increment was
w x
only a few times 28,29 . In 1992 an extremely large ER effect was found by Yang
w x w x
30 and Inoue 31 in certain LC materials. Other LC ER systems were developed
w x
later on for the purpose of mechanism study 3234 . Chlorinated paraffin
w x
oilrsilicone oil emulsion systems 35 and polymeric glycolrsilicone oil emulsion
w x
systems 36 were also found to exhibit an ER effect. The mechanism of the
w x
rheological response of the oil-in-oil emulsion system was addressed by Ha 37
quite recently. Homogeneous ER fluids are believed to offer a new and promising
type of ER fluid. However, higher viscosity under zero electric field and
liquidliquid segregation problems are obvious obstacles for such systems.
Generally speaking, most ER fluids have three components the dispersed
phase, the continuous phase and small amounts of unavoidable additives, such as
inorganic salts and water, etc. The ER fluids are either heterogeneous or homoge-
neous. In the heterogeneous group, there are inorganic, organic or polymeric
particulate materials. In the inorganic group, oxide and non-oxide materials give a
quite different ER effect, and are therefore discussed separately. Fig. 2 shows the
tree structure of the ER fluid, and we will follow this classification.
3.1. Liquid continuous phase
The liquid continuous phase of an ER fluid is usually an insulating oil. Stan-
w x .
groom 23 thought an ideal dispersing liquid material should: a have a high
boiling point and low solidifying point. In other words, it should not easily
.
evaporate within the working temperature range; b have a low viscosity for
.
keeping the viscosity of the whole suspension at a low level at zero electric field; c
have high resistance and high breakdown strength, i.e. withstand a high electric
.
3
.
field; d have a high density )1.2 grcm is best . The particle sedimentation
problem might not occur until the densities of both the liquid and the solid match
.
each other; e have high chemical stability. The liquid medium would not degrade
.
or chemically react with other materials once the ER fluid is prepared; f have an
obvious hydrophobicity, and would not adsorb too much moisture from the envi-
.
ronment; and g have low toxicity and low cost. Currently used oil materials
include silicone oil, vegetable oil, mineral oil, paraffin, kerosene, chlorinated
hydrocarbon, transformer oil, etc. Large density oil, such as fluoro or phenyl
silicone oil, is also used.
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 8
Fig. 2. The classification of ER materials.
3.2. Dispersed phase
3.2.1. Solid particulate heterogeneous ER materials
3.2.1.1. Inorganic oxide materials. Some metallic oxides or ceramic materials
sintered from several oxides were found to give a good ER effect. The composi-
tions of such ER fluids are summarized in Table 2. Most oxide ER fluids contain
water, which is a big drawback for this type of fluid.
3.2.1.2. Non-oxide inorganic. Non-oxide inorganic ER fluids were mainly developed
in the late 1980s and early 1990s. They could give an extremely strong ER effect

without any amount of water though water could also enhance the ER effect
Table 2
Oxide ER fluids
Dispersed phase Dispersant Additive Reference
w x Piezoelectric ceramic Mineral oil or xylene Water or glycerol oleates 38
. w x Iron IIrIII oxide Petroleum fractions or dibutyl Water or surfactant 39
sebacate
w x Silica Kerosene or dibutyl sebacate Water and soaps 39
w x Mineral oil, silicone oil Water and glycerol oleates 40
. w x Tin II oxide Petroleum fractions Water or surfactant 39
Titanium dioxide Mineral oil or p-xylene or
w x polyphenylmethylsiloxane Water and glycerol oleates 38
Al O , Cu O, MgO,
2 3 2
w x ZnO, La O , ZrO , Ta O Mineral oil Polybutylsuccimide 41
2 3 2 2 3
MnO , CoO, Nb O , etc.
2 2 3
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 9
Table 3
Non-oxide inorganic ER fluids
Dispersed phase Dispersant Additive Reference
w x Aluminosilicate Silicone oil or hydrocarbon Surfactant 42
SirAl ratio from
. 8:1 to 175:1
Crystalline zeolite Silicone oil or high dielectric
w . . x w x M AlO SiO wH O Constant hydrocarbon oil 43
xrn. 2 x 2 y 2
M is metal cation
w x Zeolite Silicone oil or hydrocarbon oil 44
w x Micro-glass beads Transformer oil 26
Aluminosilicate with 125%
w x crystallized water, SirAl ratio Silicone oil Polyhydroxylsiloxane 45
between 0.15 and 0.80
w x Silicate, silica-alumina Mineral oil, polyalkylene Sulfonaes 46
Paraffin mineral oil Phenates
Polyphenyl, phosphoric Phosphonates
acid ester, etc. Succinic acid, etc.
w x LiN H SO Silicone oil Block polymer 47
2 5 4
w x BN, AlN, B C Silicone oil Succimide 41
4
.
substantially . The development of these kind of ER fluids was very inspiring at
that time. Among them, aluminosilicates, especially zeolite family materials, re-
ceived a great deal of attention. Table 3 shows some non-oxide inorganic ER fluids
reported so far.
3.2.1.3. Organic and polymeric materials. Although the non-oxide inorganic ER fluid
does not contain water, and has a strong ER effect, its conductivity is too high,
especially in a high temperature environment. The density of the dispersed phase is
much higher compared than that of the dispersing medium and the suspension is
usually unstable. The particles are too hard and abrasive to the ER device. The
organic and polymeric materials were believed to be better than inorganic materi-
als, and they were investigated thoroughly. Some organic and polymeric ER fluids
are shown in Table 4.
Organic and polymeric ER fluids can be classified into two categories. The first
is materials having conjugated bonds. Two methods are generally used to control
the conductivity of the conjugated material: doping the metallic ion or metallic
oxide and controlling the carbonating temperature. Acene-quinone radical poly-
mers and polyacrylonitrile are examples of the conjugated bond. This type of
material could be highly polarized under an electric field, thus they have a large
dielectric constant. The second category is materials having a highly polarizable
group on the molecular chain, such as the hydroxyl, cyano, or amido, etc. Poly-
methyacrylic acid, starch, dextran, etc., belong to this group. Those materials are
polyelectrolytes, which have high molecular weight and high charge density. Almost
all polyelectrolytes, including both natural and synthetic, can give an obvious ER
effect, which is possibly related to their strong moisture absorptivity. However, the
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 10
Table 4
The organic and polymeric ER fluids
Dispersed phase Dispersant Additive Reference
w x Acene-quinone radicals Chlorinated hydrocarbon 48
polymer
w x Crosslinked polyvinylsilane Fluorinated silicone oil Electrolytes 49
w x Cellulosic material Transformer oil, Water or other 21
vegetable oil, etc. electrolyte
Polyphenyl Mixture of carbon-based oil Aromatic hydroxyl
w x Polyvinylideneshalides compound 50
w x Polypyrroles Transformer oil, vegetable oil 51
Chlorinated paraffins, etc.
w x Polyanilines Silicone oil 52,53
Ionic dye material Mineral oil, silicone oil, Molecule containing
w x White oil, etc. hydroxyl carboxyl 54
w x Sodium carboxylmethyl Polychlorinated biphenyls Sorbitan 55
Dextran or o-dichlorobenzene, etc. monooleate
Polymeththacrylic acid cross
w x linked with divinylbenzene Hydrycarbons Water 55
w x Starch Mineral oil, transformer Water 17
w x Hydrocarbon oil Water 56
w x Oxidized polyacrylonitrile Silicone oil 57
w x Carbonated aromatic sulfonic Fluorosilicone oil or other 58
acid or a salt modified silicone oil
ER effect of the organic and polymeric ER fluids is weaker compared with that of
the non-oxide inorganic material.
3.2.2. Liquid materials homogeneous ER materials
A homogeneous ER fluid, a liquid dispersed into an insulating oil, was thought
to be the best ER fluid so far, because this kind of ER fluid does not have a
particle sedimentation problem as the heterogeneous ER fluid usually has. How-
ever, such an ER system would not give a strong ER effect and easily segregate
into two phases. It also has a large viscosity at zero electric field, which is not good
for practical uses. LC polymer material constitutes a large portion of the homoge-
neous ER fluid. Some homogeneous ER fluids are shown in Table 5.
3.3. Addities
As described above, most ER fluids contain additives that are less than 5% by

