Smart Material Structure

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Smart Mater. Struct. 5 (1996) 607614.

Printed in the UK
A model of the behaviour of
magnetorheological materials
Mark R Jolly, J David Carlson and Beth C Mu noz
Lord Corporation, Thomas Lord Research Center, 405 Gregson Drive, Cary, North
Carolina 27511, USA
Received 28 May 1996, accepted for publication 19 July 1996
Abstract. Magnetorheological materials are a class of smart materials whose
rheological properties may be rapidly varied by application of a magnetic eld.
These materials typically consist of micron-sized ferrous particles dispersed in a
uid or an elastomer. A quasi-static, one-dimensional model is developed that
examines the mechanical and magnetic properties of magnetorheological materials.
This model attempts to account for magnetic non-linearities and saturation by
establishing a mechanism by which magnetic ux density is distributed within the
composite material. Experimental evidence of the viscoelastic behaviour and
magnetic properties of magnetorheological uids and elastomers suggests that the
assumptions made in the model development are reasonable. It is shown that the
model is semi-empirical in that it must be t to the experimental data by adjusting a
parameter that accounts for unmodelled magnetic interactions.
1. Introduction
Magnetorheological (MR) materials consist of micron
size (typically 3 to 5 microns) magnetically permeable
particles suspended in a non-magnetic medium. Upon
application of a magnetic eld, the rheological properties
of these materials are rapidly and reversibly altered. The
mechanism responsible for this bulk effect is the induced
magnetic interaction of particles within the matrix. In
the case of MR uids, this mutual interaction amongst
the particles causes the formation of chain-like structures
aligned roughly parallel to the applied magnetic eld. The
conguration and rigidity of this structure will depend upon
several factors including the strength and distribution of
the applied magnetic eld. A nite stress must develop to
yield these chain structures. Hence, MR uids are often
modelled as Bingham plastics with a eld dependent yield
stress (Spasojevic et al 1974). Yield stresses of nearly
100 kPa have been reported in commercially available
uids (Lord 1994). MR solids have also been developed
by incorporating magnetically permeable particles into
elastomer media (Rigbi and Jilken 1983, Mu noz and Jolly
1995, Jolly et al 1996). Typically, magnetic elds are
applied to the elastomer during crosslinking such that
particle chain structures form and become locked in place
upon nal cure (see gure 1). At low strains (< 2%),
these elastomers exhibit a eld dependent modulus. A eld-
induced change in modulus of up to 40% has been reported
(Jolly et al 1995).
MR uids are beginning to nd commercial success
in applications where controlled energy dissipation is
required. Such applications include controllable dampers
for primary and secondary vehicle suspensions (Carlson
et al 1994, Carlson et al 1995) and brakes and clutches
for exercise equipment (Carlson et al 1995, Gentry et
al 1995, Anon 1995, Chase 1996). MR elastomers
will ultimately nd use in applications where stiffness or
resonance tunability is sought (Jolly et al 1996, Kasahara
et al 1994). Other applications may also make use of
the unique mechanical, magnetic and thermal anisotropic
properties of these elastomer composites (Bowen et al
1995).
Most models of MR material behaviour are based
on magnetic dipole interactions between adjacent particles
within a particle chain (Harpavat 1974, Spasojevic et
al 1974). These interparticle interactions are then
averaged over the entire sample to yield a model of the
bulk magnetorheological effect. Many such models are
borrowed from models developed to understand similar
interparticle interactions within electrorheological (ER)
uids (e.g., Block and Kelly 1988 and references therein,
Gast and Zukoski 1989). Still a weakness unique to MR
material modelling is an understanding of how magnetic
ux density distributes itself within the particle network.
Such an understanding needs to take into account magnetic
nonlinearities and the effects of particle saturation. There
has been recent interest in this problem using nite
element approaches (Ginder and Davis 1994) and analytical
approximations (Carlson 1991, Ginder et al 1995, Jolly et
al 1996).
In this paper, the simple dipole model of particle energy
interaction is reviewed. From this we obtain a model of
the MR effect as a function of particle magnetization. In
particular, it is seen that MR material stress is quadratically
related to particle magnetization. This theory is then
augmented with an approximate model of the relationship
0964-1726/96/050607+08$19.50 c _ 1996 IOP Publishing Ltd 607
M R Jolly et al
Figure 1. SEM images of magnetic particles in a polymeric
elastomer matrix; (upper) randomly dispersed particles,
(lower) particles aligned by applied magnetic eld. The
width of the images is 70 microns.
between particle magnetization and average composite ux
densitya quantitiy that can be measured. This model
takes into account the gradual saturation of the permeable
particles as applied eld increases. This theory is then
compared to experimental data on MR uids and MR
elastomers.
2. Theory
The present model examines the magnetomechanical
properties of parallel chains of magnetically permeable
spherical particles. In particular, a magnetic eld is
applied parallel to the chains and the shear stress of
the chains caused by interparticle magnetic force is
modelled. It is assumed that if these particle chains are
then considered within a viscoelastic material, then the
viscoelastic properties of the composite are the sum of
the viscoelastic properties of the composite with no eld
applied (which are not modelled here) and the elastic/plastic
properties induced by interparticle magnetic force.
The model is quasi-static, therefore the dynamic effects
of the magnetic eld as well as inertial and rate-dependent
effects of the particles are ignored. The particles are
assumed to be uniform, homogenous spheres that can
be magnetically modelled as identical induced dipole
moments. It is also assumed that particles are aligned
in perfect chains (albeit with gaps between particles), and
that quasi-static shear strains and associated stresses are
uniformly distributed over the length of each particle chain.
2.1. The relationship between rheology and particle
polarization
Figure 2 shows two adjacent particles (dipoles) in a chain
along with the direction of applied magnetic eld. The
interaction energy of the two dipoles of equal strength [m[
and direction is (Rosensweig 1985)
E
12
=
[m[
2
(1 3 cos
2
)
4
1

