+ All Categories
Home > Documents > Magnetic fluid rheology and flows · 2019. 11. 20. · Magnetic fluid rheology and flows Carlos...

Magnetic fluid rheology and flows · 2019. 11. 20. · Magnetic fluid rheology and flows Carlos...

Date post: 30-Jan-2021
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
17
Magnetic fluid rheology and flows Carlos Rinaldi a,1 , Arlex Chaves a,1 , Shihab Elborai b,2 , Xiaowei (Tony) He b,2 , Markus Zahn b, * a University of Puerto Rico, Department of Chemical Engineering, P.O. Box 9046, Mayaguez, PR 00681-9046, Puerto Rico b Massachusetts Institute of Technology, Department of Electrical Engineering and Computer Science and Laboratory for Electromagnetic and Electronic Systems, Cambridge, MA 02139, United States Available online 12 October 2005 Abstract Major recent advances: Magnetic fluid rheology and flow advances in the past year include: (1) generalization of the magnetization relaxation equation by Shliomis and Felderhof and generalization of the governing ferrohydrodynamic equations by Rosensweig and Felderhof; (2) advances in such biomedical applications as drug delivery, hyperthermia, and magnetic resonance imaging; (3) use of the antisymmetric part of the viscous stress tensor due to spin velocity to lower the effective magnetoviscosity to zero and negative values; (4) and ultrasound velocity profile measurements of spin-up flow showing counter-rotating surface and co-rotating volume flows in a uniform rotating magnetic field. Recent advances in magnetic fluid rheology and flows are reviewed including extensions of the governing magnetization relaxation and ferrohydrodynamic equations with a viscous stress tensor that has an antisymmetric part due to spin velocity; derivation of the magnetic susceptibility tensor in a ferrofluid with spin velocity and its relationship to magnetically controlled heating; magnetic force and torque analysis, measurements, resulting flow phenomena, with device and biomedical applications; effective magnetoviscosity analysis and measurements including zero and negative values, not just reduced viscosity; ultrasound velocity profile measurements of spin-up flow showing counter-rotating surface and co-rotating volume flows in a uniform rotating magnetic field; theory and optical measurements of ferrofluid meniscus shape for tangential and perpendicular magnetic fields; new theory and measurements of ferrohydrodynamic flows and instabilities and of thermodiffusion (Soret effect) phenomena. D 2005 Elsevier Ltd. All rights reserved. Keywords: Magnetic fluids; Ferrofluids; Ferrohydrodynamics; Magnetization; Magnetic susceptibility; Magnetic forces; Magnetic torques; Magnetoviscosity; Magnetic thermodiffusion; Magnetic Soret effect 1. Introduction to magnetic fluids 1.1. Ferrofluid composition Magnetic fluids, also called ferrofluids, are synthesized colloidal mixtures of non-magnetic carrier liquid, typically water or oil, containing single domain permanently magne- tized particles, typically magnetite, with diameters of order 5–15 nm and volume fraction up to about 10% [1 && ,2 && ,3 & ,4 & ,5 & ]. Brownian motion keeps the nanoscopic particles from settling under gravity, and a surfactant layer, such as oleic acid, or a polymer coating surrounds each particle to provide short range steric hindrance and electrostatic repulsion between particles preventing particle agglomeration [6 & ]. This coating allows ferrofluids to maintain fluidity even in intense high-gradient magnetic fields [7] unlike magneto- rheological fluids that solidify in strong magnetic fields [2 && ,8]. Recent experiments and analysis show that magnetic dipole forces in strong magnetic fields cause large nano- particles to form chains and aggregates that can greatly affect macroscopic properties of ferrofluids even for low nanoparticle concentration [9–12]. Small angle neutron 1359-0294/$ - see front matter D 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.cocis.2005.07.004 * Corresponding author. Tel.: +1 617 253 4688; fax: +1 617 258 6774. E-mail addresses: [email protected] (C. Rinaldi), [email protected] (A. Chaves), [email protected] (S. Elborai), [email protected] (X. He), [email protected] (M. Zahn). 1 Tel.: +1 787 832 4040/3585; fax: +1 787 834 3655/+1 787 265 3818. 2 Tel.: +1 617 253 5019; fax: +1 617 258 6774. Current Opinion in Colloid & Interface Science 10 (2005) 141 – 157 www.elsevier.com/locate/cocis
Transcript
  • www.elsevier.com/locate/cocis