weight. In most cases the additives are important. Water, acid inorganic and
.
organic , alkali, salt, and surfactants, are most common. A review on ER additives
w x w x
is given by Block 2 and Tomizawa 46 . The ER additive usually can be hydro-
lyzed, and it would not be able to activate the ER system without water. The
amount of additive is very important. Less than 0.01 wt.% would not give any
enhancement and greater than 5 wt.% would give a large electric current.
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 11
Table 5
The homogeneous ER fluid
Dispersed phase Dispersant Additive Reference
w x Aluminum soap Mineral oil, silicone oil 2,6-Ditertbuthyl 59
phenol
Polyalpha-olefins, etc.
. w x Poly -glutamate Cyclic ketone 60
. w x Poly n-hexyl isocyanate p-Xylene 30
w x LC polysiloxanes Silicone oil 33
. w x LC polysiloxanes 4- Pentyloxy -4-biphenyl 33
carbonitrile
w x 4-n-Pentyl-4-cyanobiphenyl 32
w x Chlorinated paraffinrsilicone 35
oil emulsion
w x Urethane-modified polypropylener 36
silicone oil emulsion
w x Caster oilrsilicone oil emulsion 37
Besides water, other polar liquids, such as alcohol, dimethylamine, acetamide,
diethylamine, glycerol, etc., can enhance the ER effect substantially. The small
amount of polar liquid can dramatically increase the dielectric constant of the
dispersed particle, which was regarded as the possible reason that the ER effect
was manifested.
Another commonly used additive is surfactant, e.g. oleates, sorbitan, etc. An-
ionic, cationic, and non-ionic surfactant are used in the ER systems, while ampho-
teric surfactant is seldom used. Surfactant could help the ER suspension be more
stable, and it also can enhance the ER effect.
4. The positive, negative ER and photo-ER effects
Since the innovation of the ER effect in 1947 by Winslow, attention has been
paid to how to increase the rheological property induced by the electric field. The
electric-field-induced rheological property increase is termed the positive ER effect
.
see Fig. 3 . Many materials have been tried to see if a high increment of the
w x
rhelogical property could be achieved. In 1995, Boissy 61 reported a phenomenon
.
observed in the suspension containing polymethylmethacrylate PMMA powder
dispersed into the mixture of silicone oil. They found that the apparent viscosity of
the whole suspension decreases as the external electric field increases. This
phenomenon is totally opposite to that discussed above, and termed the negative
.
ER effect see Fig. 3 . Several other systems were also found to display the
w x
negative ER effect, for example, teflonrsilicone oil 62 , and magnesium
w x
hydoxidersilicone oil 63 systems. Liquid crystalline polysiloxane was found to
.
show a positive ER effect when it was dissolved in the 4- pentyloxy -4-biphenyl-
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 12
. Fig. 3. Schematic illustration of the relationship between the viscosity and the electric field: a
. negative ER effect; b positive ER effect.
carbonitrile, however, it shows a negative ER effect when it was dissolved in
. w x
N- 4-methoxybenzylidene -4-butylaniline 33 , indicating that the continuous phase
also plays a very important role in the ER response. The electrophoresis process is
usually attributed to be responsible for the negative ER effect. Just as the positive
ER effect has vast potential industrial applications, the negative ER effect could
also be used in industry in case the viscosity should be lower, though no such
device has been developed so far. The detailed mechanism of both the positive and
negative ER effects will be addressed later.
Both the positive and negative ER effect could be enhanced by the UV
w x
illumination 64,65 . This phenomenon is termed the photoelectrorheological
. w x
photo-ER effect. Komada 66,67 studied the photoelectrorheology of TiO
2
nanoparticle suspensions, and found that water plays an important role. Low water
content makes the suspension display a positive photo-induced ER effect, while a
high water content gives a negative photo-induced ER effect. Photo-generated
carriers were thought to change the electric property of the TiO and then
2
enhance the ER performance.
5. Critical parameters on the ER effect
The ER effect depends on the applied electric field strength, frequency of the
electric field, particle conductivity, particle dielectric properties, particle volume
fraction, temperature, water content, liquid medium, etc. A brief review on how
these parameters influence the ER effect is described below.
5.1. The electric field strength
The yield stress of an ER fluid greatly relies on the applied electric field
strength. The critical electric field strength, E , exists, and the ER fluid would not
c
show any ER effect until the applied electric field strength is larger than the
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 13
critical one. The yield stress was found to linearly increase with the electric field
w x
68 :
. .
sk E y E 2
y c
Where represents the yield stress of an ER fluid, k a constant, E the applied
y
w x
electric field strength. However, other researchers 17,57,48 believed that the yield
stress should be directly proportional to the square of the electric field strength.
When the electric field is high enough, the yield stress and the apparent viscosity
tend to become saturated.
5.2. Frequency of the electric field
A dc electric field is mostly used to generate a detectable ER effect. An ac
electric field is very useful to study the mechanism of the ER effect and to
determine the response time of an ER fluid. Since an ER fluid has a response time
of approximately 1 ms, its viscosity and yield stress are expected to decrease with
increasing frequency, as it would be unable to catch up with the change of the
w x
electric field at high frequency. Klass 19,20 first addressed this issue, and found
that the apparent viscosity of a silicarsilicone oil system decreases with increasing
frequency. A sharp decrease followed by recovery of the apparent viscosity was
.
found to occur at 200 Hz. A so-called Flow-Modified-Polarization FMP was
w x
proposed to explain this sharp discontinuity by Block 69 . A resonance between
the applied electric field and mechanical field should take place when the shear
rate is 4 times the frequency of the applied electric field. A weak particle
interaction was expected at this point.
w x w x
Hao 70 used the WangerMaxwell polarization 7173 to understand the
frequency dependence of the yield stress. He found that the yield stress decrease
corresponds to the decrease of dielectric constant of the whole suspension when
frequency increases. The particle conductivity determines whether the yield stress
decrease appears at high or low frequency. High conductivity particles give an
obvious ER effect even in very high frequency fields, indicating a short response
time. An oxidized polyacrylonitrilersilicone oil ER fluid was investigated under an
ac field, and a good agreement between the experimental result and the prediction
.
based on the WangerMaxwell polarization was obtained see Fig. 4
5.3. Particle conductiity
w x
Block 69 studied how the particle conductivity influences the ER effect by using
the acene-quino radical polymerrsilicone oil, and found that the static yield stress
peaks at a particle conductivity of approximately 10
y5
Srm. A similar tendency was
w x
found in the oxidized polyacrylonitrile silicone oil ER system 74 , however, the
yield stress peaks at a particle conductivity of approximately 10
y7
Srm, rather than
10
y5
Srm. Besides the influence on the ER effect, the particle conductivity also
determines the current density of the whole suspension and the response time of
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 14
Fig. 4. The shear stress obtained at 300 Vrmm and dielectric constant predicted using Wagner model
. w x solid line are showed again electric field frequency. Reprinted with permission from Ref. 70 .
the ER fluid. The response time was found to be inversely proportional to the
w x w x
particle conductivity both experimentally 75 and theoretically 70 .
5.4. Particle dielectric property
Since the ER effect is induced by an external electric field, the polarization is
believed to play an important role, and the particle dielectric property should be
dominant in the ER effect. The dielectric tool is thus frequently used for investigat-
ing how the particle dielectric properties influence the ER effect.
A linearly increasing relationship between the dielectric constant of the whole
ER suspension and the particle volume fraction was found in the silicarsilicone oil
w x
system 20 . The electric double layer overlap was thought to be a main reason, as a
dielectric dispersion was found to appear at 1000 Hz. Furthermore, the dielectric
constant of the ER suspension was found to change with the applied electric field
w x
strength. Deinega 76 found the dielectric constant increases with the electric field
.
strength 0.44 kVrmm , and levels off at high electric field. Similar results were
w x
found by Klass 20 only when the particle volume fraction is less than 10%. When
the particle volume fraction is as high as 46%, the dielectric constant decreases
with electric field strength.
The dielectric investigation was concentrated with establishing a direct correla-
.
tion between the dielectric properties of the particle and ER suspension and the
w x
ER effect. Uejima 21 claimed that the dielectric constant of the ER suspension
containing electrolyte should be larger than that of the ER suspension without
electrolyte. This is the reason that the ER effect of an ER fluid containing an
w x
electrolyte is stronger than that of an ER fluid without an electrolyte. Block 69
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 15
suggested that the polarization rate and its magnitude are very important for ER
w x
response. Filisko 8,64 even thought the ER effect might be associated with the
w x w x
low-field dielectric dispersion. Kawai 77 and Ikazaki 78 proposed that for a good
ER fluid, its dielectric relaxation frequency should be between 100 and 10
5
Hz and
the difference of the dielectric constant below and above the relaxation frequency
must be large. All the work above addressed the relationship between the dielectric
w x
properties of the whole suspension and the ER effect. Hao 74 studied how the
particle dielectric properties alters the ER effect. The oxidized polyacrylonitrile
. .
OP and aluminosilicate AS powder of different dielectric constant from 2 to
. . Fig. 5. Shear stress of the OP and AS suspensions against: a the dielectric constant; b the dielectric
w x loss tangent of dispersed particles. Reproduced with permission from Ref. 74 .
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 16
.
y5 y12
.
10 and conductivities ranging from 10 to 10 Srm were used, and it was
found that there is a complicated relationship between the ER effect and the
.
particle dielectric constantrthe dielectric loss see Fig. 5a,b . The shear stress does
not change monotonically with the dielectric constant or loss even within the same
series of samples, and the particle dielectric loss was found to play an important
role in the ER response. By analyzing the dielectric data of currently used ER
materials, an empirical criterion was proposed for selecting ER material: the
particle dielectric loss tangent should be approximately 0.10 at 1000 Hz. Also the
larger the particle dielectric constant, the stronger the ER effect.
5.5. Particle olume fraction
The yield stress and the apparent viscosity of an ER suspension are largely
dependent on the particle volume fraction. A linear relationship between the yield
stress and the particle volume fraction was derived on the basis of the fibrillation
w x
model 79 . However, other researchers found theoretically that the yield stress
goes through a maximum as the particle volume fraction increases, and the
w x w x
maximum appears at a very high particle volume fraction 80,81 . Uejima 21 found
experimentally that the yield stress peaks at particle weight percentage 10%. Block
w x w x
69 and Xu 57 found the yield stress parabolically increases with the particle
w x
volume fraction. Hao 82 found that there is a critical particle volume fraction for
the ER suspension. When the particle volume fraction exceeds this critical value, a
sharp increment of the rheological property should be observed. Percolation theory
was used to understand this phenomenon, which will be discussed in detail in the
next section.
5.6. Temperature
There are two reasons that temperature might change the ER effect substan-
tially. The first is that temperature can definitely change the polarizability of the
ER suspension, as the particle conductivity and dielectric constant vary with
temperature. The second is that temperature would directly impact particle ther-
mal motion. If the Brownian motion would be intensified at high temperature and
could become strong enough to compete with the particle fibrillation, thus the ER
effect would become weak. Whether temperature increase would intensify or
weaken the ER effect is really dependent on which factor would become dominant
at that temperature.
An obviously improved ER response was found at higher temperatures in several
w x
ER suspensions, including both the inorganic and polymeric systems 19,20,83,84 .
The readily facilitated polarization of the electric double layer in a hydrous ER
system was thought to contribute to the ER effect improvement at high tempera-