0
[r[
3
=
[m[
2

1 3
r
2
0
r
2
0
x
2

4
1

0
(r
2
0
x
2
)
3/2
(1)
where
1
is the relative permeability of the medium and
where some basic trigonometric identities have been used
in the second equality. By dening the scalar shear strain
of the particle chain as = x/r
0
, the interaction energy
can be written as
E
12
=
(
2
2)[m[
2
4
1

0
r
3
0
(
2
1)
5/2
(2)
Now it is assumed that the particles are aligned in
long chains and that there is only magnetic interaction
between adjacent particles within the chain, i.e., there
are no multipole interactions. Then the total energy
density (energy per unit volume) associated with the one-
dimensional shear strain can be calculated by multiplying
the particle-to-particle energy, given by equation (2), by the
total number of particles and dividing by the total volume:
U =
3(
2
2)[m[
2
2
2

0
d
3
r
3
0
(1
2
)
5/2
(3)
where is the volume fraction of particles in the composite
and d is the particle diameter. The stress induced by the
application of a magnetic eld can be computed by taking
the derivative of inter-particle energy density with respect
to the scalar shear strain:
=

U =
9(4
2
)[m[
2
2
2

0
d
3
r
3
0
(1
2
)
7/2
. (4)
Next, it is noted that the average particle polarization J
p
is
the dipole moment magnitude per unit particle volume, i.e.
[m[ = J
p
V
i
, where J
p
has SI units of Tesla:
=
(4
2
)J
2
p
8
1

0
h
3
(1
2
)
7/2
(5)
608
A magnetorheological material behaviour model
Figure 2. (a) Geometry of two particles of diameter d
within a particle chain. Adjacent chains are a distance y
apart and h = r
0
/d. (b) Magnetic interaction of the two
particles modelled as dipole moments m and sheared with
respect to one another.
where we have dened h = r
0
/d which is an indication of
the gap between particles in a chain. This is illustrated in
gure 2.
Note that the stress-strain relationship given in
equation (5) is independent of particle size. The size of
the gap between the particles in a chain can be seen to be a
very important parametermagnetic eld induced stress is
inversely proportional to this gap cubed. Say, for example,
that the gap between particles is equal to one quarter of a
particle diameter (h = 1.25). Gaps of this size will cause a
factor of two reduction in the magnetically induced stress.
From gure 2, simple geometry indicates that adjacent
particle chains will touch (y = d) when the particle gap is
such that h = /6. And particles will be evenly spaced
within the matrix (y = dh) when h
3
= /6.
The yield stress of the particle chains occurs at the strain
for which stress is maximum. This is found by setting the
rst derivative of stress with respect to quasi-static shear
strain to zero:

