Вы находитесь на странице: 1из 10

Soft Matter

Dynamic Article Links <

Cite this: Soft Matter, 2011, 7, 3701

REVIEW

www.rsc.org/softmatter

Magnetorheological fluids: a review


Juan de Vicente,*a Daniel J. Klingenbergb and Roque Hidalgo-Alvareza
Received 28th October 2010, Accepted 13th December 2010
DOI: 10.1039/c0sm01221a
Magnetorheological (MR) materials are a kind of smart materials whose mechanical properties can be
altered in a controlled fashion by an external magnetic field. They traditionally include fluids,
elastomers and foams. In this review paper we revisit the most outstanding advances on the rheological
performance of MR fluids. Special emphasis is paid to the understanding of their yielding, flow and
viscoelastic behaviour under shearing flows.

1. Introduction
Magnetorheological (MR) fluids are intelligent materials which
show a reversible and very fast (in a fraction of millisecond)
transition from a liquid to a nearly solid state under the presence
of external magnetic fields (magnetorheological effect1). These
fluids can exhibit changes in apparent viscosity of several orders
of magnitude for applied magnetic flux densities of order of
magnitude 1 T. This outstanding property makes them very good
candidates for applications in mechanical systems that require

a
Department of Applied Physics, Faculty of Sciences, University of
Granada, C/Fuentenueva s/n, 18071, Granada, Spain. E-mail: jvicente@
ugr.es; Fax: +34 958 243214; Tel: +34 958 240020
b
Department of Chemical and Biological Engineering and Rheology
Research Center, University of Wisconsin, 1415 Engineering Drive,
Madison, Wisconsin, 53706, USA

Juan de Vicente received PhD


degrees in Physics from the
University of Granada (Spain)
and the University of NiceSophia Antipolis (France) in
2002. He has been FPU
Predoctoral Fellow, Marie
Curie Postdoc and Marie Curie
ERG Fellow at the University of
Twente, Rheology Research
Center (University of Wisconsin),
Unilever
Corporate
Research (Colworth, UK) and
Juan de Vicente
Imperial College London. He is
recipient of the Young Investigator Award from the Social
Council and the Physics Research Award from the Academy of
Sciences. He is presently Associate Professor of Applied Physics at
the University of Granada. His research interests include magnetorheology and soft-elasto-ferrohydrodynamic lubrication.
This journal is The Royal Society of Chemistry 2011

the active control of vibrations or the transmission of torque.


Typical examples include shock absorbers, brakes, clutches,
seismic vibration dampers, control valves and artificial joints.2
Other examples involve the use of magnetic fields to control the
thermal energy transfer,3,4 biomedical applications,5 precision
polishing,68 sound propagation,9 isothermal magnetic advection,10 and chemical sensing applications1113 among others.
Conventional MR fluids are two-phase fluids prepared by
dispersing large amounts of solid, micron-sized, highly magnetizable particles (up to 50 vol%) in a non-magnetizable liquid.
Iron particles obtained from the thermal decomposition of iron
pentacarbonylso-called
carbonyl
iron
particlesare
commonly used because of their large saturation magnetization
(m0Ms 2.1 T). Typical carrier liquids are mineral and silicone
oils, polyesters, polyethers, synthetic hydrocarbons and water.
Additives are also necessarily added to inhibit sedimentation and

Daniel J: Klingenberg

Dan Klingenberg is a Professor


of Chemical and Biological
Engineering at the University of
Wisconsin-Madison. He received
a BS in Chemical Engineering
from the University of MissouriRolla in 1985, and a PhD in
Chemical Engineering from the
University of Illinois-Urbana in
1991. He is currently the associate chair of the Rheology
Research Center at the University of Wisconsin. His research
interests include electro- and
magnetorheological fluids, and
flexible fiber suspensions.

Soft Matter, 2011, 7, 37013710 | 3701

aggregation, as well as to provide additional lubricating properties. These include thixotropic agents, surfactants and polymers.
The MR effect is most commonly attributed to the fieldinduced magnetization of the suspended particles. In the absence
of magnetic fields, the suspensions have a relatively low viscosity.
When a magnetic field is applied, the particles magnetize and
attract one another along the field lines and form anisometric
aggregates that span the system. The resulting material typically
exhibits a large yield stressthe minimum shear stress required
to make the suspension flow. This behavior is also typically
characterized as a large shear rate-dependent apparent viscosity
and enhanced viscoelasticity induced by the applied magnetic
field.
Generally speaking, MR fluids must have large saturation
magnetization and small coercivity/remnant magnetization, be
active over a wide temperature range, and be stable against
settling, irreversible flocculation and chemical degradation/
oxidation. The field-dependent mechanical strength of MR fluids
depends on the composition, particle size and volume fraction.
Other types of MR fluids are occasionally employed. Inverse
ferrofluids14 (or magnetic holes) are a class of MR fluids formed
by dispersing micron-sized non-magnetizable particles in a ferrofluid.15 Their mechanical properties can be controlled by
varying the strength of the magnetic field and/or the saturation
magnetization of the ferrofluid. Interest in using inverse ferrofluids comes from the fact that many types of non-magnetizable
particles are available, and thus particle size, shape and functionality are easily tunable.1618 Ferrofluid emulsions are formed
using two immiscible fluids, one of them being magnetic field
responsive.19,20 They have been used in the past as model colloids
to investigate the equilibrium structure of MR fluids.21
Magnetic composites also exist that are the solid counterparts
of ferrofluids and MR fluids, namely magnetic gels22 and elastomers,23 respectively. Magnetic elastomers are made by
dispersing magnetic particles in a polymer solution or melt and
applying a magnetic field prior to crosslinking. Field-responsive
composite materials such as superparamagnetic polymer-based

Roque Hidalgo-Alvarez received


a Master in Science and a PhD
in Physics from the University of
Granada. He joined the Physics
Department at the University of
Granada in September 1975. He
was a postdoctoral fellow at
Wageningen University, the
Netherlands from 1984 to 1985.
His research and teaching
interests lie in the general area
of Soft Matter with a special
emphasis on Colloids. He has
Roque Hidalgo-Alvarez
expertise in the experimental
characterization of colloidal
systems using electrokinetic
techniques, but his research interests also include colloidal aggregation processes, liquid matter topics, and thermodynamics of nonequilibrium.
3702 | Soft Matter, 2011, 7, 37013710

particles have been used in colloidal aggregation and optical


trapping studies24,25 as well as in magneto-electrorheological
applications where a synergistic effect has been reported that is
attributed to highly homogeneous field-induced structures.26,27
By introduction of graphite microparticles into the elastomeric
matrix composites become electroconductive. This property can
be exploited in the fabrication of magnetoresistors, magnetic
field sensors and transducers.28,29
Magnetorheology has been reviewed numerous times.3033 This
review mainly focuses on the most relevant progress during the
last few years and those foreseen for the future regarding
formulations, flow, yielding and viscoelastic properties.