ture, while the change of the particle intrinsic property e.g. ion mobility inside the
.
particle was thought to be responsible for the ER effect enhancement in an
w x
anhydrous ER system. Hao 85 found that the temperature dependence of the
yield stress is dependent on the particle conductivity. As mentioned in Section 5.3,
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 17
there is an optimal particle conductivity at which the ER effect peaks. A decrease
in shear stress with elevated temperature would be found in a suspension with the
particle conductivity greater than the optimal value, while an increase would be
found if the particle conductivity is less than the optimal value.
Moist ER suspensions are believed to have a narrow temperature range some-
what between 20 and q70C, possibly due to the water solidification and
evaporation. An anhydrous ER suspension could work in a wide temperature
range, however, it is limited by the large conductance at high temperature, as most
anhydrous ER fluids are made from ionic materials.
5.7. Water content
Before the development of anhydrous ER fluid, water was believed to play a key
role in the ER response. Many efforts were made to study how trace amount water
affects the ER effect. The yield stress of the ER suspension was unambiguously
w x
found to go through a maximum with water content increasing 22,55,85 . A
w x
different case was found by Uejima 21 the viscosity increases with water
content and saturates when the water content exceeds a certain value. The
function of water in the ER effect was primarily assumed to increase the dielectric
constant of the particle, which in turn results in a strengthening of particle
w x
interaction 21,22 . Another hypothesis is that water could stick the particles
w x
together due to the property of high surface tension of water 24 .
5.8. Liquid medium
The liquid medium gives a pronounced difference for the ER effect. It changes
the particle sedimentation due to the density of the liquid and the viscosity of
whole suspension at zero electric field. The one example is that the same solid
particulate material displays an ER effect in one medium, however, the ER effect
w x
disappears when the particle is dispersed into another medium 86 . The dielectric
mismatch of the solid-to-liquid dielectric constant was believed to be a main reason
for such differences.
6. Physics of the ER material
The ER effect is induced by an electric field in a colloidal suspension. The
mechanism of the ER effect should be related to the physical processes occurring
in the suspension. Elucidating the physical basis of those processes would be very
helpful to correctly understand the ER mechanism. In this section, the forces
involved in the ER response, the percolation transition due to the particle volume
percentage increase, the phase transition induced by the electric field, and the
conductive mechanism of the ER suspension, as well as the polarization, are
discussed.
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 18
6.1. Forces releant to the ER effect
An ER fluid works under a high electric field, and the particles should be
polarized to such an extent that the electrostatic force between them would be very
strong. However, other forces act in ER fluids. Examples are the hydrodynamic
force on the particle due to the viscous continuous phase, Brownian force due to
the thermal motion of the continuous phase, short-range repulsive forces arising
w x
from the Born repulsion or steric interaction, adhesion due to water 24 or
w x
surfactant 87 , colloidal interaction such as the van der Waals attraction and
DLVO-electrostatic repulsion forces. The structure and rheological properties of
ER suspensions depend on the competition among all of the forces mentioned
above. Various dimensionless groups have been employed to describe the relative
w x
importance of those forces. For example, the Mason number 3,88 , Mn s
.
2

6r E where is viscosity of the dispersing medium, is the shear


0 sm m
y
sp sm
rate, is permittivity of free space, s , and are the static
0 sm sp
q2
sp sm
dielectric constants of dispersing medium and dispersed particles, E is the applied
.
electric field , is used to scale relative importance of hydrodynamic to the elec-
w x
3