=
(4
4
27
2
4)J
2
p
8
1

0
h
3
(1
2
)
9/2
= 0 (6)
from which yield strain and corresponding yield stress can
be found:

y
=

27

665
2
3/2
= 0.389 (7a)

y
=
0.1143J
2
p

0
h
3
. (7b)
Experimental evidence has indicated that the actual particle
chain yield strain is 0.010.02 in MR elastomers (Jolly et al
1995) and substantially less than 0.01 in uids (Weiss et al
1994). Theories that a majority of strain and the ultimate
yielding of a given chain occurs between a single (or a few)
pairs of particles (Tang et al 1995) are typical justication
for this large discrepancy between theory and experimental
evidence.
A small strain approximation of stress given by
equation (5) is
.
J
2
p
2
1

0
h
3
< 0.1 (8)
for which a 3.6% error results when strain = 0.1. The
preyield modulus G of the particle network is simply stress
divided by strain, or, from equation (8)
G .
J
2
p
2
1

0
h
3
< 0.1. (9)
Furthermore, the maximum possible eld induced change
in stress (and modulus) occurs when the aligned particles
become magnetically saturated, i.e. when J
p
= J
s
.
It can be seen that the maximum eld-induced stress
rises quadratically with saturation magnetization J
s
of the
particle material. This observation advocates the use of a
particulate material with a high saturation magnetization.
Pure iron has the highest saturation magnetization of
known elements with J
s
= 2.1 T. However, alloys of
iron and cobalt (e.g. Permendur) have higher saturation
magnetizations with J
s
2.4 T.
2.2. The relationship between particle polarization and
average composite ux density
The average composite ux density is relatively straightfor-
ward to measure experimentally. For example, ux meters
employing eld probes or sense coils can typically measure
the ux density at the composite sample directly. Alterna-
tively, if assumptions are made regarding the reluctance of
the magnetic circuit, the ux density at the composite sam-
ple can be calculated as a function of the applied magneto-
motive force. On the other hand, the particle polarization
is impossible to measure directly and is very difcult to
compute. In this section, theory is developed on the basis
that saturation begins in the particles polar regions at a
very low applied eld and increases towards total particle
saturation as applied eld increases. Average particle polar-
ization is then computed as a function of average composite
ux density.
Consider the schematic in gure 3. We, again, isolate a
single chain of particles from an array of parallel chains. It
is assumed that uniform saturation occurs in each particle
in region (s) shown in gure 3. Furthermore, distance
s decreases from r to zero as average composite ux
density increases from zero to some large level. Letters
in parentheses in gure 3 correspond to subscripts used in
the following development. The magnetic model is one-
dimensional, therefore vector quantities are expressed as
scalar magnitudes.
609
M R Jolly et al
Figure 3. Particle saturation model. Quantities are
consistent with those dened in gure 2. Letters in
parentheses indicate different regions: (s) region where
particle is saturated (shaded), (u) region where particle is
unsaturated, (m) matrix region with assumed relative
permeability
m
= 1. Letters in parentheses also
correspond to subscripts used in theory development.
It is assumed that the ux through the unsaturated
and saturated particle regions is uniformly distributed and
remains constant, i.e.,
p
= B
p
A
p
= B
s
A
s
= B
u
A
u
.
Given that the reluctance of the matrix R
m
and the
reluctance of the particle R
p
are in parallel, the ux
distribution between the two regions can then be expressed
as

p
= B
s
A
s
=
1
1 R
p
/R
m
BA (10)
where A
s
is the particle area at x = s, and A = y
2
is the
total area. It is further assumed that the reluctance of the
particle is the sum of the reluctance of the unsaturated and
saturated regions of the particle:
R
p
= R
u
R
s
=
2