    Current Opinion in Colloid & Interface

    Magnetic fluid rheology and flows

    Carlos Rinaldi a,1, Arlex Chaves a,1, Shihab Elborai b,2, Xiaowei (Tony) He b,2, Markus Zahn b,*

    a University of Puerto Rico, Department of Chemical Engineering, P.O. Box 9046, Mayaguez, PR 00681-9046, Puerto Ricob Massachusetts Institute of Technology, Department of Electrical Engineering and Computer Science and Laboratory for

    Electromagnetic and Electronic Systems, Cambridge, MA 02139, United States

    Available online 12 October 2005

    Abstract

    Major recent advances: Magnetic fluid rheology and flow advances in the past year include: (1) generalization of the magnetization

    relaxation equation by Shliomis and Felderhof and generalization of the governing ferrohydrodynamic equations by Rosensweig and

    Felderhof; (2) advances in such biomedical applications as drug delivery, hyperthermia, and magnetic resonance imaging; (3) use of the

    antisymmetric part of the viscous stress tensor due to spin velocity to lower the effective magnetoviscosity to zero and negative values; (4)

    and ultrasound velocity profile measurements of spin-up flow showing counter-rotating surface and co-rotating volume flows in a uniform

    rotating magnetic field.

    Recent advances in magnetic fluid rheology and flows are reviewed including extensions of the governing magnetization relaxation and

    ferrohydrodynamic equations with a viscous stress tensor that has an antisymmetric part due to spin velocity; derivation of the magnetic

    susceptibility tensor in a ferrofluid with spin velocity and its relationship to magnetically controlled heating; magnetic force and torque

    analysis, measurements, resulting flow phenomena, with device and biomedical applications; effective magnetoviscosity analysis and

    measurements including zero and negative values, not just reduced viscosity; ultrasound velocity profile measurements of spin-up flow

    showing counter-rotating surface and co-rotating volume flows in a uniform rotating magnetic field; theory and optical measurements of

    ferrofluid meniscus shape for tangential and perpendicular magnetic fields; new theory and measurements of ferrohydrodynamic flows and

    instabilities and of thermodiffusion (Soret effect) phenomena.

    D 2005 Elsevier Ltd. All rights reserved.

    Keywords: Magnetic fluids; Ferrofluids; Ferrohydrodynamics; Magnetization; Magnetic susceptibility; Magnetic forces; Magnetic torques; Magnetoviscosity;

    Magnetic thermodiffusion; Magnetic Soret effect

    1. Introduction to magnetic fluids

    1.1. Ferrofluid composition

    Magnetic fluids, also called ferrofluids, are synthesized

    colloidal mixtures of non-magnetic carrier liquid, typically

    water or oil, containing single domain permanently magne-

    tized particles, typically magnetite, with diameters of order

    1359-0294/$ - see front matter D 2005 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.cocis.2005.07.004

    * Corresponding author. Tel.: +1 617 253 4688; fax: +1 617 258 6774.

    E-mail addresses: [email protected] (C. Rinaldi),

    [email protected] (A. Chaves), [email protected] (S. Elborai),

    [email protected] (X. He), [email protected] (M. Zahn).1 Tel.: +1 787 832 4040/3585; fax: +1 787 834 3655/+1 787 265 3818.2 Tel.: +1 617 253 5019; fax: +1 617 258 6774.

    5 –15 nm and volume fraction up to about 10%

    [1&&,2&&,3&,4&,5&].

    Brownian motion keeps the nanoscopic particles from

    settling under gravity, and a surfactant layer, such as oleic

    acid, or a polymer coating surrounds each particle to provide

    short range steric hindrance and electrostatic repulsion

    between particles preventing particle agglomeration [6&].