trostatic polarization forces. The Peclet number 3,89 , Pe s 6 a rk T where


sm b
.
a is particle radius, k is the Boltzmann constant, T is temperature , indicates the
b
ratio of the hydrodynamic to thermal forces.
Among these forces, the electrostatic force receives particular attention, and is
regarded as a main source of the ER effect. Calculation of the electrostatic force
by using various physical assumptions andror models, for example, the point-
w x w x
dipole approximation 90,91 , multipole and multibody effect 9295 , sphere within
w x w x w x
lattice 93,9699 , energy method 100 , non-linear conduction 101,102 , etc., have
been tried before, deriving so-called polarization and conduction models which will
be discussed in Section 7.
6.2. Phase transition
Phase transition is observed in colloidal suspensions. As the particle volume
fraction increases, the equilibrium phase changes from a disordered state, to
.
coexistence with a crystalline phase closed-packed particle arrangement , then to a
w x .
glass, and finally to a fully crystalline state 103,104 . The face-centered cubic fcc
w x
lattice was found to be the most stable structure both theoretically 105 and
w x
experimentally 104,106,107 , compared with other crystal forms such as, body-
. .
centered tetragonal bct and hexagonal closed-packed hcp structure.
The electric-field-induced phase transition in an ER suspension was found to be
w x
different from in a general colloidal suspension. Tao 108 predicted theoretically
that the bct structure has an energy lower than that of the fcc and other structures,
w x
based on dipolar interaction energy calculations. The laser diffraction method 109
was employed to experimentally determine the crystal structure within the fibril-
lated columns by using a uniform glass microspherersilicone oil system, and a bct
structure was found as expected. The bct structure was also observed by using
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 19
w x
confocal scanning laser microscopy quite recently 110 . This is a very important
finding in the ER field. It points out for the first time that an external electric field
can induce a phase transition in ER fluids, changing from a meta-stable liquid to a
crystallized solid.
6.3. Percolation transition
A disordered or amorphous system can be understood with the aid of percolation
theory, which deals with how the short-range finite connectivity would finally
w x
change to a long-range infinite connectivity, above a threshold probability 111 .
The phase transition in the ER system described earlier could be the analog of a
disorderorder transition, which can also be understood by using the percolation
theory. An investigation of the percolation transition in an ER fluid was carried
w x
out by Hao 82 . It was found that the electrorheological properties the complex
U
.
viscosity , the real modulus G and the imaginary modulus G increase sharply
once the particle volume fraction exceeds a critical value . is a constant and
c c
does not change with the applied electric field strength. Fig. 6 shows the dimen-
.
sionless real modulus GrG against the dimensionless particle volume fraction
c
.
r . The data overlap regardless of the applied electric field strength, indicat-
c
ing that an electric field induced phase transition occurs when the particle volume
fraction increases. The critical volume fraction is calculated to be 37% based on
the experimental data. According to the percolation theory, a network microstruc-
ture, rather than the generally accepted fibrillated chain structure, will be formed
in an ER fluid when the particle volume fraction is above the critical point of 37%.
w x
This was experimentally verified 82 . The fibrillated chain can only be formed
when the particle volume fraction is less than 37%. Accordingly, three kinds of
particleparticle clusters or paths were presumably thought to be formed in an ER
.
fluid: a continuous paths, which start at one electrode and connect with another
.
electrode; b branched paths, which start from one electrode and end between the
.
two electrodes; and c isolated paths, whose two ends exist in ER fluid but do not
connect with any electrodes. The weight fraction of various paths can be calculated
w x
112 and are shown in Fig. 7. The percolation path was found to start to appear at
a particle volume fraction of 24%, and to become dominant at the volume fraction
of approximately 40%, where the whole ER suspension was occupied by a percola-
tion network. The calculated critical volume fraction value agrees well with the
experimental measured one.
6.4. Conductie mechanism
w x
The conductive properties of an ER suspension were studied by Hao 112 , and it
was found that the conductive behavior obeys the Quasi-One-Dimensional Vari-
.
able Range Hopping Quasi-1d-VRH model. The Variable Range Hopping model
.
assumes that once the energy barrier between two molecules particulates is too
wide, electrons can move by the hopping mechanism rather than by the tunneling
mechanism. The electron hopping process from the localized state below the Fermi
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 20
. Fig. 6. Dimensionless real modulus GrG against the dimensionless particle volume fraction
c
. w x r . Reproduced with permission from Ref. 82 .
c
energy to the unoccupied state over the Fermi energy is determined by the
temperature and energy barrier between the beginning state and the terminal
state. The most probable hopping range is exponentially related to temperature.
The Quasi-1d-VRH model assumes electrons hop only in a one-dimensional chain
and cannot hop in another direction. In view of the morphology of the ER
suspensions, this model sounds reasonable, as the fibrillated chains formed in the
ER fluids could provide a one-dimensional path for charge carriers to transfer
from one electrode to another. Also, the Quasi-1d-VRH model is consistent with
Fig. 7. Weight fraction of various paths in the ER suspension calculated by using Florys gelation
theory under the assumption that bct lattice is formed. X is the number of particles in the path. Wg is
w x the percolation path. Reproduced with permission from Ref. 112 .
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 21
Fig. 8. The dc current passing through the oxidized polyacrylonitrilersilicone oil ER fluid at 0.5
kVrmm and the applied oscillatory mechanical field of angular frequency s 1, and strain amplitude
50%. The shear stress recorded against time and strain against time are also shown here.
w x Reproduced with permission from Ref. 112 .
the percolation transition in the ER fluids, as this model is a consequence of strong
disorder. A very interesting phenomenon is that the dc current passing through the
w x
ER fluids were found to oscillate with the applied mechanical field 112 see Fig.
.
8 . The oscillatory frequency of the dc current is equal to that of the stress and
double that of the applied strain. The relationship between the oscillatory dc
current and the applied mechanical field was established by using the Quasi-1d-
w x
VRH-model 113 , agreeing well with the experimental results. A real time elec-
tronicmechanical signal inter-transferring sensor that used this phenomenon,
w x
such as seismograph, was proposed 113 .
6.5. Polarization process
Polarization processes are extremely important in an ER system. Generally, four
kinds of polarizations exist: electronic, atomic, the Debye and the interfacial
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 22
.
polarizations the WagnerMaxwell polarization . If the particulate material is an
ionic solid, ionic displacement polarization should also be considered. The Debye
and the interfacial polarizations are rather slow processes compared with elec-
tronic and atomic polarizations. Usually, the former two polarizations are called
the slow polarizations, appearing at low frequency fields, whereas the latter two are
termed fast polarizations, appearing at high frequencies.
It would be very intriguing to clarify which polarization process is responsible for
w x
the ER effect. Hao 114 investigated how the particle conductivity changes the
response time of the ER suspension and how the particle surface properties
influence the ER effect. Hao concluded that the interfacial polarization con-
tributes to the ER effect. This finding is consistent with his previous proposal that
a large dielectric loss is required for a good ER material, because only a material
having a large dielectric loss could give a large interfacial polarization once it is
dispersed into a liquid. This can be easily understood by considering where the
dielectric loss comes from.
The total polarization of heterogeneous material, P, can be expressed as:
.
P s P qP qP qP , 3
I D A E
where P , P , P , P stand for the interfacial, Debye, atomic, and electronic
I D A E
polarization, respectively. Correspondingly, the total dielectric constant, , can be
expressed as,
s q q q q
s I D A E
.
s q q q , 4
s I D
where , , and are artificially regarded as the dielectric constant induced
I D A E
by the interfacial, the Debye, the atomic and the electronic polarizations, respec-
tively. The static dielectric constant is represented by and the high-frequency
s
dielectric constant is represented by s q . For solid particles, dipole
A E
orientation is almost impossible, because the solidification usually fixes the molecule
with such rigidity in the lattice that there is little or no orientation of the dipoles,
w x
even in an extremely strong electric field 115 . Thus the Debye contribution can be
.
negligible, and Eq. 4 can be written as:
.
s q q 5
s I
Because the dielectric constant and the dielectric loss are not independent, and
the dielectric loss originally results from slow polarizations, i.e. the interfacial
polarization in this case, a suspension that has a large dielectric loss tangent
implies that the ratio of to is large. In other words, a large dielectric loss of
I
dispersed particles would result in a large dielectric loss of the whole suspension,
which means a large interfacial polarization.
In homogeneous two-component ER systems, the interfacial polarization is also
w x
found to be responsible for the ER effect 116 .
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 23
7. The mechanism of the ER effect
The mechanism of the ER effect has been targeted since the innovation of ER
fluids. Various models or mechanisms were proposed previously to explain the
observed ER phenomena. In this section, the discussion is focused on the physical
mechanism that relates the microscopic material-intrinsic properties to the macro-
scopic properties. The phenomenological models, such as rheological models that
characterize the ER particular rheological performance, will not be discussed in
detail. The mechanism will be discussed chronologically to show readers how the
ER mechanism has evolved.
7.1. Fibrillation model
w x
The fibrillation model was proposed by Winslow 1 based on his observation
that the fibrillated chains were formed in the ER suspension. The particle could be
polarized and be aligned as a dipole along the direction of the electric field. The
interaction between the polarized particulates would be dramatically increased,
resulting in the obvious ER effect. The particle could bear some net charge, arising
from the non-uniform polarization or ionic adsorption, thus electrophoresis could
contribute to the particle motion for particle re-arrangement. This model can be
called a primary polarization model, because the particle polarization was empha-
sized.
( )
7.2. Electric double layer EDL model
w x
The EDL model was primarily proposed by Klass 19,20 to explain why water
plays a key role in the ER response and why the ER effect could take place on a
millisecond time scale. Klass thought that the fibrillation process would be rather
slow compared with the ER response time, and thus the fibrillation model is
inadequate to describe ER phenomena. If water is in the ER suspension, each
particle would be surrounded by the EDL. The EDL could be polarized and
distorted. The neighboring distorted EDLs might overlap each other, generating
strong electrostatic repulsive forces and thus resulting in the ER effect. A more
w x w x
detailed EDL model was developed by Uejima 21 and by Deinega 22 to explain
the dependency of the ER effect on water, field frequency and temperature. The
ER effect could be enhanced if the charge carrier only moves within the diffuse
region, while it could be weakened once the charge carrier could transfer between
particles. However, the EDL model is unable to explain how the EDL overlap
could result in an increase of rheological properties of several orders of magnitude.
7.3. Waterrsurfactant bridge mechanism
w x
The water bridge mechanism was proposed by Stangroom 24 to explain moist
ER fluids. He thought that a good ER fluid should meet the following require-
. .
ments: a the liquid medium must be hydrophilic; b the solid particles must be
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 24
water-like and porous material with the capability of adsorbing and keeping some
.
amount of water; and c the water amount on the particle surface would de-
termine the ER effect. Under an electric field, ions from water could move out of
pores and migrate from one particle to another. Thus an adhesive water bridge
could be formed between particles. This water bridge is very strong due to the high
w x
surface tension of water. A similar surfactant bridge was observed by Kim 87 . See
w x
et al. 117 slightly modified this theory by considering the electrostatic energy
distribution in the area surrounding particles.
7.4. Polarization model
Both the EDL and water bridge mechanisms lost their physical basis when
anhydrous ER fluids were invented in 1985. The fibrillation model received much
attention again and many attempts were made to quantitatively calculate the
electrostatic polarization force between particles within a chain structure. A
thorough overview of the calculated results on the basis of various slightly different
w x
polarization models was given by Parthasarathy and Klingenberg 7 . Only the basic
concept of this model will be introduced here.
The objective of the polarization models is to relate the material parameters,
such as the dielectric properties of both the liquid and solid particles, the particle
volume fraction, the electric field strength, etc., to the rheological properties of the

whole suspension, in combination with other microstructure features e.g. a fibril-