2. MR formulations
As with any micron-scale particulate suspension in which
a density mismatch between the particulate and the surrounding
fluid exists, settling in MR fluids must be considered. Also,
interparticle aggregation is favored due to the large surface areato-volume ratios that tend to reduce the surface energy resulting
in some cases in severe redispersibility problems. These two
problems are generally aggravated by any remnant magnetization of the particles.34 As a result, various solutions have been
proposed in the literature to prevent particle settling and/or
agglomeration.
Reducing the magnetizable particle size should allow for an
enhanced kinetic stability hence reducing sedimentation and
abrasiveness. However, a limitation in size exists when
approaching the nanometre range due to the prevalence of
Brownian forces (see Section 3). Another approach involves lowdensity core/shell-structured and functionalized magnetic particles that may improve the dispersibility, and avoid oxidation and
corrosion by passivating the surface as well as prevent particles
from coming in close contact.35,36
Thixotropic networks can be prepared by using nanostructured fumed silica, anisotropic carbon fibers, acicular iron
oxide nanoparticles, and surfactants such as stearate and
oleate.3739 These space filling gel-like networks prevent the
magnetizable particles from settling. Viscoplastic media have
also been used to combat irreversible sedimentation. Particularly, greases are demonstrated to be promising candidates for
viscoplastic media because of their inherent yield stress.40 Ionic
liquids are also interesting carrier fluids because unlike conventional ones, the properties of ionic liquids can be tuned by
varying the composition of their ions.41
MR suspensions composed of magnetizable fibers show
improved stability against sedimentation, compared to suspensions of spherical magnetizable particles at the same concentration. Fiber suspensions also tend to exhibit a larger MR response
(e.g., larger field-induced yield stress) than suspensions of
spherical particles.42 Such an enhancement has also been
observed in particle-level simulations of fibers and spheres.43 This
behavior for MR fiber suspensions is consistent with that
reported previously for electrorheological suspensions consisting
of fibrous particles.44,45
While small amounts of particulate additives can be used to
improve the stability of MR fluids, mixtures of a variety of
particles at higher concentrations can also be employed to
enhance the rheological properties. Mixtures of magnetizable
This journal is The Royal Society of Chemistry 2011

particles of the same material (e.g., iron) but of two different


diameters (bidisperse suspensions; both in the range of a micron
to a few tens of microns) can be exploited to enhance MR
behavior.46,47 Such bidisperse mixtures show larger field-induced
yield stresses and smaller off-state viscosities than monodisperse
suspensions of either the large spheres or the small spheres. See
et al.48 observed similar behavior for bidisperse suspensions of
electrorheological fluids. Particle-level simulations of bidisperse
suspensions reproduce the enhancement for mixtures.49,50 The
simulations reveal that the enhancement appears to arise because
the smaller diameter spheres tend to break-up large aggregates of
the larger spheres, producing more anisotropic structures.
Suspensions of mixtures of magnetizable spheres with diameters
in the range of a few to tens of microns with much smaller
magnetizable spheres with diameters in the range of a few tens of
nanometres also show larger field-induced stresses than obtained
for monodisperse suspensions at the same total particle volume
fraction. The effect is observed when both the large and small
particles are composed of the same material5154 as well as when
different magnetizable materials are employed.5558 Ngatu et al.59
reported that the addition of magnetizable nanofibers to
a suspension of spherical magnetizable particles did not produce
an enhancement in the field-induced yield stress, although the
mixtures did exhibit improved resistance to sedimentation.
More recently, Ulicny et al.60 have shown that the addition of
nonmagnetizable spheres to conventional MR fluids can produce
an increase in the field-induced yield stress. Particle-level simulations in three dimensions qualitatively reproduce the experimental results. However, the phenomenon is not observed in
two-dimensional simulations of monolayers, which suggests that
the underlying mechanism differs from that causing enhancement in bidisperse suspensions. The ability to control and
enhance the field-induced yield stress in MR fluids by adding
nonmagnetizable particles can be exploited to produce more
active fluids, less dense fluids or less expensive fluids. We note
that this phenomenon appears to be similar to that reported by
Levin et al.,61 where the addition of nonmagnetic abrasive
particles to MR fluids caused an increase in the viscous stress.
In addition to improving the sedimentation stability of MR
fluids and enhancing their rheological properties, another major
challenge in magnetorheology is improving the durability of MR
fluids. Surprisingly, this aspect of MR technology has attracted
relatively little attention despite its importance for the success of
applications.
A lack of durability of MR fluids is typically attributed to the
tendency of iron particles to oxidize. Carlson62 describes in-usethickening, which is the increase in off-state viscosity of MR
fluids over time. This can ultimately result in device failure, as the
MR fluid becomes an unmanageable paste. The cause is believed
to be spalling of the brittle oxide surface layer from the iron
particles. The increase in off-state viscosity may be caused by the
increase in solid volume, and perhaps colloidal forces acting
between the small particles generated.63 Ulicny et al.64 examined
the durability of an MR fan clutch over extended periods. The
field-induced torque capacity decreased slowly with time. Prior
to the durability test, the particles were composed of nearly pure
iron, containing only 0.4 wt% oxygen. Following the durability
test, the particles contained 7 wt% oxygen. Prior to the durability
test, the particles appeared smooth and spherical. Following the
This journal is The Royal Society of Chemistry 2011

durability test, the particles developed a coreshell structure,


with the outer layer containing iron oxide. The authors speculated that the time-dependent decrease in torque capacity may be
caused by the decrease in particle magnetizability with increasing
iron oxide content. Sunkara et al.63 examined the oxidation of
iron particles within MR fluids and its impact on the fieldinduced rheological properties. Experiments were employed in
which the particles were oxidized in a controlled environment.
The magnetic field-induced yield stress decreased with increasing
extent of oxidation. The rheological behavior is consistent with
that predicted using a model in which the particles have a core
shell structure, with the core composed of iron and the shell
composed of iron oxide (magnetite). As the particles oxidize, the
shell grows, resulting in weaker particle magnetizability and
decreased field-induced yield stresses.

3. Physical mechanisms and micromechanical


models
There is only one widely accepted mechanism to account for MR
field-induced magnetization and tunable anisotropic interaction.
This is the so-called particle magnetization model. According to
this model, the MR effect is attributed to the magnetic permeability mismatch between the constituent (continuous and
dispersed) phases. Assuming that the magnetic phase is the solid
one, it is important to remark that, according to the large particle
size, the particles behave as magnetic multidomains. Their
magnetic moments are field-induced and their Brownian motion
is generally negligiblethat is, the thermal forces exerted by the
solvent molecules (fkT) are typically much smaller than
magnetic and hydrodynamic forces.
A considerable simplification is achieved by neglecting
multipole and multibody magnetostatic interactions between
particles. In the linear magnetization regime, an isolated
magnetizable sphere of radius a acquires a magnetic moment
m 4pm0mcrba3H0, where m0 4p  107 Tm A1 is the
permeability of vacuum, mcr is the relative permeability of the
continuous phase, b (mpr  mcr)/(mpr + 2mcr) is the contrast
factor or coupling parameter, mpr is the relative permeability of
the particles, and H0 is the magnetic field strength. In contrast, at
high fields, the particle magnetisation saturates and the magnetic
4
moment is independent of the field strength, m pm0 mcr a3 Ms .
3
The magnitude of the magnetic interaction energy between two
magnetic moments in the linear regime relative to thermal energy
is the so-called l parameter:15
l

pm0 mcr b2 a3 H02


2kB T

(1)

where 0 < b < 1 for conventional (strong) MR fluids and 0.5 <
b < 0 for inverse ferrofluids (weak MR fluids). For sufficiently
large values of l, magnetostatic particle interactions dominate
over thermal motion, resulting in chain-like particle aggregates.
For sufficiently small values of l, Brownian motion dominates
and field-induced aggregates are absent. The structure evolution
of dilute MR fluids (approximately 0.15 vol%) has been
extensively investigated under DC uniaxial fields using videomicroscopy observations,65 optical tweezers25 and (2D) light
scattering.66 Interestingly, the equilibrium structure in MR fluids
Soft Matter, 2011, 7, 37013710 | 3703