    This coating allows ferrofluids to maintain fluidity even in

    intense high-gradient magnetic fields [7] unlike magneto-

    rheological fluids that solidify in strong magnetic fields

    [2&&,8]. Recent experiments and analysis show that magnetic

    dipole forces in strong magnetic fields cause large nano-

    particles to form chains and aggregates that can greatly

    affect macroscopic properties of ferrofluids even for low

    nanoparticle concentration [9–12]. Small angle neutron

    Science 10 (2005) 141 – 157

  • C. Rinaldi et al. / Current Opinion in Colloid & Interface Science 10 (2005) 141–157142

    scattering (SANS) distinguishes between magnetic and non-

    magnetic components of ferrofluids allowing density,

    composition, and magnetization profiles to be precisely

    determined [13,14]. Nanoparticle dynamics, composition,

    and magnetic relaxation affect magnetic fluid rheology

    which can be examined using magnetic field induced

    birefringence [15–19].

    The study and applications of ferrofluids, invented in

    the mid-1960s, involve the multidisciplinary sciences of

    chemistry, fluid mechanics, and magnetism. Because of the

    small particle size, ferrofluids involved nanoscience and

    nanotechnology from their inception. With modern advan-

    ces in understanding nanoscale systems, current research

    focuses on synthesis, characterization, and functionaliza-

    tion of nanoparticles with magnetic and surface properties

    tailored for application as micro/nanoelectromechanical

    sensors, actuators, and micro/nanofluidic devices [20&];

    and for such biomedical applications as [20&,21&&] nano-

    biosensors, targeted drug-delivery vectors [22], magneto-

    cytolysis of cancerous tumors, hyperthermia [23],

    separations and cell sorting, magnetic resonance imaging

    [24,25], immunoassays [26], radiolabelled magnetic fluids

    [27], and X-ray microtomography for three dimensional

    analysis of magnetic nanoparticle distribution in biological

    applications, a crucial parameter for therapeutic evaluation

    [28].

    1.2. Applications

    Conventional ferrofluid applications use DC magnetic

    fields from permanent magnets for use as liquid O-rings in

    rotary and exclusion seals, film bearings, as inertial

    dampers in stepper motors and shock absorbers, in

    magnetorheological fluid composites, as heat transfer

    fluids in loudspeakers, in inclinometers and accelerome-

    ters, for grinding and polishing, in magnetocaloric pumps

    and heat pipes [1&&,2&&,3&,4&,5&], and as lubrication in

    improved hydrodynamic journal bearings [29]. Ferrofluid

    is used for cooling over 50 million loudspeakers each year.

    Almost every computer disk drive uses a magnetic fluid

    rotary seal for contaminant exclusion and the semiconduc-

    tor industry uses silicon crystal growing furnaces that

    employ ferrofluid rotary shaft seals. Ferrofluids are also

    used for separation of magnetic from non-magnetic

    materials and for separating materials by their density by

    using a non-uniform magnetic field to create a magnetic

    pressure distribution in the ferrofluid that causes the fluid

    to act as if it has a variable density that changes with

    height. Magnetic materials move to the regions of

    strongest magnetic field while non-magnetic materials

    move to the regions of low magnetic field with matching

    effective density. Magnetomotive separations use this

    selective buoyancy for mineral processing, water treatment

    [30], and sink-float separation of materials, one novel

    application being the separation of diamonds from beach

    sand.

    2. Governing ferrohydrodynamic equations

    2.1. Magnetization

    2.1.1. Langevin magnetization equilibrium

    Ferrofluid equilibrium magnetization is accurately

    described by the Langevin equation for paramagnetism

    [1&&,2&&,3&,4&,5&,31&&]

    M0 ¼ MS cotha �1

    a

    � �; a ¼ l0mH=kT ð1Þ

    where in equilibrium M̄0 and H̄ are collinear; Ms=Nm =

    Md/ is the saturation magnetization when all magneticdipoles with magnetic nanoparticle volume Vp and magnet-

    ization Md have moment m =MdVp aligned with H̄; N is

    the number of magnetic dipoles per unit volume; and / isthe volume fraction of magnetic nanoparticles in the

    ferrofluid. For the typically used magnetite nanoparticle

    (Md=4.46�105 A/m or l0Md=0.56 T), a representativevolume fraction of / =4% with nanoparticle diameterd =10 nm (Vp =5.25�10�25 m3) gives a ferrofluidsaturation magnetization of l0Ms=l0Md/ =0.0244 T andN =/ /Vp�7.6�1022 magnetic nanoparticles/m3.