s
0
1
(r
2
x
2
)
dx
2

0
r s
(r
2
s
2
)
.
2

0
r s
(r
2
s
2
)
(11)
and the reluctance of the matrix is
R
m
=
2r

0
(y
2
r
2
)
(12)
where the following additional assumptions have been
made:
p
1 implying that reluctance of the unsaturated
region is negligible; ux lines are perpendicular to the
secants that dene the saturated region; the relative
permeability of the matrix and the saturated portion of the
particle are unity, i.e.,
m
=
s
= 1; and for the sake
of brevity the gap between particles within a chain are
assumed negligible, i.e., h = 1. Inclusion of h > 1 is not
difcult but would detract from the clarity of exposition
sought here.
The ratio of reluctances required by equation (10) can
now be calculated as
R
p
R
m
= k
2/3 1
(1 )
(13)
where = s/r, and particle volume fraction = 2r
2
/3y
2
has been substituted, and k 1 is a factor that has been
added to accommodate for magnetic interaction between
adjacent particle chains and packing arrangements that are
more complex than simple linear chains, i.e., columnar
structures with more than two contact points per particle
that may locally be face center cubic or body center
cubic arrangements. Parameter k may also be used to
accommodate for situations where
m
,= 1. Such a situation
may arise when there are particles within the matrix that
are not within a chain. These particles may contribute
little to the desired eld dependency of the composite yet
may act to increase the effective permeability of the matrix
(analogous to the concept of effective dielectric constant
in electrorheological uids discussed by Adriani and Gast
1988). The following area ratio can be calculated
A
s
A
=
(r
2
s
2
)
y
2
=
3
2

1
2

. (14)
Now, by substituting equations (13) and (14) into (10), the
quantity as a function of B can be solved for in terms of
known parameters.
We now seek to develop a relationship between B and
H, the applied eld. It is known that

0
H = B

J
f

(15)
where J
f
) is the average polarization density of the
composite. Since it has been assumed that
m
= 1, all
polarization is due to the polarization of the particles such
that

0
H = B

J
p

(16)
where J
p
) is the average polarization density of the
particles and may be calculated by

J
p

=
1
2/3r
3

s
0
J
x

r
2
x
2

dx

r
s
J
s

r
2
x
2

dx

(17)
where J
s
is the saturation polarization of the particles and
J
x
is the polarization density at a distance x within the
unsaturated region of the particle and may be written as
J
x
=
(r
2
s
2
)
(r
2
x
2
)
B
s
. (18)
Since, in equation (11), it has been assumed that the
reluctance of the unsaturated portion of the particle is
negligible compared to that of the saturated region (i.e.,
all of the eld appears across the saturated region), it can
be stated that B
s
J
s

0
H, which when combined with
(16) and (17) yields

0
H = B
3/2
r
3

J
s

0
H


s
0

r
2
s
2

dx
J
s

r
s

r
2
x
2

dx

. (19)
Integration of equation (19) with some algebra yields

0
H =
B (1
3
)J
s
1 3/2
3
(20)
and the average polarization density (or intrinsic induction)
of the composite is J
f
) = B
0
H = J
p
), where, from
equation (16), the average particle polarization density is:
J
p
=