is determined by only f and l as demonstrated by Furst and


Gast.67 This is illustrated in Fig. 1 where the dominant mechanisms of lateral chain aggregation are mapped as a function of
f and l.
From an experimental point of view, it has been recognized
that the rate of field increase and container size profoundly
impact the final structuration.68 Labyrinthine structures appear
far away from equilibrium if the rate of field increase is large.
Moreover, Promislow and Gast21 demonstrated that a DC field
increased at a very slow rate does not necessarily produce the
lowest energy suspension structure. They used instead pulsed
magnetic fields to investigate the structuration in the absence of
flow.
The kinetic aggregation process is typically divided in two well
differentiated regimes. Initially, the average length of the
aggregates increases as a power law according to Smoluchowski
equation. Later, single-width chain-like structures laterally
aggregate to form columnar structures. Mechanisms explaining
this lateral aggregation are the coalescence through torquedriven zippering motion and thermal fluctuations of particle
positions.69 Since thermal Landau Peierls fluctuations are not
relevant in the case of conventional (e.g. iron-based) MR fluids,
the coarsening is thought to be driven by topological defects that
first appear during the initial tip-to-tip stage of the aggregation
process. These defects lead to a local variation in the magnetic
field surrounding the chain-like structures. Using magnetic
energy minimization principles, Bossis and coworkers developed
a model to explain this field-induced phase separation.65 Furst
and Gast67 measured the lateral interactions between dipolar
chains directly using optical tweezers.
Another dimensionless number (so-called Mason number,
Mn) is relevant when the MR fluid is subjected to flow. The
Stokesian approximation is generally employed to capture
hydrodynamic forces since Reynolds numbers are typically
small. For steady shear flow, Mn is basically a dimensionless
shear rate that can be defined as the ratio of hydrodynamic drag
(estimated as Stokes drag) and magnetostatic forces (estimated
as the dipole force magnitude) acting on the particles:70

Fig. 1 Dominant mechanisms of lateral interactions as a function of the


dipolar interaction parameter l and the initial volume fraction f.
Adapted with permission from E. M. Furst and A. P. Gast, Phys. Rev. E,
62, 6916, 2000. Copyright (2000) by the American Physical Society.67 HT
stands for the Halsey and Toor model in which aggregation is driven by
fluctuations of relaxed chains.137,138 MHT stands for the modified Halsey
and Toor model in which fluctuation-driven aggregation occurs before
chains can relax.139,140

3704 | Soft Matter, 2011, 7, 37013710

Mn

8hc g_
m0 mcr b2 H 2

(2)

where hc is the continuous phase viscosity and g_ is the magnitude


of the shear rate tensor. It is worth noting that different definitions exist for the Mason number in the case of large magnetic
fields.71 As expected, Mn and l parameters are related through
_ 3/kBT via Mnl 2Pe/3. It has been
Peclet number Pe 6phcga
recently demonstrated that other forces apart from hydrodynamic and magnetic (e.g. van der Waals, acidbase, frictional
and body forces) may play a significant role in the redispersability and off-state properties of MR fluids.7274 However, in
the case of negligible short range forces, any rheological property
of MR fluids will depend on only f, l and Mn.
In the absence of a magnetic field, MR fluids typically behave
as Newtonian fluids. However, when a magnetic field is applied
transverse to the direction of flow, a yielding, shear thinning and
viscoelastic behavior is generally observed, as a consequence of
the magnetic field-induced structuration in the colloid. Let us
now briefly review these three features in the case of simple
shearing flows.

3.1.

Yielding

One of the most relevant rheological properties of a MR fluid is


the yield stress that must be overcome to initiate gross material
deformation or flow.75 The yield stress is the minimum stress
value required for the onset of flow and is of special interest in
torque transfer applications. It is proportional to the force
required to break field-induced structures.
Most relevant yield stress models published in the MR literature can be classified in two groups. Macroscopic models
employ magnetic energy minimization principles and assume
homogeneous structures consisting of spheroidal, cylindrical or
layered particle aggregates. Basically, these models are based on
a mesoscopic description of the structure only taking into
account the shape anisotropy of the strained aggregates under
small deformation.76 Microscopic models take into account
interparticle interactions, which include for example single
particle width chains sheared under an external field.77,78 In most
yield stress models, the shear stress is considered to be dominated
by the magnetostatic interactions between the particles, and
shear-induced deformation is assumed to be affine. Furthermore,
interactions between field-induced structures are ignored, hence
limiting the analysis to low particle concentrations. Other yield
stress models have been elaborated in the literature involving
fractal aggregate concepts79 and percolation theory.80 According
to the percolation theory, an anisotropic network microstructure
rather than fibrillated chains should be formed when the volume
fraction is larger than the percolation threshold. The presence of
a magnetic field strengthens the network without significantly
affecting its shape.
Both macroscopic and microscopic models were demonstrated
in the past to apply well in the prediction of the yield stresses in
inverse ferrofluids.76,81 However, for conventional MR fluids,
macroscopic models are not able to reproduce yielding behaviour
observed experimentally. Meanwhile, microscopic models that
include the anisotropic polar magnetostatic interactions between
particles can provide reasonable agreement with experiments.78
This journal is The Royal Society of Chemistry 2011

At small field strengths, the yield stress is predicted to be


essentially proportional to the magnetic field strength squared.81
As the applied field strength increases and the particle or the
ferrofluid magnetization begins to saturate (in the case of
conventional MR fluids or inverse ferrofluids, respectively), the
yield stress will increase
sub-quadratically with the external field
p
1=2 3=2
strength, sy 6fm0 Ms H0 , eventually becoming field
strength-independent at large field strengths, sy 0.086fm0M2s.78
All micromechanical models based on the fibrillation model
(i.e., gap-spanning single particle-width chains) predict a yield
stress that increases linearly with the volume fraction for all
magnetic field strengths. However, experiments reported in the
literature reveal a linear behaviour only for diluted MR fluids at
low fields.82 In general, a more rapid than linear increase with
volume fraction is observed for more concentrated conventional
MR fluids that is thought to be associated with thick columnar
structures and/or non-affine motion of the aggregates. In the case
of inverse ferrofluids it has been documented that the yield stress
presents a maximum with increasing volume fraction.76
There are many ways for experimentally measuring the yield
stress, either using direct or indirect approaches, involving steady
shear and oscillatory tests.83 One can define three yield stresses as
obtained through shearing flow experiments (Fig. 2):84 the
elastic-limit yield stress, the static yield stress and the dynamic
yield stress. The elastic-limit yield stress represents the maximum
shear stress that can be applied while still obtaining complete
recovery when the stress is removed. The static (or frictional)
yield stress is the minimum stress required to cause the fluid to
flow. This yield stress is frequently associated to the slip of the
aggregates on the plates rather than the structure collapse/break
under shear. The static yield stress is typically estimated by using
creep tests, the so-called tangent method, and low-shear extrapolation of stress-controlled data in double logarithmic representations of stress versus shear rate. This latter procedure
corresponds to the threshold stress associated with the sudden
decrease in viscosity upon increasing the stress plotted on a log
log scale. Finally, the dynamic yield stress corresponds to the
stress needed to continuously break the aggregates which reform
in the presence of the magnetostatic forces once the stress exceeds
the static yield stress. It can be calculated by fitting a viscoplastic

constitutive model, such as the Bingham, HerschelBulkley or


Casson equations, to experimental data at nonzero shear rates.
The dynamic yield stress is typically larger than the static yield
stress. Even though this is undoubtedly the most widely used
yield stress estimator, the determination of the dynamic yield
stress is controversial and always involves the use of indirect
methods.85

3.2.