    At low magnetic fields the magnetization is approx-

    imately linear with H̄

    M̄M 0

    H̄H¼ v0 ¼ l0m2N= 3kTð Þ ¼

    p18kT

    � �/l0M

    2d d

    3 ð2Þ

    where v0 is the magnetic susceptibility, related to magneticpermeability as l =l0(1+v0). For our representative num-bers at room temperature we obtain v0�0.42 and l /l0�1.42. When the initial magnetic permeability is large,the interaction of magnetic moments is appreciable so that

    Eq. (2) is no longer accurate. Shliomis [31&&,1&&] considers

    the case of monodispersed particles and uses a method

    similar to that used in the Debye–Onsager theory of polar

    fluids to replace Eq. (2) with

    v0 2v0 þ 3ð Þv0 þ 1

    ¼ l0m2N= kTð Þ ¼k

    6kT

    � �/l0M

    2d d

    3: ð3Þ

    Langevin magnetization measurements in the linear low

    field region (ab1) provide an estimate of the largestmagnetic particle diameters and in the high field saturation

    regime (aH1) gives an estimate of the smallest magneticdiameters [1&&,32&]

    limaH1

    M0

    MS,1� 1

    a¼ 1� 6kT

    kl0MdHd3

    limab1

    M0

    MS,

    a3¼ kl0MdHd

    3

    18kTð4Þ

    This method allows estimation of the ferrofluid mag-

    netic nanoparticle size range using a magnetometer. Other

    methods include transmission electron microscopy, atomic

  • C. Rinaldi et al. / Current Opinion in Colloid & Interface Science 10 (2005) 141–157 143

    force microscopy, and forced Rayleigh scattering of light

    [33].

    2.1.2. Magnetization relaxation

    When the ferrofluid is at rest and the magnetization is

    not in equilibrium due to time transients or time varying

    magnetic fields, the conventional magnetization equation

    for an incompressible fluid derived by Shliomis is

    [31&&,34&&]

    BM̄M

    Btþ v̄vIlð ÞM̄M � x̄x � M̄M þ

    M̄M � M̄M 0� �

    s¼ 0 ð5Þ

    where s is a relaxation time constant, v̄ is the ferrofluidlinear flow velocity, and x̄ is the ferrofluid spin velocitywhich is an average particle rotation angular velocity.

    Soon after Eq. (5) was derived, another magnetization

    equation was derived microscopically using the Fokker–

    Planck equation in the mean field approximation [35&&].

    Defining half the flow vorticity as X̄X ¼ 12l� v̄v, the two

    magnetization equations are in good agreement when

    Xsb1, but (5) leads to non-physical results for XsH1.Another analysis by Felderhof used irreversible thermody-

    namics of ferrofluids with hydrodynamics and the full set

    of Maxwell equations to derive a closely related but not

    identical equation to Eq. (5) [36&,37&]. Shliomis [38&,39&]

    states that Felderhof’s relaxation equation leads to an

    incorrect limiting value of rotational viscosity when the

    Langevin parameter a in Eq. (1) becomes large whileFelderhof [40&] replies that Shliomis’ approximate theory

    for dilute ferrofluids does not apply in the dense regime

    and has limited validity in the dilute regime. Recent

    rotational measurements with an applied DC magnetic

    field perpendicular to capillary Poiseuille flow agree with

    the Shliomis result and differ from Felderhof’s result

    [41&]. Earlier similar measurements with an alternating

    magnetic field had a qualitative fit to the Shliomis

    rotational viscosity result but required a frequency

    dependent fitting parameter to successfully fit theory and

    experiment [42].

    More recently, Shliomis has derived a third magnet-

    ization equation from irreversible thermodynamics which

    describes magnetoviscosity in the entire range of magnetic

    field strength and flow vorticity [43&]. It is stated that this

    new magnetization equation is an improvement over that

    of Ref. [35&&] by being valid even far from equilibrium. Yet

    another analysis by Müller and Liu uses standard non-

    equilibrium thermodynamics with magneto-dissipative

    effects to derive a magnetization relaxation equation using

    their ferrofluid dynamics theory [44]. They state that

    Shliomis’ Debye theory of Eq. (5) and his effective-field

    theory approach are special cases of the new set of

    equations.