J
p

=
3/2
3
B (1
3
)J
s
1 3/2
3
. (21)
610
A magnetorheological material behaviour model
Figure 4. Magnetic induction curves for three MR uids of
different iron volume percents. Experimental (thin lines)
and model (thick lines) results are shown.
3. Magnetorheological uids
3.1. Magnetic properties of MR uids
The magnetic properties of three MR uid samples were
assessed using a magnetic hysteresigraph (KJS Assoc. Inc.,
Model HG-500, per ASTM A773-91). The uid samples
contained three different iron loadings: 10%, 20% and 30%
by volume. The magnetic induction curves (also called the
BH curves) of the three samples are shown in gure 4.
Also shown are theoretical model results computed from
equation (20). As the theory predicts, magnetic induction
increases with eld strength until saturation of the uid
occurs. This can be seen to occur at an intrinsic induction
of J
s
. For example, the MR uid sample loaded with 30%
iron (which in bulk form saturates at 2.1 T) can be seen
to saturate at about (0.3)(2.1 T) = 0.63 T. A very gradual
increase in uid induction is observed after saturation of the
uid. This effectwhich has been observed during testing
of other MR uids (Carlson 1996)is attributed to particle
restructuring over time (Felici et al 1994) and increasing
eld. The permeabilities of these uids were found to be
between approximately 2
0
and 6
0
when measured at
very low elds. These permeabilities are consistent with
results of similar experiments conducted by G okt urk et al
(1993).
Parameter k, which was introduced in equation (13),
was adjusted to t the model to the experimental data.
Recall that parameter k is used to account for magnetic
interactions other than those between adjacent particles
within a chain. The values of k used to t the model were
5.0, 4.6, and 4.4 for the 30%, 20% and 10% iron by volume
MR uid samples, respectively. Since these values are
accurate within about 10%, it can be seen that the model,
when applied to MR uids, is relatively insensitive to the
parameter k. In other words, a value of, say, k = 4.7 would
provide reasonable model accuracy over a wide range of
iron volume fractions.
Figure 5. Experimental yield stress as a function of applied
magnetic eld for two MR uids containing approximately
30% iron by volume. Theoretical model results are also
shown for k = 8 (solid) and k = 5 (broken).
3.2. Magnetorheological effect
Experimental evidence has conrmed that MR uids in
the presence of a magnetic eld exhibit both a pre-yield
regime, characterized by an elastic response, and a post-
yield regime, characterized by a viscous response. The
transition between these two regimes has been reported to
occur at a yield strain on the order of 0.5% (Weiss et al
1994). Since MR uids typically operate in a continuous
yield regime, they are usually characterized by their eld
dependent yield stress.
Figure 5 shows published results of yield stress for
two MR uids formulated with approximately 30% (V/V)
carbonyl iron (Lord 1994, Weiss et al 1993). Also
shown are theoretical model results for two different values
of parameter k (from equations (7b) and (21)). The
model results are seen to be in relatively good agreement
with the experimental data. The gure suggests that the
experimentally applied elds were not sufcient to saturate
the MR uid. This is suggested by the trend demonstrated
by the model and also by the observation that increases
in eld at higher experimental elds are still resulting in
corresponding substantial increases in yield stress.
It is anticipated that experimental hysteresigraph results
can be used to calibrate the theoretical model. As was
done in gure 4, such a calibration entails choosing a
k that allows for a good t between the model and the
experimental induction curves. The values of k, which
are weakly dependent upon iron volume fraction, can then
be used with the model for characterization and design
purposes. Recall that k = 5 was used in conjunction with
the 30% (V/V) iron MR uid sample in gure 4. The
experimental data in gure 5 can be seen to fall to the right
of the theoretical model result for k = 5. It is speculated
that this discrepancy is a result of losses in the magnetic
circuit used to generate the experimental results.
611
M R Jolly et al
Figure 6. Double lap shear test specimen. The elastomer
segments (shaded) are 20 mm by 7.5 mm by 1.0 mm thick
and are sandwiched between an inner iron plate and two
outer iron plates. The vertical arrow indicates the direction
of applied eld.
4. Magnetorheological elastomers
Double lap shear specimens of MR elastomer were prepared
in a mould using silicone oil loaded with a specied volume
percent of carbonyl iron particles of a 34 m mean
diameter (see gure 6). This formulation was then cured
in the presence of a magnetic eld of 0.8 10
6
A m
1
applied in the direction shown in gure 6. This eld
causes parallel particle chains to form while the silicone
is still liquid. Upon crosslinking (during which polymer
molecules interconnect), this particle structure is locked in
place. Three double lap shear specimens of MR elastomer
were prepared containing 10, 20 and 30% iron by volume.
Testing of the double lap shear specimens was