Flow behavior

Magnetized MR fluids behave as strongly shear thinning materials with a viscosity that decreases with the shear rate.
For l [ 1, experimental viscosity data at various shear rates and
field strengths typically collapse to a function of only the Mason
number suggesting that this number satisfactorily captures the
scaling of both hydrodynamic and magnetostatic forces. Many
examples exist in the literature for a wide range of MR
fluids70,71,81,82,86 as well as ER fluids.8790
In the limit of l / N, Marshall et al.91 employed dimensional
analysis to show that when the Bingham model suffices to explain
the steady shear flow behavior and the yield stress scales as
fm0mcrb2H2, the dimensionless viscosity can be written
h/hN 1 + Mn*(f)MnD

(3)

with D 1. Here, hN is the (field strength-independent) highshear viscosity and Mn*(f) is the critical Mason number that
determines the transition from magnetization to hydrodynamic
control of suspension structure. Slight refinements to this model
exist in the literature, in particular, by calculating the high-shear
viscosity using the KriegerDougherty equation.92 Micromechanical chain-like models were later developed to understand
the particle volume fraction dependence of Mn*(f) by balancing
magnetostatic and hydrodynamic forces and torques on fieldinduced structures.70,86,93,94 These models give Mn* Cfhc/hN
where different values for C are derived depending on specific
assumptions and/or simplifications in the mechanical stability
conditions in the problem: C 8.82,93 C 8.485,94 C 5.25,86
and C 1.91.70 In all cases, the C coefficients presented here have
been calculated assuming the bare point-dipole approximation.

Fig. 2 (a) Typical yield stresses under stress growth (start-up) shearing flow tests. (b) An example of the shear stress growth after start-up as a function
of the shear strain at a shear rate of g_ 0.0103 s1 for inverse ferrofluids. The magnetic field is 76.7 kA m1 and the volume fraction of the nonmagnetic
particles is f 0.18. Particle radii: 53 nm (+), 84 nm (O), 138 nm (C), 189 nm (>). Reprinted with permission from B. J. de Gans, N. J. Duin, D. van
den Ende and J. Mellema, J. Chem. Phys., Vol. 113, Page 2032, (2000). Copyright (2000), American Institute of Physics.16

This journal is The Royal Society of Chemistry 2011

Soft Matter, 2011, 7, 37013710 | 3705

such low-shear plateaus are not observed, perhaps because


sufficiently small shear rates (which can require very long times)
are simply not examined. In this sense, other more convenient
tools for inferring low-shear viscosities involve small-amplitude
unsteady shear tests.
3.3.

Fig. 3 Dimensionless viscosity (h/hN) curves as a function of Mason


number (Mn) for an inverse ferrofluid containing 19.3 vol% 378 nm
radius silica particles in a 44 mPas ferrofluid81. Lines correspond to
theoretical models: solid line, Martin and Anderson;93 dashed line, de
Vicente et al.;94 dotted line, de Gans et al.;86 dash-dotted line, Volkova
et al.70 Also shown is the best fit to h/hN 1 + Mn*MnD with Mn*
0.202 and exponent D 0.80, and the Mn value associated to the
breakage of the last doublet (Mnb).141

In general, eqn (3) satisfactorily fits experimental data with


1 < D < 2/3 (e.g. Felt et al.82 D 0.740.83, de Gans et al.86
D 0.80.9 and Volkova et al.70 D 0.740.87). The exponents obtained experimentally are greater than D < 2/3 as
predicted by Halsey et al.,89 obtained by minimization of the total
magnetic energy of spheroidal aggregates; the experimentally
determined exponents also satisfy D > 1, consistent with eqn
(3). Fig. 3 shows a typical example of a viscosity curve corresponding to a model inverse ferrofluid.
Interestingly, a low-shear plateau is occasionally observed
experimentally under steady shear flow (see Fig. 3), contrary to
the existence of a yield stress in MR fluids.86 This low-shear
plateau has been theoretically predicted using single-width chain
models93 and Stokesian dynamics simulations on monolayer ER
suspensions for l < 10.95 However, in most studies on MR fluids,

Fig. 4 Pipkin diagram of the dynamic rheological behavior of MR


fluids. The curves demarcate regions of different rheological behavior.
Reprinted from J. Non-Newtonian Fluid Mechanics, 81, M. Parthasarathy and D. J. Klingenberg, Large amplitude oscillatory shear of ER
suspensions, 83, Copyright (1999), with permission from Elsevier.96

3706 | Soft Matter, 2011, 7, 37013710

Viscoelasticity

Oscillatory flow provides valuable tools for investigating the


structure of MR fluids. In contrast to steady, unidirectional
flows, oscillatory flows can provide access to a wide range of
temporal scales for a particular structure, and in general there is
less ambiguity than in yield stress determinations from simple
shearing plots. Furthermore, from a practical point of view, MR
devices often operate in oscillatory mode.
Parthasarathy and Klingenberg96 mapped the frequency and
strain dependence of the rheological properties of field-responsive fluids in the form of a Pipkin diagram (Fig. 4). At small
strains the system exhibits linear viscoelasticity. The transition
from the linear (LVE) to the non-linear viscoelastic regime
(NLVE) is characterized by the appearance of high order
harmonics, thus defining a critical strain (gc,1). For larger strain
amplitudes, a transition from the nonlinear viscoelasticity to
viscoplasticity (VP) is associated with a second critical strain
amplitude (gc,2). At very high frequencies and/or strain amplitudes, the elastic properties of the fluid are negligible and the
system behaves as a Newtonian fluid (N).
The linear viscoelastic behavior of MR fluids has been extensively investigated using small amplitude oscillatory shear
(SAOS). In the presence of a magnetic field, a large storage
modulus, G0 , appears, which is associated with the field-induced
structures, and is typically at least an order of magnitude larger
than the loss modulus, G00 . Under the assumption that stresses
are solely determined by the magnetic field (hence neglecting
dissipation), micromechanical models discussed in Section 3.1
have been used in the past to predict G0 values by simply taking
the low strain limit for the ratio s/g. For intermediate magnetic
field strengths, Ginder et al.78 predicted that G0 3fm0MsH. The
nonquadratic dependence on the field strength arises as the
particle magnetization begins to saturate near the poles in chainlike aggregates. At large field strengths where the particles
magnetization is completely saturated, the storage modulus
becomes independent of field strength, given by G0 0.3fm0M2s.
This behavior is consistent with a variety of experimental reports.
Other contributions coming from multipolar and multibody
interactions are manifested in some cases.97 As expected, in the
case of weak MR fluids (e.g. inverse ferrofluids) macroscopic
micromechanical models also satisfactorily predict experimental
observations.18
Amplitude sweep tests reveal a small linear viscoelastic region
(strain amplitudes in the range of approximately 0.01 to 0.1%)
that shifts to even lower strain amplitudes upon increasing the
magnetic field strength. Experimental data are generally found to
collapse when plotting the dimensionless scaled viscoelastic
moduli G0 /H and G00 /H vs. the scaled stress amplitude s/H3/2,98 in
agreement with Ginder et al.78 According to their viscoelastic
behavior, MR fluids are classified as Type III Complex
Fluids.99,100 The dependence of the viscoelastic moduli on the
shear strain/stress amplitude varies with the particle volume
This journal is The Royal Society of Chemistry 2011