    2.1.3. Magnetization relaxation time constants

    When a DC magnetic field H̄ is applied to a ferrofluid,

    just like a compass needle, each magnetic nanoparticle

    with magnetic moment m̄ experiences a torque l0m̄� H̄which tends to align m̄ and H̄. There are two important

    time constants that determine how long it takes m̄ to align

    with H̄,

    sB ¼ 3gVh=kT ; sN ¼ s0e KVp=kTð Þ: ð6Þ

    The Brownian rotational relaxation time, sB, describesthe hydrodynamic process when the magnetic moment is

    fixed to the nanoparticle and surfactant layer of total

    hydrodynamic volume Vh, including surfactant or other

    surface coatings, and the whole nanoparticle rotates in a

    fluid of viscosity g to align m̄ and H̄. The Néel timeconstant, sN, is the characteristic time for the magneticmoment inside the particle to align with H̄, due to flipping

    of atomic spins without particle rotation. The parameter K is

    the particle magnetic anisotropy and Vp is the volume of

    particle magnetic material. The literature gives different

    values for the anisotropy constant K of magnetite, over the

    range of 23,000 to 100,000 J m�3 while s0 approximatelyequals 10�9 s. Recent work has used Mossbauer spectro-

    scopy to show that the value of K is size dependent,

    increasing as particle size decreases and gives a value of

    K =78,000 J m�3 for 12.6 nm diameter magnetic nano-

    particles [45&]. Other work has measured the effective

    anisotropy constant in diluted ferrofluids using electron spin

    resonance spectrometry and AC complex magnetic suscept-

    ibility measurements [46]. The total magnetic time constant

    s including both Brownian and Néel relaxation is given byRefs. [1&&,2&&,3&,4&,5&,31&&]

    1

    s¼ 1

    sBþ 1

    sN`s ¼ sBsN

    sB þ sN: ð7Þ

    The time constant s is dominated by the smaller of sBor sN. For K =78,000 J m

    �3 with 12.6 nm diameter

    magnetic nanoparticles at T=300 K, sN¨368 ms. For aferrofluid with representative suspending medium viscosity

    of g =0.001 Pa-s and with a surfactant thickness d =2 nm,sB�1.7 ls. For these parameters we see that the effectivetime constant is dominated by sB. Torque measurementswith ferrofluid particles of order 10 nm diameter have

    estimated, by comparison to theory, effective magnetic

    relaxation times of the order of the computed Brownian

    relaxation time [47&,48&]. Brownian and Néel relaxation are

    lossy processes leading to energy dissipation so that the

    complex magnetic susceptibility has an imaginary part.

    These losses lead to heat generation using time varying

    magnetic fields which can be used for localized treatment

    of cancerous tumors [21&&,23].

    In rotating magnetic fields, the time constant s of Eq.(7) results in the magnetization direction lagging H̄,

    therefore a time average torque acts on each nanoparticle

    causing the particles and surrounding fluid to spin. This

    leads to new physics as the fluid behaves as if it is filled

    with nanosized gyroscopes that stir and mix the fluid

  • C. Rinaldi et al. / Current Opinion in Colloid & Interface Science 10 (2005) 141–157144

    balanced by the non-symmetric part of the viscous stress

    tensor.

    Rotational Brownian motion of ferrofluid nanoparticles

    in a time dependent magnetic field can have thermal noise

    induced rotation due to rectification of thermal fluctuations

    [49,50]. A plastic sphere filled with ferrofluid was

    suspended on a fiber and placed in a horizontal magnetic

    field of the form H =HDC+H1[cos xt +asin(2xt+b)].When the magnetic field was turned on, the sphere rotated.

    The explanation given was that rotation was due to thermal

    ratchet behavior [50]. However, Shliomis describes the true

    cause of rotation to be a magnetic torque due to the

    nonlinearity of ferrofluid magnetization as given by Eq. (1)

    [51].