conducted using a Dynastat material testing system. This
system applies a xed oscillatory strain to the specimen
and measures the amplitude and phase of the output
force, from which stress and modulus can be calculated.
Testing involved recording the complex modulus of various
specimens at various frequencies, strains and applied
magnetic elds. Magnetic elds were applied to the test
specimens using a horseshoe-shaped electromagnet with
poles that efciently focused magnetic ux across the
elastomer segments in the direction shown in gure 6. The
induction eld (in Tesla) across the elastomer segments was
experimentally measured using an integrating ux meter
(Walker Scientic, Model MF-3D) and a sense coil wrapped
directly around the elastomer segments.
Figure 7 shows the effect of induced composite ux
density on elastic modulus for the three test specimens. As
can be seen, the maximum change in modulus increases
monotonically with increasing volume percentage of iron.
While the change in modulus increases to nearly 0.6 MPa
as iron loading increases to 30% (V/V), we see from
table 1 that the zero eld modulus also increases with
particle loading. As a result, the percentage of maximum
increase in modulus for the three samples remains relatively
constant between 30% and 40%. The model results provide
reasonable predictions for the maximum change in modulus
obtained at large saturating ux densities for the 20% (V/V)
and 30% (V/V) iron samples. Once again, in using the
present model at intermediate elds, the parameter k, which
accounts for magnetic interaction between neighbouring
particle chains, must be adjusted in order to t the
model to the experimental data. In addition, parameter
h, which accounts for gaps between particles in a chain,
can be adjusted to match the maximum change in modulus
predicted by the model to that observed experimentally.
Figure 7. The response of elastic modulus to applied eld
for 10% (h = 1.29, k = 5), 20% (h = 1.08, k = 12) and 30%
(h = 1.0, k = 15)) iron by volume elastomer specimens.
Experimental data (, , ) and theoretical model data
() are shown.
Because the maximum change in modulus is independent
of parameter k, these two parameters (k and h) can be
adjusted independently to best match the model with the
data. It can be seen that an increase in particle volume
fraction results in a pronounced increase in k, implying that
magnetic interaction between neighboring particle chains
is increasing. This is consistent with intuition because
as particle volume fraction increases, the distance between
neighbouring chains decreases. It is also noted that there is
much greater sensitivity to k when the model is applied to
MR elastomers as compared to MR uids. This is attributed
to variations in particle structure for the two cases, resulting
from the corresponding variations in the level of particle
mobility in the presence of a eld. It was also observed
that as volume fraction of iron increases, the parameter
h decreases suggesting that gaps between particles within
a chain decreases. It is speculated that at lower volume
fractions, particle forces that induce tight, gap-free particle
chains are lower within the precured elastomer as a result
of an increase in the mean distance between particles.
Onset of particle chain yielding in elastomer composites
has been observed in the vicinity of 12% strain (Jolly et
al 1996). As shown in gure 8, this yielding is manifested
by a pronounced drop off in the magnetorheological effect
and an increase in eld dependent energy dissipation
(tan ). The theoretical yield strain has been observed
to be much higher than experimental evidence indicates.
As with MR uids, an explanation for this discrepancy
is that the majority of strain and the ultimate yielding of
a given chain occurs between a single (or a few) pairs
of particles. In other words, at a local level, it is likely
that the theoretical yield strain is achieved, but the strain
averaged over an entire chain is considerably smaller. This
phenomenon is attributed to imperfect chain structures and
non-uniformity in particle geometry and chain structure.
Still, the solids exhibit a yield strain that is several times
higher in magnitude than that of analogous MR uids.
612
A magnetorheological material behaviour model
Table 1. Enumeration of the nominal absolute modulus (no eld applied), the experimental maximum change in modulus for
specimens tested at 1.0% strain and 2.0 Hz, and the maximum change in modulus predicted by the model assuming no gaps
between particles (h = 1).
Nominal Max. change in
modulus, Max. change in modulus,
no eld modulus, exper. model, h = 1
Test specimen (MPa) (MPa) (% change) (MPa)
30% (V/V) iron 1.80 0.56 (31%) 0.55
20% (V/V) iron 0.74 0.29 (39%) 0.36
10% (V/V) iron 0.26 0.08 (30%) 0.17
Figure 8. The effect of peak strain on eld induced change in modulus () and change in tan (). This data corresponds to
the 30%(V/V) iron specimen measured at 10 Hz and 0.85 T.
5. Conclusions
A model that examines the behaviour of MR materials
has been presented. The model is based on the dipolar
interaction of particles within an assumed structure. This
dipolar model was then augmented with a proposed
mechanism of how magnetic ux density distributes itself