fraction. Suspensions below the percolation transition exhibit


a particle reorganization under the field that is associated with
a viscoelastic enhancement. On the other hand, for highly
concentrated MR fluids, flocculation already exists even in the
absence of the field. Hence, upon shearing, the network structure
is disrupted and viscoelasticity decreases.101
The frequency dependence of linear viscoelastic material
functions is not clear. Various experiments and simulations
suggest that moduli can remain constant, pass through
a maximum, or decrease or increase depending on the system
under study. A frequency independent loss tangent has been
reported to correspond to a gelation transition point.102 For
magnetic field strengths below gel transition, a 2 : 1 power law
dependence of G0 and G00 on the oscillation frequency in the
terminal zone indicates that the relaxation times associated with
the self-similar structure reside within this frequency range.
Understanding the non-linear regime (large amplitude oscillatory shear, or LAOS) is also important because the linear
viscoelastic region is very small (i.e. limited to very small strain
amplitudes), and hence of limited value for practical applications. An experimental investigation is described by Li and
coworkers.103 A microscopic visco-elasto-plastic model inspired
by classical soft-glassy rheology models was successfully used by
McKinley and coworkers to predict the LAOS (and shear creep)
behavior of MR fluids using a scaling transformation involving
a dimensionless stress s/sy and a dimensionless strain gG0 /sy,
where sy is the yield stress.104
Oscillatory and steady state shear flows are found to be related
through a modified CoxMerz rule. Curves of h vs. g_ closely overlap
with jhj vs. gu plots.105 This is expected since LAOS and steady
shear behavior share a common structural mechanism: aggregation
and fragmentation of clusters.106 However, in the literature there are
examples that do not obey CoxMerz rule as well.107

4. Magnetorheometry
Most devices that use MR fluids can be classified as having either
fixed poles (pressure-driven flow modePDF) or relatively
moveable poles (direct-shear modeDS and squeeze-film
modeS). Diagrams of these three basic operational standard
flow modes are shown in Fig. 5. Examples of PDF mode devices

include servo-valves, dampers and shock absorbers. Examples of


DS mode devices include clutches, brakes, chucking and locking
devices. The third mode of operation also known as biaxial
elongational flow mode appears in slow motion and/or high force
applications.108
The classical DS mode has been studied thoroughly in the
literature109 and, as a result, several products are already present
in the market. However, relatively limited research attention has
been given to the behavior of MR-fluids in pressure-driven flows,
in spite of the facts that this flow mode is the most common in
current applications (e.g., automotive dampers) and much larger
and realistic shear rates can be imposed (up to 20 000 s1, in
contrast to conventional rotational rheometry that encompasses
shear rates only up to a few thousand s1).110 Also, very scarce
information exists in the literature regarding the S mode, even
though it has been suggested that the yield stress that could be
achieved would be up to ten times larger than that attainable
with either the DS or PDF modes.111,112
Simple shear flow or approximations thereof have been used as
a standard tool to investigate the MR effect. Ramp-up stresscontrolled or strain-controlled experiments are typically performed under torsional shear flow between parallel disks, coneplate or concentric-cylinder geometries to obtain the associated
apparent viscosity material function and hence the flow curves
and resulting yield stresses by extrapolation at low-shear rates
(see Section 3.1). Unsteady (time-dependent) shear flows have
also been investigated in the past including stress growth (startup), stress decay/relaxation (cessation of flow), retardation
(creep), relaxation (step strain) and SAOS tests. The rheological
transient response in which a pulse-type magnetic field is applied
to MR fluids at a fixed shear rate and shear stress is reported in
Ulicny et al.113 and See and Gordin114 respectively. MR fluids do
creep below the yield point implying that the sample shows
irreversible plastic deformation in response to the applied stress.
A part of this time-dependent deformation is recovered when the
applied stress is removed which provides a measure of the elastic
properties of the material.115 Apart from widely used SAOS
techniques, step strain tests were also carried out by Li et al.116 to
investigate the viscoelastic properties of MR fluids covering both
the preyield and postyield regimes. The stress relaxation modulus
was found to obey the time-strain factorability and the damping

Fig. 5 Basic operational modes for controllable MR fluid devices: (a) pressure driven flow mode, (b) direct shear mode, and (c) biaxial elongational flow
mode. Reprinted from Materials and Design, 28, A. G. Olabi and A. Grunwald, Design and application of magneto-rheological fluid, 2658, Copyright
(2007), with permission from Elsevier.108,109

This journal is The Royal Society of Chemistry 2011

Soft Matter, 2011, 7, 37013710 | 3707

function was well fitted by a Soskey and Winter model. The


transient response of a MR fluid after a double-step shear strain
has been investigated by See.117 Combined flows have also been
investigated118 and field-induced normal stresses have also been
reported in the absence and presence of shearing and compression flows.119121 The relaxation time spectrum of MR fluids was
obtained by Keentok and See.122

5. Concluding remarks and future challenges


MR fluids constitute a stimulating family of magnetic-field
responsive materials with applications in three well differentiated
areas: (i) colloids with adjustable rheological properties, (ii)
patterned anisotropic self-assembled materials, and (iii) sensors
for monitoring mechanical vibrationss. They present a rich
rheological behavior showing yield stress, shear thinning and
viscoelastic response that can be externally tuned through
magnetic fields.
Typical particle sizes employed in MR fluid formulations are
in the range of micrometres. This results in a strong magnetization of the particles since their field-induced magnetic moment
depends on the particle volume. On the other hand, the typical
particle size for ferrofluids is around tens of nanometres.
Contrarily to MR fluids, these nanoparticles exhibit a permanent
magnetic moment as a consequence of their magnetic monodomain character where the Neel relaxation mechanism dominates. The investigation of monodispersed systems with typical
sizes in between these two extremes has not been done yet in
detail mainly due to the absence of appropriate chemical
approaches to prepare kinetically stable and monodisperse
enough magnetic colloids. Just recently, some publications have
appeared in the literature that allow us to prepare these particles
(using chemical precipitation techniques123,124 and polyol
processes125,126) and this allows their use in MR formulations and
the study of the interesting field existing in between ferrofluids
and classical MR fluids.
Traditionally, particles used in MR fluid formulations are
spherical. Recently, it has been demonstrated that the incorporation of nonspherical particles to MR formulations has significant benefits. Important efforts have been done in recent years
to obtain well defined monodisperse non-spherical particles
(spheroidal and plate-like shapes) that only differ in shape in
isolation from other parameters such as the size and saturation
magnetization.42,127,128 Further exploration of non-spherical
particulate mixtures may provide even more advantages for
optimizing the various properties of MR fluids, in particular
their durability and abrasiveness. Currently, some work exists on
the lubrication capabilities of ferrofluids,129 however, abrasiveness and lubrication performance of MR fluids are still widely
unknown.
MR fluids have been traditionally investigated under shearing
flow conditions. Most MR fluids present a yield stress in the
typical observational time scale that can be explained in terms
of plastic constitutive equation under macroscopic and microscopic approaches. The yield stress is proportional to the
magnetic field strength squared at small field strengths and
increases sub-quadratically when the particles poles begin to
saturate. At very large field strengths, the yield stress becomes
field strength-independent. Under flow conditions a typical
3708 | Soft Matter, 2011, 7, 37013710