.
lated chain . Using an idealized physical model ER system a uniform, hard
.
dielectric sphere real dielectric constant , diameter 2r dispersed in a Newto-
p
.
nian continuous medium real dielectric constant , the derived electrostatic
m
force was found to be dependent on the dielectric constant mismatch between the
w x
particle and continuous medium 9098 . A universal form could be written as:
2
2 2
. .
F sk 2r f E S 6
m
. .
where F is the electrostatic force, k a constant, f s g y 1 r g q2 and g s
r , S is a factor related to the particle microstructure. The shear modulus
p m
w x
would increase linearly with the dielectric constant ratio r 96 , indicating that
p m
a high particle dielectric constant would give a strong ER effect.
The material of an extremely high dielectric constant was thus used experimen-
tally as the solid particulate phase of the ER fluid, and a strong ER effect was
anticipated according to the derivation of the polarization model. A barium

titanate suspension BaTiO , its dielectric constant approx. 2000, depending on its
3
.
crystallization state , however, presents a surprising result: inactive under a dc field
w x w x
118 , and active after adsorbing a small amount of water 119 or being stimulated
w x
by an ac field 86,120 , suggesting that the polarization model still has much room
to improve. Obviously, the polarization model fails to describe other important ER
experimental observations, such as the rheological property dependence on the
electric field frequency and the particle conductivity.
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 25
7.5. Conduction model
Modifications were made to the polarization model because of its many limita-
tions. The conductivity mismatch between particle and liquid medium, rather than
the dielectric constant mismatch, was thought to be a dominant factor for dc and
w x
low frequency ac excitation 93,96,98 . However, the generalized polarization model
is still very poor in many aspects; for example, it could not explain why the negative
w x w x
ER effect could occur in some suspensions. Atten 101 and Foulc 102 thus
proposed a conduction model in which the ER effect was thought to be determined
by the particle-to-liquid conductivity ratio r if ) ; if - , a
p m p m p m
negative ER effect would be expected. This model was extensively developed by
w x w x
Tang 121 and Wu 122 .
The conduction model could successfully explain ER phenomena that are
unexplainable by using the polarization model. It could predict the current density,
the yield stress and the temperature dependence of the ER suspension. However,
w x
as indicated by Khusid and Acrivos 123 , the conduction model can only be used
for the situation where the suspension microstructure has been fully formed. The
conduction model only considers the particle interaction, regardless of the mi-
crostructure change after an electric field is applied. It therefore could not give an
explanation of the dynamic phenomena, such as the response time of ER fluid.
More important, some experimental results provide evidence against this mecha-
.
nism. For example, a magnesium hydroxiderpoly methylphenylsiloxane suspen-
sion should have exhibited a positive ER effect according to the conduction model,
w x
however, it displays an obvious negative ER effect 63 .
The fatal shortcoming of the polarization and conduction models is that both of
them are static, and do not take dynamic processes occurring in ER fluids into
w x
account. Khusid 123 considered dynamic events in ER fluids and examined the
effects of the conductivity on both the electric field induced particle aggregation
process and the interfacial polarization process. An excellent qualitative theory was
derived, and it is much more powerful than the polarization and conduction
models, though some discrepancies with the experimental results still exist. This is
because Khusids two presumptions are not always valid in ER fluids. Both
dispersed particles and the liquid medium were assumed to have no intrinsic
dielectric dispersion, and the variation of the applied electric field was assumed to
be very slow compared with the polarization rate.
w x
There are other models available 124,125 , however, none of them, including the
polarization and conduction model, could explain all the current ER findings. They
all suffer from a severe limitation: They cannot predict the yield stress based on
the physical properties of ER suspension components and on the operating
.
conditions field strength, temperature, frequency, etc. . They thus could not
provide a clear clue or implication on how to formulate a good ER suspension.
7.6. Dielectric loss model
w x
Hao and co-workers 114,126,127 proposed a dielectric loss model to understand
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 26
the ER mechanism on the basis of their experimental findings. Two dynamic
processes were emphasized in this model. The first step is the particle polarization
process, in which the particle dielectric constant is dominant. The second step is
particle turning, i.e. the polarized particle could have the capability to align along
.
the direction of the electric field see Fig. 9 . This step was determined by the
particle dielectric loss. The second step is the most important one, which distin-
guishes the ER particle from non-ER particle. In other words, both the ER particle
and non-ER particle could be polarized under an electric field, however, the ER
particle could re-orientate along the electric field direction, building the fibrillated
bridges between two electrodes. The non-ER particle does not have such an ability.
The possible reason is that the ER particle has a comparatively high dielectric loss
tangent, approximately 0.1 at 1000 Hz, which could generate a large amount of
bounded surface charge. The particle could turn even under a weak electric field
w x
114 . The ER particle turning under an electric field has ever been detected by
w x
using X-ray diffraction 128,129 . The non-ER particle could not gain enough
surface charge due to its low dielectric loss. Although they still could be polarized,
the total inter-particle force would be cancelled out owing to the diversity of
.
particle dipole vectors see Fig. 9 .
For the purpose of understanding why a large dielectric loss is experimentally
Fig. 9. Schematic illustration of the ER particles and the non-ER particle behaviors before and after
. . an external electric field is applied. a ER particle; b non-ER particle. Reprinted with permission
w x from Ref. 114 . Copyright 1998 American Chemical Society.
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 27
w x
necessary for the ER response, a theoretical approach was developed 126 . Since

the ER fluid changes from a liquid state the particle randomly distributed in the
. .
liquid medium to a solid state fibrillated chain of bct lattice , the entropy of the
w x
ER system may dramatically decrease. Based on this fact, Hao et al. 126
theoretically came to the conclusion that the particle dielectric loss tangent
maximum value should be larger than 0.10, which agrees well with the empirical
w x
criteria put forward before 74 . The criteria for the positive ER effect and negative
ER effect were derived on the assumption that the interfacial polarization would
contribute mainly to the ER effect. Assuming that the static dielectric constants of
the liquid medium and the particle are , and , respectively, and T is
sm sp
d
sp
temperature, a strong positive ER effect will occur if )4 and )0. A
sp sm
dT
d
sp
.
weak or no ER effect is expected if )4 , -0, and Eq. 7 is satisfied;
sp sm
dT
d
sp
.
A negative ER effect is anticipated if )4 , -0, and Eq. 8 is satisfied.
sp sm
dT
3
2 2 2
. .
d rdT 1 q 3 2 q q54 y1.5 y
.
sp sm sp m sp sm sp sm
.
- 7
2 2
.
d rdT 27 y4
sm sm sp sm
3
2 2 2
. .
d rdT 1 q 3 2 q q54 y1.5 y
.
sp sm sp m sp sm sp sm
.
) 8
2 2
.
d rdT 27 y4
sm sm sp sm
d
sp
If -4 , physically would unlikely be less than zero, thus a weak or
sp sm
dT
negative ER effect would become possible at this condition. A recent experimental
w x
result 130 shows that merely changing the particle volume fraction could switch
the suspension from showing a positive ER effect to a negative ER effect, which is
possible, according to the above criteria.
A yield stress equation was derived on the basis of the dielectric loss mechanism
described above. Assuming that only interfacial polarization would contribute to
the ER effect and the ER particle would form the bct structure under an electric
w x
field, a yield stress equation could be expressed as 127 :
126
2
E
2
s
y
.
4 1 y y18

2 2 2
. . .
3 q y 2 2 T y1.5 y 1 q y2
sm m sm sm
y q q
3 3
. .
2 q 2 q
~
=
2
. .
T p y4 q2
p sm
3

.
2 q
.
9
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 28
where is the yield stress, and are the liquid medium and solid particle
y m p
w .
linear expansion coefficient, respectively, s r , p s n y1 p y n q
sp sm s
y 1 y1
sp p
. x
2 p , n is a constant, between 4 and 19.7, p s , and p s .
s
q 2 q2
sp p
.
Eq. 9 obviously indicates that the yield stress of an ER fluid would increase
with the square of the applied electric field, the particle volume fraction and the
dielectric constant of the liquid medium, which agrees very well with the previous
w x
experimental results 69,82,88 . If p is a positive value, the yield stress thus will
increase with increasing, as the numerator increases much faster than the
denominator with , which is consistent with the prediction given by the polariza-
w x
tion model 79,96 . However, if p is a negative value, then the yield stress will
decrease with increasing, which can not be explained by the polarization model,
however, it was experimentally observed. Furthermore, for some extreme cases, the
yield stress would become negative, leading to a negative ER effect. The parameter
p only becomes positive when the dielectric loss tangent of the dispersed solid
material is larger than 0.1, as we demonstrated before both experimentally and
w x
theoretically 74,126,131 . This requirement now is included in our present yield
stress expression.
The normalized yield stress dependence on the particle volume fraction in a
.
silicone oil ER system can be predicted by using Eq. 9 and is shown in Fig. 10
d
sp
under the assumption that the particle-to-oil dielectric ratio s10 and s 0.4.
dT
When the particle volume fraction is small, the yield stress increases slightly with
the particle volume fraction increasing, which is in a good agreement with the
w x
experimental results 69,88 . When the particle volume fraction is approximately
0.35, the yield stress dramatically increases, indicating that a critical volume
fraction may exist in the ER suspension. The critical volume fraction value given by
the present yield stress equation is quite close to the previous data, approximately
w x
0.37, derived from the rheological studies 82 , and was calculated to be approxi-