within the particle network. This mechanism takes into
account magnetic non-linearities associated with the gradual
saturation of the permeable particles as applied eld
increases. The theory is capable of examining both the
magnetic and mechanical properties of MR materials.
The developed theory was compared with experimental
data taken on MR uids and elastomers. Magnetic and post-
yield properties of MR uids were examined. In particular,
the model was t to experimental magnetic induction curves
by adjusting a parameter that accounts for unmodelled
magnetic interactions between neighboring particle chains.
Indications that these magnetic interactions decreased with
decreasing particle volume fraction were observed. It was
proposed that this t of the model with magnetic induction
curves be a basis for calibrating the model for various
particle loadings. The calibrated model was then observed
to be in relatively good agreement with published yield
stress data for MR uids.
Experimental pre-yield data taken on MR elastomers
was also compared with model results. Once again it was
observed that the model could be calibrated to provide a
good t with experimental data. The maximum change in
modulus was observed to increase monotonically with iron
loadingnearly 0.6 MPa was observed with an iron loading
of 30% (V/V). However, it was also observed that the
zero eld modulus increases with particle loading such that
the percentage of maximum increase in modulus remained
relatively constant between 30% and 40%.
Ongoing work on deterministic modelling of MR
material behaviour will examine the magnetic properties
of more complex particle structures.
References
Adriani P M and Gast A P 1988 A microscopic model of
electrorheology Phys. Fluids 31 275768
Anon 1995 Brake cuts exercise-equipment cost Design News
Dec. 4
Block H and Kelly J P 1988 Electro-rheology J. Phys. D: Appl.
Phys. 21 166177
Bowen C P, Shrout T R, Newnham R E and Randall C A 1995
Tunable electric eld processing of composite materials J.
Intelligent Mater. Syst. Struct. 6 15968
Carlson J D 1991 Private communication
613
M R Jolly et al
Carlson J D 1994 The promise of controllable uids Proc.
Actuator 94 (Bremen, Germany 1994) pp 26670
Carlson J D 1996 Private communication
Carlson J D, Chrzan M J and James F O 1994
Magnetorheological uid devices US Patent No. 5,284,330
Carlson J D, Catanzarite D M and St Clair K A 1995
Commercial magneto-rheological uid devices 5th Int.
Conf. ER, MR Suspensions Assoc. Technol. (Shefeld,
England, 1995)
Chase V D 1996 Cutting edge Appl. Manuf. May p 6
Felici N, Foulc J N and Atten P 1994 A conduction model of
electrorheological effect Electrorheological Fluids:
Mechanisms, Properties, Technology and Applications ed R
Tao and G D Roy (Singapore: World Scientic) pp 13952
Gast A P and Zukoski C F 1989 Electrorheological uids as
colliodal suspensions Adv. Colloid Interface Sci. 30 153202
Gentry S B, Mazur J F and Blackburn B K 1995 Muscle training
and physical rehabilitation machine using
electro-rheological magnetic uid US Patent No. 5,460,585
Ginder J M and Davis L C 1994 Shear stresses in
magnetorheological uids: role of magnetic saturation Appl.
Phys. Lett. 65 34102
Ginder J M, Davis L C and Elie L D 1995 Rheology of
magnetorheological uids: models and measurements 5th
Int. Conf. ER, MR Suspensions Assoc. Technol. (Shefeld,
England, 1995)
G okt urk H S, Fiske T J and Kalyon D M 1993 Electric and
magnetic properties of a thermoplastic elastomer
incorporated with ferromagnetic powders IEEE Trans.
Magn. 29 41706
Harpavat G 1974 Magnetostatic forces on a chain of spherical
beads in a uniform magnetic eld IEEE Trans. Magn. 10
91922
Jolly M R, Carlson J D, Munoz B C and Bullions T A 1996 The
magnetoviscoelastic effect of elastomer composites
consisting of ferrous particles embedded in a polymer
matrix J. Intell. Mater. Syst. Struct. submitted
Kasahara A, Yamada A, Yoshizawa T, Wada K and Yamasaki H
1994 Viscoelastic substance and objective lens driving
apparatus with the same US Patent No. 5,337,865
Lord 1994 Lord Product Information: MR Fluid MRX-100 and
MRX-135CD, Lord Corp., Cary, NC, USA
Mu noz B C and Jolly M R 1995 Mechanical response of
elastomers to magnetic elds 4th Ann. Workshop: Advances
in Smart Mater. for Aerospace Appl. (NASA Langley
Research Center, 1995), 210th ACS National Meeting
(Chicago, IL, 1995)
Rigbi Z and Jilken L 1983 The response of an elastomer lled
with soft ferrite to mechanical and magnetic inuences J.
Magnetism Magn. Mater. 37 26776
Rosensweig R E 1985 Ferrohydrodynamics (Cambridge:
Cambridge University Press)
Spasojevic D, Irvine T F and Afgan N 1974 The effect of a
magnetic eld on the rheodynamic behaviour of
ferromagnetic suspensions Int. J. Multiphase Flow 1 60722
Tang X, Chen Y and Conrad H 1995 Structure and interaction
force in a model magnetorheological system 5th Int. Conf.
ER, MR Suspensions Assoc. Technol. (Shefeld, England,
1995)
Weiss K D et al 1993 High strength magneto- and
electro-rheological uids Proc. SAE SAE paper no 932451
Weiss K D, Carlson J D and Nixon D A 1994 Viscoelastic
properties of magneto- and electro-rheological uids J.
Intell. Mater. Syst. Struct. 5 7725
614

You might also like