shear thinning behavior is observed as a consequence of the


balance between shear/magnetic driven fragmentation/aggregation processes. A good scaling with Mason number is generally
found reinforcing the idea that the particle magnetization model
drives the structure evolution. Generally speaking, chain-like
micromechanical models do overestimate the shear thinning
exponent and experimental observations. On the other hand the
linear viscoelastic behavior reveals a field dependent storage
modulus. Even though many results exist for shear flow
behavior, some open questions concern the determination of
yield stress, the shear thinning exponent less than one, the field
dependent behavior of the loss modulus and the non-linear
viscoelastic behavior.
Among the three basic modes of operationvalve, shear and
squeeze modesthere is no doubt that shear mode has been the
most studied. However, with the advent of new devices it has
been recently demonstrated that the MR damping forces under
squeeze flow mode are superior by an order of magnitude than
the other two.130,131 Surprisingly, detailed investigations of MR
performance under squeeze flow are still lacking. Moreover, the
search for new operation modes is very active today and as
a consequence of these efforts an alternative valve configuration
called Magnetic Gradient Pinch has been recently identified for
controlling MR fluids.132
In practice, neither of the three modes appear in isolation. In
fact, conventional applications involve the presence of complex
flows. As a first approach, to get a full understanding of the
complete picture it is necessary to deepen our understanding of
the MR performance under combined flows created by superposition of standard flows such as valve, shear and squeeze flows
(cf. Fig. 5).118
Customarily, MR fluids are magnetized through the application of uniaxial homogeneous DC magnetic fields.121 Biaxial and
triaxial AC fields have been recently used to form new fieldstructured composites for directed; efficient heat transfer133 and
non-homogeneous magnetic fields have been used in High
Gradient Magnetic Separation techniques.134 Surprisingly, the
rheological behavior of MR fluids under these operating conditions has not been investigated in detail yet.135
Finally, MR fluids have been used in the past as model systems
to a better understanding of flocculation phenomena.95 Model
MR fluids have been recently used to demonstrate that shear
thickening in dense suspensions can be masked by a yield stress
arising from the tunable strength of interparticle attractions,136
and in general, these materials could be used as a benchmark for
the validation of glass transition theories and soft glassy
rheology mainly because the interparticle interactions that drive
the glassy dynamics can be easily tuned externally through the
application of magnetic fields.

Acknowledgements
This work was supported by MICINN MAT 2009-14234-C03-03
and MAT 2010-15101 projects (Spain), by the European
Regional Development Fund (ERDF) and by Junta de
Andaluca P07-FQM-02496, P07-FQM-03099 and P07-FQM02517 projects (Spain). This work was also supported in part by
the National Science Foundation (award no. 0932680).
This journal is The Royal Society of Chemistry 2011

References
1 J. Rabinow, AIEE Trans., 1948, 67, 13081315.
2 D. J. Klingenberg, AIChE J., 2001, 47, 246249.
3 M. Heine, J. de Vicente and D. J. Klingenberg, Phys. Fluids, 2006,
18, 023301.
4 B. N. Reinecke, J. W. Shan, K. K. Suabedissen and
A. S. Cherkasova, J. Appl. Phys., 2008, 104, 023507.
5 J. Liu, G. A. Flores and R. Sheng, J. Magn. Magn. Mater., 2001,
225, 209217.
6 W. Kordonski and D. Golini, J. Intell. Mater. Syst. Struct., 2002, 13,
401404.
7 S. Jha and V. K. Jain, Int. J. Mach. Tool Manufact., 2004, 44, 1019
1029.
8 W. I. Kordonski, A. B. Shorey and M. Tricard, J. Fluids Eng., 2006,
128, 2026.
9 F. Donado, J. L. Carrillo and M. E. Mendoza, J. Phys.: Condens.
Matter, 2002, 14, 21532157.
10 K. J. Solis and J. E. Martin, Appl. Phys. Lett., 2010, 97, 034101.
11 D. H. Read and J. E. Martin, Anal. Chem., 2010, 82(5), 21502154.
12 D. H. Read and J. E. Martin, Adv. Funct. Mater., 2010, 20(10), 1577
1584.
13 D. H. Read and J. E. Martin, Anal. Chem., 2010, 82(16), 69696975.
14 A. T. Skjeltorp, Phys. Rev. Lett., 1983, 51, 23062309.
15 R. E. Rosensweig, Ferrohydrodynamics, Dover, New York, 1997.
16 B. J. de Gans, N. J. Duin, D. van den Ende and J. Mellema, J. Chem.
Phys., 2000, 113, 20322042.
17 R. Saldivar-Guerrero, R. Richter, I. Rehberg, N. Aksel, L. Heymann
and O. S. Rodriguez-Fernandez, J. Chem. Phys., 2006, 125, 084907.

18 J. Ramos, J. de Vicente and R. Hidalgo-Alvarez,
Langmuir, 2010,
26(12), 93349341.
19 A. P. Gast and C. F. Zukoski, Adv. Colloid Interface Sci., 1989, 30,
153202.
20 J. Bibette, J. Magn. Magn. Mater., 1993, 122, 3741.
21 J. H. E. Promislow and A. P. Gast, Langmuir, 1996, 12, 40954102.
22 M. Zrinyi, Colloid Polym. Sci., 2000, 278, 98103.
23 W. H. Li, Y. Zhou and T. F. Tian, Rheol. Acta, 2010, 49, 733740.
24 J. H. E. Promislow, A. P. Gast and M. Fermigier, J. Chem. Phys.,
1995, 102, 54925498.
25 E. M. Furst and A. P. Gast, Phys. Rev. Lett., 1999, 82, 41304133.
26 K. Minagawa, T. Watanabe, K. Koyama and M. Sasaki, Langmuir,
1994, 10, 39263928.
27 W. J. Wen, N. Wang, W. Y. Tam and P. Sheng, Appl. Phys. Lett.,
1997, 27, 25292531.
28 I. Bica and H. J. Choi, Int. J. Mod. Phys. B, 2008, 22, 50415064.
29 I. Bica, J. Ind. Eng. Chem., 2010, 16, 359363.
30 P. J. Rankin, J. M. Ginder and D. J. Klingenberg, Curr. Opin.
Colloid Interface Sci., 1998, 3, 373381.
31 G. Bossis, S. Lacis, A. Meunier and O. Volkova, J. Magn. Magn.
Mater., 2002, 252, 224228.
32 I. Bica, J. Ind. Eng. Chem., 2006, 12(4), 501515.
33 L. Vekas, Adv. Sci. Technol., 2008, 54, 127136.
34 P. P. Phule, M. P. Mihalcin and S. Genc, J. Mater. Res., 1999, 14,
30373041.
35 J. S. Choi, B. J. Park, M. S. Cho and H. J. Choi, J. Magn. Magn.
Mater., 2006, 304, 374376.
36 W. P. Wu, B. Y. Zhao, Q. Wu, L. S. Chen and K. A. Hu, Smart
Mater. Struct., 2006, 15, N94N98.
37 P. P. Phule, US Pat., 5985168, 1999.
38 B. D. Chin, J. H. Park, M. H. Kwon and O. O. Park, Rheol. Acta,
2001, 40, 211219.
39 J. de Vicente, M. T. L
opez-L
opez, F. Gonzalez-Caballero and
J. D. G. Dur
an, J. Rheol., 2002, 47, 10931109.
40 P. J. Rankin, A. T. Horvath and D. J. Klingenberg, Rheol. Acta,
1999, 38, 471477.
41 C. Guerrero-Sanchez, T. Lara-Ceniceros, E. Jimenez-Regalado,
M. Rasa and U. S. Schubert, Adv. Mater., 2007, 19, 17401747.
42 R. C. Bell, J. O. Karli, A. N. Vavreck, D. T. Zimmerman, G. T. Ngatu
and N. M. Wereley, Smart Mater. Struct., 2008, 17, 015028.
43 Y. K. Kor and H. See, Rheol. Acta, 2010, 49, 741756.
44 R. C. Kanu and M. T. Shaw, J. Rheol., 1998, 42(3), 657670.
45 Y. Otsubo, Colloids Surf., A, 1999, 153, 459466.
46 R. T. Foister, US Pat., 5667715, 1997.
47 A. M. Trendler and H. Bose, Int. J. Mod. Phys. B, 2005, 19, 1416
1422.