mately 0.4 with the aid of Florys gelation theory and the percolation concept see
.
Fig. 7 .
The influence of the particle-to-oil dielectric constant ratio, , on the yield stress
.
of an ER fluid may also be computed on the basis of Eq. 9 . The numerical
relationship between the yield stress and the particle-to-oil dielectric constant ratio
of the silicone oil ER system is showed in Fig. 11. It clearly shows that at low
particle-to-oil dielectric constant ratio the yield stress increases dramatically with
increasing, however, the yield stress gradually levels off after is more than 60.
The sharp increase of the yield stress that takes place in the range of -50,
indicates that materials of static dielectric constant of approximately 150 might
potentially display the best ER effect. Materials of large dielectric constant may
not always display a good ER effect if other properties are not beneficial to the ER
effect. It should be noted that the conclusion above is derived under the assump-
tion that the parameter p is positive. If p is negative, the larger particle-to-oil
dielectric constant ratio would not generate a larger yield stress. The best example
would be the BaTiO rsilicone oil system.
3
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 29
Fig. 10. The normalized yield stress, rE
2
=10
2
vs. particle volume fraction at the particle-to-oil
y
. dielectric constant ratio s 10, and d rdT s0.4. This is computed on the basis of Eq. 9 . Reprinted
sp
w x with permission from Ref. 127 . Copyright 2000 American Chemical Society.
w x .
As indicated by Hao et al. 127 , Eq. 9 could also well explain why the ER
effect would approximately increase with the difference between the dielectric
constants below and above the relaxation frequency, which was experimentally
w x .
observed by Ikazaki and Kawai 77,78 . Eq. 9 explains why the yield stress peaks
for the particle conductivity, and for different particulate material, the yield stress
w x .
peaks at different conductivity 69,74 . Eq. 9 also explains why the yield stress
w x
goes through a maximum value as temperature increases 132134 , and why the
polarization rate controls the ER effect. It accounts for why the slower polarization
rate will give a stronger ER effect, etc. In other words, the yield stress equation
based on the dielectric loss model could give a very good explanation of almost all
the currently observed ER phenomena.
The yield stress equation developed by Hao provides a rational relationship
between the yield stress and the physical parameters of ER fluids, thus offering a
.
clear indication of how to formulate a high performance ER fluid. From Eq. 9 ,
one should use a liquid medium and a solid material that are of high dielectric
w . x
2 2
Fig. 11. The numerical relationship between the normalized yield stress, 4 1 y y18 r E ,
y
. and the particle-to-oil dielectric constant ratio, , of the silicone oil ER system, based on Eq. 9 at
. w x room temperature . Reprinted with permission from Ref. 127 . Copyright 2000 American Chemical
Society.
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 30
.
constant large dielectric constant ratio of the solid material to the liquid , and
keep the particle-to-liquid dielectric constant ratio at approximately 50. In addi-