    2.2. Complex magnetic susceptibility

    We consider the low field linear region of the Langevin

    curve so that in equilibrium M̄0=v0H̄, with v0 given by Eqs.(2) or (3). We examine the sinusoidal steady state magnet-

    ization response for a uniform magnetic field at radian

    frequency Xf in the ferrofluid layer with planar Couette flowshown in Fig. 1, with upper surface driven at constant

    velocity V in the z direction. The ferrofluid flow velocity is

    then of the form m̄ =mz(x)īz. The magnetization is independ-ent of y and z because the applied magnetic field is assumed

    uniform. Then the second term of Eq. (5) is zero. For

    magnetization and magnetic field driven at field frequency

    Xf, the complex amplitude sinusoidal steady state forms ofH̄ and M̄ are

    H̄H ¼ Re ĤH xīx þ ĤH zīz�e jXf tg��

    M̄M ¼ Re M̂M xīx þ M̂M zīz�e jXf tg:��

    ð8Þ

    Phase shifts between M̄ and H̄ components cause a

    non-zero torque, T̄=l0M̄� H̄, which will result in a timeaverage spin velocity of the form x̄ =xy(x)ī y. Thecomplex magnetic susceptibility tensor is then found by

    x

    Ferrofluid

    d

    y

    v ωyz

    Fig. 1. A planar ferrofluid layer between rigid walls, in planar Couette flow driven

    by uniform x and z directed DC magnetic fields.

    solving M̄ as a function of H̄ in Eq. (5) in the linear limit

    of Eqs. (2) or (3)

    M̂M xM̂M z

    � �¼ v0

    jXf s þ 1� �2 þ xys� �2h i

    jXf s þ 1� �

    xys� xys jXf s þ 1

    � �� � ĤH xĤH z

    � �:

    ð9Þ

    If xy=0, the tensor relationship reduces to a complexscalar magnetic susceptibility

    v ¼ v V� jvW ¼ M̄MH̄H

    ¼ v01þ jXf s� �`v V

    ¼ v01þ Xf s

    � �2h i ; vW ¼ v0Xf s1þ Xf s

    � �2h i ð10Þwhere vV is the real part of v related to magnetic field energystorage and vW is the imaginary part of v which causespower dissipation. Note that vW has a peak value of v0 /2when Xf s =1. Spectroscopy measurements of the real andimaginary parts of v as a function of frequency provideinformation about the magnetization dynamics [52–54&,55–

    57]. Other work uses a kinetic approach to microrheology

    that adds stress retardation (mechanical memory) to a

    viscoelastic fluid or weak gel containing a ferrosuspension

    to produce new and interesting non-Newtonian properties

    and dynamic susceptibility effects [58].

    2.3. Heating

    Increasing concentrations of ferrofluid magnetic material

    have a greater rate of heating and higher temperature

    increase when placed in an alternating magnetic field [59].

    The time average dissipated power per unit volume for

    fields at sinusoidal radian frequency Xf is [60–62]

    < Pd > ¼ Re1

    2l0jXf

    ˆ̄MM̄MM I ˆ̄HH̄HHT� �

    ð11Þ

    where the superscript asterisk (*) denotes a complex

    conjugate.

    Hz

    z

    V

    Hx or Bx

    by the x =d surface moving at z directed velocity V, is magnetically stressed

  • C. Rinaldi et al. / Current Opinion in Colloid & Interface Science 10 (2005) 141–157 145

    For an oscillating magnetic field with Ĥx = Ĥz =H0 and

    magnetization given in Eq. (9) for a ferrofluid with spin

    velocity xy, the time average dissipated power is then

    < Pd > ¼l0v0H

    20X

    2f s 1þ X2f � x2y

    � �s2

    h i1þ 2 X2f þ x2y

    � �s2 þ X2f � x2y

    � �2s4

    � � :ð12Þ

    In a clockwise rotating field with Ĥx =H0 and Ĥz =� jH0,the dissipated power is

    < Pd > ¼l0v0H

    20Xf s Xf þ xy

    � �1þ Xf þ xy

    � �2s2

    h i1þ 2 X2f þ x2y

    � �s2 þ X2f � x2y

    � �2s4

    � � :ð13Þ

    Note that the magnetic susceptibility in Eqs. (9) and (10)

    and the dissipated power in Eqs. (11)– (13) can be

    modulated by magnetic field amplitude, frequency, phase

    and direction and by control of the spin velocity which itself

    can be set by flow vorticity and by magnetic field

    parameters such as amplitude and frequency of a rotating

    uniform magnetic field.