This journal is The Royal Society of Chemistry 2011

48 H. See, A. Kawai and F. Ikazaki, Rheol. Acta, 2002, 41, 5560.


49 D. Kittipoomwong, D. J. Klingenberg and J. C. Ulicny, J. Rheol.,
2005, 49(6), 15211538.
50 C. Ekwebelam and H. See, Rheol. Acta, 2009, 48, 1932.
51 N. M. Wereley, A. Chaudhuri, J.-H. Yoo, S. John, S. Kotha,
A. Suggs, R. Radhakrishnan, B. J. Love and T. S. Sudarshan,
J. Intell. Mater. Syst. Struct., 2006, 17, 393401.
52 J. L. Viota, J. D. G. Duran and A. V. Delgado, Eur. Phys. J. E: Soft
Matter Biol. Phys., 2009, 29, 8794.
53 K. H. Song, B. J. Park and H. J. Choi, IEEE Trans. Magn., 2009,
45(10), 40454048.
54 F. Jonsdottir, K. H. Gudmundsson, T. B. Dijkman,
F. Thorsteinsson and O. Gutfleisch, J. Intell. Mater. Syst. Struct.,
2010, 21(11), 10511060.
55 J. M. Ginder, L. D. Elie and L. C. Davies, US Pat., 5549837, 1996.
56 M. T. L
opez-L
opez, J. de Vicente, G. Bossis, F. Gonzalez-Caballero
and J. D. G. Duran, J. Mater. Res., 2005, 20(4), 874881.
57 X. Wang and F. Gordaninejad, Rheol. Acta, 2006, 45, 899908.
58 D. Susan-Resiga, D. Bica and L. Vekas, J. Magn. Magn. Mater.,
2010, 322, 31663172.
59 G. T. Ngatu, N. M. Wereley, J. O. Karli and R. C. Bell, Smart
Mater. Struct., 2008, 17, 045022.
60 J. C. Ulicny, K. S. Snavely, M. A. Golden and D. J. Klingenberg,
Appl. Phys. Lett., 2010, 96, 231903.
61 M. L. Levin, D. E. Polesskii and I. V. Prokhorov, J. Eng. Phys.
Thermophys., 1997, 70, 769772.
62 J. D. Carlson, Proc. 8th Int. Conf. on Electrorheological Fluids and
Magnetorheological Suspensions, World Scientific, Singapore, 2001,
pp. 6369.
63 S. R. Sunkara, T. W. Root, J. C. Ulicny and D. J. Klingenberg,
J. Phys.: Conf. Ser., 2009, 149, 012081.
64 J. C. Ulicny, M. P. Balogh, N. M. Potter and R. A. Waldo, Mater.
Sci. Eng., A, 2007, 443, 1624.
65 S. Cutillas, G. Bossis and A. Cebers, Phys. Rev. E: Stat. Phys.,
Plasmas, Fluids, Relat. Interdiscip. Top., 1998, 57, 804811.
66 J. E. Martin and J. Odinek, J. Rheol., 1995, 39(5), 9951009.
67 E. M. Furst and A. P. Gast, Phys. Rev. E: Stat. Phys., Plasmas,
Fluids, Relat. Interdiscip. Top., 2000, 62, 69166925.
68 M. Mohebi, N. Jamasbi and J. Liu, Phys. Rev. E: Stat. Phys.,
Plasmas, Fluids, Relat. Interdiscip. Top., 1996, 54(5), 54075413.
69 G. Bossis, O. Volkova, S. Lacis and A. Meunier, Magnetorheology:
Fluids, Structures, and Rheology, in Ferrofluids, ed. S. Odenbach,
Springer, Bremen, Germany, 2002, p. 202.
70 O. Volkova, G. Bossis, M. Guyot, V. Bashtovoi and A. Reks,
J. Rheol., 2000, 44, 91104.
71 D. J. Klingenberg, J. C. Ulicny and M. A. Golden, J. Rheol., 2007,
51, 883893.
72 J. de Vicente, J. D. G. Duran, A. V. Delgado, F. Gonzalez-Caballero
and G. Bossis, Int. J. Mod. Phys. B, 2002, 16, 25762582.
73 D. J. Klingenberg, C. K. Olk, M. A. Golden and J. C. Ulicny,
J. Phys.: Conf. Ser., 2009, 149, 012063.
74 A. J. F. Bombard, L. S. Antunes and D. Gouv^ea, J. Phys.: Conf.
Ser., 2009, 149, 012038.
75 G. Bossis, P. Kuzhir, S. Lacis and O. Volkova, J. Magn. Magn.
Mater., 2003, 258, 456458.
76 G. Bossis, E. Lemaire, O. Volkova and H. Clercx, J. Rheol., 1997, 41,
687704.
77 D. J. Klingenberg and C. F. Zukoski, IV, Langmuir, 1990, 6, 1524.
78 J. M. Ginder, L. C. Davis and L. D. Elie, Int. J. Mod. Phys. B, 1996,
10, 32933303.
79 C. Gehin, J. Persello, D. Charraut and B. Cabane, J. Colloid
Interface Sci., 2004, 273, 658667.
80 T. Hao, Y. Chen, Z. Xu, Y. Xu and Y. Huang, Chin. J. Polym. Sci.,
1994, 12, 97105.

81 J. Ramos, D. J. Klingenberg, R. Hidalgo-Alvarez
and J. de Vicente,
J. Rheol., 2011, 55, 127152.
82 D. W. Felt, M. Hagenbuchle, J. Liu and J. Richard, J. Intell. Mater.
Syst. Struct., 1996, 7, 589593.
83 T. G. Mezger, The Rheology Handbook, Vincentz, Coatings
compendia, 2nd edn, 2006.
84 R. T. Bonnecaze and J. F. Brady, J. Rheol., 1992, 36(1), 73115.
85 P. C. F. Moller, A. Fall and D. Bonn, Europhys. Lett., 2009, 87(3),
38004.
86 B. J. de Gans, H. Hoekstra and J. Mellema, Faraday Discuss., 1999,
112, 209224.