tion, the solid material must have a large dielectric loss tangent at least its
.
maximum value must be larger than 0.1 for keeping the parameter p positive. For
the purpose of increasing the p value as much as possible, a slow relaxation time
and suitable conductivity are also essential.
.
In spite of its successes, Eq. 9 has some deficiencies. Since our equation is
derived from the internal energy change under a static electric field, it is unsuitable
to account for the frequency dependence of the yield stress. The internal energy
change under an oscillating electric field is complicated.
8. Application of the ER fluid
The ER fluid could be readily and rapidly converted between a liquid and a solid
state continuously and reversibly via an electric field. It thus could be used as a
mechanical electronic interface for transferring and controlling mechanical move-
ment. Many ER devices have been reported and patented, including clutches
w x w x w x w x
17,135,136 , brakes 137 , damping devices 138 , actuators 139 , fuel injection
w x w x w x
valves 140 , hydraulic valves 17,39 and robotic controlling systems 39,140 .
Among these devices, the ER damping devices have received intensive attention,
because they do not require the ER fluid to have a very high yield stress or a very
w x
wide temperature range. Many damping devices, such as shock absorbers 141 ,
w x w x w x
robot arms 142 , engine mounts 143 , bearing dampers 144 , seismic controlling
w x
frame structures 145 , etc., have been proposed. In addition, an ER fluid could
w x w x
also be used for a photonic crystal 146 , light shutter 147 , mechanical polisher
w x w x w x w x
148 , display 149 , ink jet printer 150,151 , human muscle stimulator 152 ,
w x
mechanical sensor or seismograph 113 , etc. More ER devices will surely be
developed in the future.
.
The ER devices have not yet been commercialized. The major problems are: a
.
the yield stress is not high enough; b the working temperature range is not wide
enough, because the ER effect decreases sharply once the temperature is larger
.
than 100C; c the ER suspension stability against sedimentation is very poor.
.
Once segregation occurs, no ER effect is available; and d the ER fluid sometimes
malfunctions once contaminated. The particle sedimentation problem is regarded
as the biggest one at the current stage. Many efforts were made to improve the
w x
sedimentation properties 153,154 . The ER technology will boom if these limita-
tions are overcome.
9. Summary and outlook
An electrorheological fluid is a smart material of vast application potential. The
heterogeneous ER fluid has been studied for a long time, and there is a large body
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 31
of literature on the subject. The main problem for the heterogeneous ER fluid is
the sedimentation of particles. The homogeneous ER fluid was developed quite
recently. It is believed to be promising, compared with the heterogeneous ER fluid.
Anhydrous ER fluids could work in a wide temperature range, and are superior to
moist ER fluids. The electric field strength, frequency of electric field, the particle
conductivity, the particle dielectric properties, particle volume fraction and temper-
ature, as well as water content are variables that can affect the ER effect
substantially. The electric-field-induced phase transition occurs in ER fluids, and
the crystal that is formed is of bct lattice structure. The percolation transition also
occurs as the particle volume fraction increases to 37%. The dc current passing
through an ER fluid may oscillate with the applied mechanical field, which can be
explained by the Quasi-1d-VRH conduction model. Four polarizations including
the interfacial, the Debye, the atomic and the electronic polarizations may take
place in ER fluids, and the interfacial polarization is found to contribute to the ER
effect. The electric double layer and water bridging models were proposed earlier,
however, they could only be used for moist ER systems. An idealized polarization
model and a generalized polarization model, as well as a similar conduction model
were proposed, but suffered from a number of limitations. The dielectric loss
model, giving the best description on the mechanism of ER effect so far, can
explain almost all currently available experimental data. It offers clear guidance on
how to formulate a high performance ER fluid for industrial purposes. The yield
stress equation derived on the basis of this model is powerful, but it does not
explain the frequency dependence of the ER effect, indicating that much more
work remains to be accomplished.
Future studies will continue to emphasize the development of high performance
ER fluids with a strong ER effect and no sedimentation trouble. Physical models
for the mechanism of ER effect will be refined and the design of the ER fluid
devices will be optimized. A breakthrough in ER fluid preparation will boost this
field significantly and will definitely speed up the commercialization of ER devices.
The ER sensor, the ER damping device, the ER ink jet printer and ER polisher for
semi-conductive industrial parts will become the front runners, because they will
not require the ER fluid to work in extreme conditions, and they have vast
industrial needs.
Acknowledgements
I am very grateful to those with whom I worked before for their contribution to
this work. A special debt of gratitude goes out to Dr Yuanze Xu, Dr Mikio
Nakamura and Dr Fumikazu Ikazaki. I would also like to thank the National
Science Foundation of China, the Toho University, The Agency of Science and
Technology of Japan for financial support. I thank Dr Elizabeth McCandish for
reading this paper.
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 32
References
w x . 1 W.M. Winslow, J. Appl. Phys. 20 1949 1137.
w x . 2 H. Block, J.P. Kelley, J. Phys. D.: Appl. Phys. 21 1988 1661.
w x . 3 A.P. Gast, C.F. Zukoski, Adv. Colloid Interface Sci. 30 1989 153.
w x . 4 T.C. Jordan, M.T. Shaw, IEEE Trans. Electr. Insul. 24 1989 849.
w x . 5 T.C. Halsey, Science 23 1992 761.
w x . 6 C.F. Zukoski, Annu. Rev. Mater. Sci. 23 1993 45.
w x . 7 M. Parthasarathy, D.J. Klingenberg, Mater. Sci. Eng. R17 1996 7.
w x . 8 F.E. Filisko, D.R. Gamota, ASME 153 1992 5.
w x . 9 A.W. Duff, Phys. Rev. 4 1896 23.
w x . 10 E.N. Andrade, C. Dodd, Nature 143 1939 26.
w x . 11 E.N. Andrade, C. Dodd, R. Soc. Lond. Proc. A 187 1946 296.
w x . 12 E.N. Andrade, C. Dodd, R. Soc. Lond. Proc. A 204 1951 449.
w x . 13 E.N. Andrade, J. Hart, R. Soc. Lond. Proc. A 225 1954 463.
w x . 14 L. Onsager, R.M. Fuoss, J. Phys. Chem. 36 1932 2689.
w x . 15 M. von Smoluchowski, Kolloid-Z 18 1916 190.
w x . 16 B.E. Conway, A. Dobry-Duclaux, in: F.R. Eirich Ed. , Rheology: Theory and Applications, 3,
Academic Press, NY, 1960.
w x 17 W.M. Winslow, U.S. Patent 2417850, 1947.
w x . 18 Yu.F. Deinega, G.V. Vinogradov, Colloid J. 24 1962 570.
w x . 19 D.L. Klass, T.W. Martinek, J. Appl. Phys. 38 1967 67.
w x . 20 D.L. Klass, T.W. Martinek, J. Appl. Phys. 38 1967 75.
w x . 21 H. Uejima, Jpn. J. Appl. Phys. 11 1972 319.
w x . 22 Yu.F. Deinega, G.V. Vinogradov, Rheol. Acta 23 1984 636.
w x 23 J.E. Stangroom, I. Harness, GB Patent 2153372, 1985.
w x . 24 J.E. Stangroom, Phys. Techn. 14 1983 290.
w x 25 H. Block, J.P. Kelly, GB Patent 2170510, 1985.
w x 26 D.G. Bytt, GB Patent 2189803, 1987.
w x 27 F.E. Filisko, W.E. Amstrong, U.S. Patent 4744914, 1988.
w x . 28 J.J. Wysocki, J. Adama, W. Hass, J. Appl. Phys. 40 1969 3865.
w x . 29 T. Honda, T. Sasada, K. Kurokawa, Jpn. J. Appl. Phys. 17 1978 1525.
w x . 30 I.K. Yang, A.D. Shine, J. Rheol. 36 1992 1079.
w x 31 A. Inoue, S. Maniwa, European Patent 478034A1, 1992.
w x . 32 K. Negita, J. Phys. Chem. 105 1996 7837.
w x . 33 N. Yao, A.M. Jamieson, Macromolecules 30 1997 5822.
w x . 34 K. Tajiri, K. Ohta, T. Nagaya, H. Orihara, J. Rheol. 41 1997 335.
w x . 35 X. Pan, G.H. McKinley, J. Colloid Interface Sci. 195 1997 101.
w x 36 H. Kimura, A. Aikawa, Y. Masubuchi, J. Takimoto, K. Koyama, T. Uemura, J. Non-Newtonian
. Fluid Mech. 76 1998 199.
w x . 37 J. Ha, S. Yang, J. Rheol. 44 2000 235.
w x . 38 G.G. Petrzhik, O.A. Chertkova, A.A. Trapeznikov, Dokl. Akad. Nauk USSR 253 1980 173.
w x 39 W.M. Winslow, U.S. Patent 2661596, 1953.
w x 40 W.M. Winslow, U.S. Patent 3047507, 1962.
w x 41 M. Kanbara, H. Tomizawa, European Patent 0361931, 1990.
w x 42 D. Gillies, L. Sutcliffe, P. Bailey, GB Patent 2219598, 1989.
w x 43 F.E. Filisko, W.F. Armstrong, European Patent 0313351, 1989.
w x 44 F.E. Filisko, W.F. Armstrong, European Patent 0265252, 1988.
w x 45 J. Goossens, G. Oppermann, US Patent 4702855, 1987.