Soft Matter, 2011, 7, 37013710 | 3709

87 C. J. Gow and C. F. Zukoski, J. Coll. Int. Sci., 1990, 136, 175188.


88 D. J. Klingenberg, D. Dierking and C. F. Zukoski, J. Chem. Soc.,
Faraday Trans., 1991, 87(3), 425430.
89 T. C. Halsey, J. E. Martin and D. Adolf, Phys. Rev. Lett., 1992,
68(10), 15191522.
90 D. J. Klingenberg, MRS Bull., 1998, August, 3034.
91 L. Marshall, C. F. Zukoski and J. W. Goodwin, J. Chem. Soc.,
Faraday Trans. 1, 1989, 85, 27852795.
92 J. W. Goodwin, G. M. Markham and B. Vincent, J. Phys. Chem. B,
1997, 101, 19611967.
93 J. E. Martin and R. A. Anderson, J. Chem. Phys., 1996, 104(12),
48144827.
94 J. de Vicente, M. T. L
opez-L
opez, J. D. G. Duran and F. GonzalezCaballero, Rheol. Acta, 2004, 44, 94103.
95 Y. Baxter-Drayton and J. F. Brady, J. Rheol., 1996, 40(6), 1027
1056.
96 M. Parthasarathy and D. J. Klingenberg, J. Non-Newtonian Fluid
Mech., 1999, 81, 83104.
97 J. de Vicente, M. T. L
opez-L
opez, J. D. G. Duran and G. Bossis, J.
Colloid Interface Sci., 2005, 282, 193201.
98 S. S. Deshmukh and G. H. McKinley, Proc. XIVth Int. Congr. on
Rheology, 2004.
99 K. Hyun, S. Kim, K. H. Ahn and S. J. Lee, J. Non-Newtonian Fluid
Mech., 2002, 107, 5165.
100 H. G. Sim, K. H. Ahn and S. J. Lee, J. Non-Newtonian Fluid Mech.,
2003, 112, 237250.
101 T. Hao and Y. Xu, J. Colloid Interface Sci., 1997, 185, 324331.
102 B. D. Chin and H. H. Winter, Rheol. Acta, 2002, 41, 265275.
103 W. H. Li, H. Du, G. Chen, S. H. Yeo and N. Guo, Rheol. Acta, 2003,
42, 280286.
104 S. S. Deshmukh, Development, Characterization and Applications
of Magnetorheological Fluid based smart Materials on the
Macro-to-Micro Scale, PhD thesis, MIT, 2007.
105 D. Doraiswamy, A. N. Mujumdar, I. Tsao, A. N. Beris,
S. C. Danforth and A. B. Metzner, J. Rheol., 1991, 35(4), 647685.
106 W. H. Li, H. Du and N. Q. Guo, Mater. Sci. Eng., A, 2004, 371, 9
15.
107 C. Balan, D. Broboana, E. Gheorghiu and L. Vekas, J. NonNewtonian Fluid Mech., 2008, 154, 2230.
108 M. R. Jolly and J. D. Carlson, Actuator 96, 5th Int. Conf. on New
Actuators, ed. H. Borgmann and K. Lenz, Axon Technologies
Consult GmbH, 1996.
109 A. G. Olabi and A. Grunwald, Mater. Des., 2007, 28, 26582664.
110 X. Wang and F. Gordaninejad, Rheol. Acta, 2006, 45, 899908.
111 X. Z. Zhang, X. L. Gong, P. Q. Zhang and Q. M. Wang, J. Appl.
Phys., 2004, 96, 23592364.
112 E. C. McIntyre and F. E. Filisko, J. Intell. Mater. Syst. Struct., 2007,
18, 12171220.
113 J. C. Ulicny, M. A. Golden, C. S. Namuduri and D. J. Klingenberg,
J. Rheol., 2005, 49, 87104.
114 H. See and A. Gordin, J. Rheol., 2008, 36, 5964.

3710 | Soft Matter, 2011, 7, 37013710

115 W. H. Li, H. Du, G. Chen and S. H. Yeo, Mater. Sci. Eng., A, 2002,
333(12), 368376.
116 W. H. Li, H. Du, G. Chen, S. H. Yeo and N. Q. Guo, Smart Mater.
Struct., 2002, 11, 209217.
117 H. See, Colloid Polym. Sci., 2003, 281, 788793.
118 P. Kulkarni, C. Ciocanel, S. L. Vieira and N. Naganathan, J. Intell.
Mater. Syst. Struct., 2003, 14, 99104.
119 J. de Vicente, F. Gonzalez-Caballero, G. Bossis and O. Volkova,
J. Rheol., 2002, 46, 12951303.
120 H. See and R. Tanner, Rheol. Acta, 2003, 42, 166170.
121 H. M. Laun, C. Gabriel and G. Schmidt, J. Non-Newtonian Fluid
Mech., 2008, 148, 4756.
122 M. Keentok and H. See, Korea Aust. Rheol. J., 2007, 19(3), 117123.
123 M. Andres-Verges, R. Costo, A. G. Roca, J. F. Marco, G. F. Goya,
C. J. Serna and M. P. Morales, J. Phys. D: Appl. Phys., 2008, 41,
134003.
124 F. Vereda, J. de Vicente, M. P. Morales, F. Rull and R. Hidalgo
Alvarez,
J. Phys. Chem. C, 2008, 112, 58435849.
125 F. Fievet, Fine Particles, ed. T. Sugimoto, Marcel and Dekker, New
York, 2000, pp. 460496.
126 F. Bozon-Verduraz, F. Fievet, J.-Y. Piquemal, R. Brayner, K. El
Kabouss, Y. Soumare, G. Viau and G. Shafeev, Braz. J. Phys.,
2009, 39, 134140.
127 J. de Vicente, J. P. Segovia-Gutierrez, E. Andablo-Reyes, F. Vereda

and R. Hidalgo-Alvarez,
J. Chem. Phys., 2009, 131, 194902.
128 J. de Vicente, F. Vereda, J. P. Segovia-Gutierrez, M. P. Morales and

R. Hidalgo-Alvarez,
J. Rheol., 2010, 54, 13371343.

129 E. Andablo-Reyes, J. de Vicente, R. Hidalgo-Alvarez,
C. Myant,
T. Reddyhoff and H. A. Spikes, Tribol. Lett., 2010, 39, 109114.
130 Y. Tian, Y. Meng, H. Mao and S. Wen, Phys. Rev. E: Stat.,
Nonlinear, Soft Matter Phys., 2002, 65, 031507.
131 J. de Vicente, J. A. Ruiz-L
opez, E. Andablo-Reyes, J. P. Segovia
Gutierrez and R. Hidalgo-Alvarez,
submitted.
132 F. D. Goncalves and J. D. Carlson, J. Phys.: Conf. Ser., 2009, 149,
012050.
133 J. E. Martin and G. Gulley, J. Appl. Phys., 2009, 106, 084301.
134 J. Faraudo and J. Camacho, Colloid Polym. Sci., 2010, 288, 207215.

135 E. Andablo-Reyes, R. Hidalgo-Alvarez
and J. de Vicente, Soft
Matter, 2011, 7, 880883.
136 E. Brown, N. A. Forman, C. S. Orellana, H. Zhang, B. W. Maynor,
D. E. Betts, J. M. DeSimone and H. M. Jaeger, Nat. Mater., 2010,
9(3), 220224.
137 T. C. Halsey and W. Toor, J. Stat. Phys., 1990, 61, 12571281.
138 T. C. Halsey and W. Toor, Phys. Rev. Lett., 1990, 65, 28202823.
139 J. E. Martin, J. Odinek and T. C. Halsey, Phys. Rev. Lett., 1992, 69,
15241527.
140 J. E. Martin, J. Odinek, T. C. Halsey and R. Kamien, Phys. Rev. E:
Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top., 1998, 57, 756
775.
141 O. Volkova, S. Cutillas and G. Bossis, Phys. Rev. Lett., 1999, 82,
233236.

This journal is The Royal Society of Chemistry 2011

Вам также может понравиться