w x 46 H. Tomizawa, M. Kanbara, N. Yoshimura, J. Mitsui, H. Hirano, European Patent 0342041, 1989.
w x 47 J.D. Carlson, US Patent 4772407, 1988.
w x 48 H. Block, J.P. Kelley, US Patent 1501635, 1987.
w x 49 W. Podszun, R. Bloodworth, G. Oppermann, US Patent 5503763, 1996.
w x 50 J.W. Pialet, US Patent 5558811, 1996.
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 33
w x 51 C.P. Bryant, K. Lal, J.W. Pialet, US Patent 5435932, 1995.
w x . 52 C.J. Gow, C.F. Zukoski, J. Colloid Interface Sci. 126 1990 175.
w x 53 J.W. Pialet, C.P. Bryant, K. Lal, European Patent 0562067, 1997.
w x 54 J.D. Carlson, K.D. Weiss, J.E. Bares, WO Patent 9312192, 1993.
w x 55 J.E. Stangroom, GB Patent 1570234, 1980.
w x 56 J.E. Westhaver, US Patent 3970573, 1976.
w x . 57 Y. Xu, R. Liang, J. Rheol. 35 1991 1355.
w x 58 S. Endo, European Patent 0964053, 1999.
w x 59 D. Prick, H. Grasshoff, H. kohnz, P. Finmans, T. Carstensen, D. Jakubik et al., US Patent
5800731, 1998.
w x 60 K. Haji, M. Sasaki, M. Matsuno, European Patent 0798368, 1997.
w x . 61 C. Boissy, P. Atten, J-N. Foulc, J. Electrostatics 35 1995 13.
w x . . 62 C.W. Wu, H. Conrad, J. Rheol. 41 2 1997 267.
w x . 63 J. Trlica, Q. Quadrat, P. Bradna, V. Pavlinek, P. Saha, J. Rheol. 40 1996 943.
w x . 64 F.E. Filisko, in: R. Tao Ed. , Proc. Intern. Conf. On Electrorheological Fluids, World Sci, 1992,
p. 116.
w x 65 L. Carreira, V.S. Mihajlov, US Patent 3553708, 1971.
w x . 66 Y. Komoda, T.N. Rao, A. Fujishima, Langmuir 13 1997 1371.
w x . 67 Y. Komoda, N. Sakai, T.N. Rao, D.A. Tryk, A. Fujishima, Langmuir 14 1998 1081.
w x 68 J.E. Stangroom, GB Patent 2119392, 1983.
w x . 69 H. Block, J.P. Kelley, A. Qin, T. Watson, Langmuir 6 1990 6.
w x . 70 T. Hao, J. Colloid Interface Sci. 206 1998 240.
w x . 71 K.W. Wagner, Arch. Electrotechnik 2 1914 371.
w x . 72 S.O. Morgan, Trans. Am. Electrochem. Soc. 65 1934 109.
w x . 73 R.W. Sillars, JIEE 80 1937 378.
w x . 74 T. Hao, Z. Xu, Y. Xu, J. Colloid Interface Sci. 190 1997 334.
w x . 75 K.D. Weiss, D.A. Nixon, J.D. Carlson, A.J. Margida, Polym. Preprints 35 1994 325.
w x . 76 Yu. F. Deinega, K.K. Popko, N. Ya Kovganich, Heat-Transfer-Sov. Res. 10 1978 50.
w x 77 A. Kawai, K. Uchida, K. Kamiya, A. Gotoh, S. Yoda, K. Urabe et al., Int. J. Mod. Phys. B 10
. 1996 2849.
w x 78 F. ikazaki, A. Kawai, T. Kawakami, K. Edamura, K. Sakuri, H. Anzai et al., J. Phys. D: Appl.
. Phys. 31 1998 336.
w x . 79 D.J. Klingenberg, C.F. Zukoski, Langmuir 6 1990 15.
w x . 80 A.M. Kraynik, R.T. Bonnecaze, J.F. Brady, in: R. Tao Ed. , Proc. Intern. Conf. On ER fluids,
World Scientific, 1992, p. 59.
w x . 81 G. Bossis, E. Lemaire, O. Volkova, H. Clercx, J. Rheol. 41 1997 687.
w x . 82 T. Hao, Y. Chen, Z. Xu, Y. Xu, Y. Huang, Chin. J. Polym. Sci. 12 1994 97.
w x . 83 U.Y. Treasurer, F.E. Filisko, L.H. Radzilowski, J. Rheol. 35 1991 1051.
w x . 84 H. Conrad, A.F. Sprecher, Y. Choi, Y. Chen, J. Rheol. 35 1991 1393.
w x . 85 A.V. Lykov, Z.P. Shulman, R.G. Gorodkin, A.D. Matsepuro, J. Eng. Phys. 18 1970 979.
w x . 86 T. Garino, A. Adolf, B. Hance, in: R. Tao Ed. , Proc. Intern. Conf. on ER Fluids, World
Scientific, 1992, p. 167.
w x . 87 Y.D. Kim, D.J. Klingenberg, J. Colloid Interface Sci. 183 1996 568.
w x . 88 L. Marshall, C.F. Zukoski, J.W. Goodwin, J. Chem. Soc. Faraday Trans. 85 1989 2785.
w x . 89 D.M. Heyes, J.R. Melrose, Mol. Sim. 5 1990 293.
w x . 90 P.M. Adriani, A.P. Gast, Phys. Fluid 31 1988 2757.
w x . 91 D.J. Klingenberg, S. van Frank, C.F. Zukoski, J. Chem. Phys. 94 1991 6160.
w x . 92 Y. Chen, A.F. Sprecher, H. Conrad, J. Appl. Phys. 70 1991 6796.
w x . 93 R.A. Anderson, Langmuir 10 1994 2917.
w x . 94 H.H. Clarx, G. Bossis, Phys. Rev. E 48 1993 2721.
w x . 95 R. Tao, Q. Jiang, Phys. Rev. Lett. 73 1994 2721.
w x . 96 L.C. Davis, Appl. Phys. Lett. 60 1992 319.
w x . 97 L.C. Davis, J. Appl. Phys. 73 1993 680.
w x . 98 L.C. Davis, J. Appl. Phys. 72 1992 1334.
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 34
w x . 99 M.J. Chrzan, J.P. Coulter, Int. J. Mod. Phys. B 6 1992 2651.
w x . 100 R.T. Bonnecaze, J.F. Brady, J. Rheol. 36 1992 73.
w x . 101 P. Atten, J-N. Foulc, N. Felici, Int. J. Mod. Phys. B 8 1994 2731.
w x . 102 J-N. Foulc, P. Atten, N. Felici, J. Electrostatic 33 1994 103.
w x . 103 Z. Cheng, W.B. Russel, P.M. Chalkin, Nature 401 1999 893.
w x . 104 P.N. Pusey, W. van Megan, Nature 320 1986 340.
w x . 105 L.V. Woodcok, Nature 385 1997 141.
w x . 106 W.L. Vos, M. Megens, C.M. Kats, P. Bosecke, Langmuir 13 1997 6004.
w x . 107 H. Miguez, F. Meseguer, C. Lopez, A. Mifsud, J.S. Moya, L. Vazquez, Langmuir 13 1997 6009.
w x . 108 R. Tao, J.M. Sun, Phys. Rev. Lett. 67 1991 398.
w x . 109 T. Chen, R.N. Zitter, R. Tao, Phys. Rev. Lett. 68 1992 2555.
w x . 110 U. Dassanayake, S. Fraden, A. van Blaaderen, J. Chem. Phys. 112 2000 3851.
w x 111 R. Zallen, The Physics of Amorphous Solids, Chapter 6, Wiley & Sons, New York, 1983.
w x . 112 T. Hao, Y. Xu, J. Colloid Interface Sci. 181 1996 581.
w x . 113 T. Hao, J. Phys. Chem. 102 1998 1.
w x . 114 T. Hao, A. Kawai, F. Ikazaki, Langmuir 14 1998 1256.
w x 115 H. Frohlich, Theory of Dielectrics, Clarendon Press, Oxford, UK, 1958.
w x . 116 T. Hao, A. Kawai, F. Ikazaki, J. Colloid Interface Sci. submitted .
w x . 117 H. See, H. Tamura, M. Doi, J. Phys. D: Appl. Phys. 26 1993 746.
w x . 118 Y. Otsubo, K. Watanabe, J. Soc. Rheol. Jpn. 18 1990 111.
w x . 119 C.F. Zukoski, Annu. Rev. Mater. Sci. 23 1993 45.
w x . 120 P.J. Rankin, D.K. Klingenberg, J. Rheol. 42 1998 639.
w x . 121 X. Tang, C. Wu, H. Conrad, J. Rheol. 39 1995 1059.
w x . 122 C. Wu, H. Conrad, J. Phys. D: Appl. Phys. 29 1996 3147.
w x . 123 B. Khusid, A. Acrivos, Phys. Rev. E 52 1995 1669.
w x . 124 H. See, T. Saito, Rheol. Acta 35 1996 233.
w x . 125 H. Ma, W. Wen, W.Y. Tam, P. Sheng, Phys. Rev. Lett 77 1996 2499.
w x . 126 T. Hao, A. Kawai, F. Ikazaki, Langmuir 15 1999 918.
w x . 127 T. Hao, A. Kawai, F. Ikazaki, Langmuir 16 2000 3058.
w x . 128 W. Wen, K. Lu, Appl. Phys. Lett. 68 1996 1046.
w x . 129 Y. Lan, X. Xu, S. Men, K. Lu, Phys. Rev. E 60 1999 4336.
w x . 130 V. Pavlinek, P. Saha, O. Quadrat, J. Stejskal, Langmuir 16 2000 1447.
w x . 131 T. Hao, Appl. Phys. Lett. 70 1997 1956.
w x . 132 T. Hao, H. Yu, Y. Xu, J. Colloid Interface Sci. 184 1996 542.
w x . 133 H. Conrad, Y. Li, Y. Chen, J. Rheol. 39 1995 1041.
w x . 134 P. Goonon, J.-N. Foulc, J. Appl. Phys. Lett. 87 2000 3563.
w x . 135 D.L. Hartsock, R.F. Novak, G.J. Chaundy, J. Rheol. 35 1991 1305.
w x . 136 W.A. Bullough, A.R. Johnson, R. Tozer, J. Makin, in: M. Nakano, K. Koyama Eds. , Proc.
Intern. Conf. on Electrorheological Fluids, World Sci, 1998, p. 623.
w x . 137 K. Shimada, T. Fujita, M. Iwabuchi, M. Nishida, K. Okui, in: M. Nakano, K. Koyama Eds. , Proc.
Intern. Conf. on Electrorheological Fluids, World Sci, 1998, p. 680.
w x . 138 D.A. Brooks, in: M. Nakano, K. Koyama Eds. , Proc. Intern. Conf. on Electrorheological Fluids,
World Sci, 1998, p. 689.
w x . 139 E. Wendt, K.W. Busing, in: M. Nakano, K. Koyama Eds. , Proc. Intern. Conf. on Electrorheo-
logical Fluids, World Sci, 1998, p. 780.
w x 140 S.H. Nicholas, US Patent 5019119, 1991.
w x . 141 B. Khusid, A. Acrivos, Y. Khodorkovsky, M. Beltran, in: M. Nakano, K. Koyama Eds. , Proc.
Intern. Conf. on Electrorheological Fluids, World Sci, 1998, p. 705.
w x . 142 J. Furusho, N. Takeue, G. Zhang, M. Sakaguchi, in: M. Nakano, K. Koyama Eds. , Proc. Intern.
Conf. on Electrorheological Fluids, World Sci, 1998, p. 713.
w x . 143 Y.S. Jeon, Y.T. Choi, C.C. Cheong, M.S. Suh, in: M. Nakano, K. Koyama Eds. , Proc. Intern.
Conf. on Electrorheological Fluids, World Sci, 1998, p. 721.
w x . 144 J.M. Vance, D. Ying, J. Eng. Gas Turb. Power 122 2000 337.
w x . 145 Y.L. Xu, W.L. Qu, J.M. Ko, Earthquake Eng. Struct. Dyn. 29 2000 557.
( ) T. Hao rAdances in Colloid and Interface Science 97 2002 135 35
w x . 146 R. Tao, in: M. Nakano, K. Koyama Eds. , Proc. Intern. Conf. on Electrorheological Fluids,
World Sci, 1998, p. 811.
w x . 147 K. Yamaguchi, B. Jeyadevan, T. Fujita, A. Nishihara, in: M. Nakano, K. Koyama Eds. , Proc.
Intern. Conf. on Electrorheological Fluids, World Sci, 1998, p. 819.
w x 148 Y. Akagami, S. Nishimura, Y. Ogasawana, T. Fujita, B. Jeyadevan, K. Nuri, K. Itoh, in: M.
. Nakano, K. Koyama Eds. , Proc. Intern. Conf. on Electrorheological Fluids, World Sci, 1998, p.
803.
w x 149 K. Akashi, H. Anzai, K. Edamura, Y. Otsubo, European Patent 0697615, 1996.
w x 150 S. Sohn, US Patent 5510817, 1996.
w x 151 R.W. Gundlach, E.G. Rawson, US Patent 6048050, 2000.
w x 152 C. Norbert, R. Axel, DE Patent 19830559, 2000.
w x . 153 M. Qi, M.T. Shaw, J. Appl. Polym. Sci. 65 1997 539.
w x . 154 H. Xie, J. Guan, J. Guo, J. Appl. Polym. Sci. 64 1997 1641.

You might also like