    2.4. Conservation of linear and angular momentum

    equations

    In time transient or dynamic flow conditions, fluid

    viscosity causes the magnetization M̄ to lag the magnetic

    field H̄. When M̄ and H̄ are not collinear, the torque density

    T=l0M̄� H̄ leads to novel flow phenomena because theviscous stress tensor has an anti-symmetric part. The general

    pair of force and torque equations for ferrofluids are then

    [1&&,2&&,3&,4&,5&,34&&]

    qBm̄mBt

    þ m̄mIlð Þm̄m�

    ¼ �lpþ l0M̄M IlH̄H þ 2fl� x̄x

    þ k þ g � fð Þl lIm̄mð Þ þ g þ fð Þl2m̄mð14Þ

    IBx̄xBt

    þ m̄mIlð Þx̄x�

    ¼ l0M̄M � H̄H þ 2f l� v̄v � 2x̄xð Þ

    þ k Vþ g Vð Þl lIx̄xð Þ þ g Vl2x̄xð15Þ

    where q is the fluid mass density, I is the fluid moment ofinertia density, v̄ is the linear flow velocity, x̄ is the spinvelocity, and p is the hydrodynamic pressure. The viscosity

    coefficients are the usual shear viscosity g and the dilationalviscosity k while gV and kV are the analogous shear and bulkcoefficients of spin viscosity. The coefficient f is called thevortex viscosity and from microscopic theory for dilute

    suspensions obeys the approximate relationship, f =1.5g/,

    where / is the volume fraction of particles [63,34&&]. Recentwork has performed computer simulations of ferrofluid

    laminar pipe flows to show magnetic field induced drag

    reduction [64,65] and fluorescent microscale particle image

    velocimetry was developed for ferrofluid micro-channel

    flows [66].

    Most practical ferrofluid flows are incompressible, so

    that l I v̄ =0, and are viscous dominated so that inertialeffects are negligible. Flow geometry often results in

    l I x̄ =0, as is the case in Fig. 1 where x̄ =xy(x)īy.Dimensional analysis gives g V̈ g: 2/2 where : is of the

    order of the distance between particles and / is the particleand surfactant volume fraction. Because, with usual volume

    fractions of order / =0.01, the distance : is comparable toparticle diameter of order 10 nm, gV becomes very small andso is often neglected in Eq. (15). In these limits and

    assuming viscous dominated flow, Eqs. (14) and (15) can be

    combined to the simpler forms:

    0,�lpþ l0M̄M IlH̄H þl02l� M̄M � H̄H

    � �þ gl2v̄v ð16Þ

    x̄x,l0 M̄M � H̄H� �

    4fþ 1

    2l� v̄v: ð17Þ

    Note that the spin velocity depends on magnetic torque

    and flow vorticity. Using Eqs. (16) and (17) in Eq. (5)

    yields

    BM̄M

    Btþ v̄vIlð ÞM̄M ¼ X̄X � M̄M �

    M̄M � M̄M 0� �

    s

    � 14f

    l0M̄M � M̄M� H̄H� �

    ; X̄X ¼ 12l� v̄v:

    ð18Þ

    Recent work by Rosensweig has extended Eqs. (14)

    and (15) based on integral balance equations and

    thermodynamics without arbitrary definitions of electro-

    magnetic energy density and stress. The analysis included

    electric and magnetic fields using the Minkowski expres-

    sion, D̄� B̄, for electromagnetic momentum density andused an appropriate formulation of entropy production in

    which the electromagnetic field is distinguished from

    equilibrium. The usual constitutive relationships result

    from the analysis without any empirical assumptions,

    including the magnetization relaxation Eqs. (5) and (18)

    [67&]. Rosensweig then further extended the analysis to

    include Galilean relativity effects to first order in the ratio

    of fluid velocity to light speed [68&].

    A similar analysis was performed by Felderhof,

    following the semirelativistic hydrodynamic equations of

    motion of de Groot and Mazur [69], based on the choice

    of e0l0Ē� H̄ for electromagnetic momentum density[36&,70&]. The Rosensweig and Felderhof analyses are

    on the whole very similar, and in the limit of small v /c,

    the two analyses are identical.

  • C. Rinaldi et al. / Current Opinion in Colloid & Interface Science 10 (2005) 141–157146

    3. Magnetic forces and torques

    3.1. Magnetic force density

    The magnetic force density acting on an incompressible

    ferrofluid is

    F̄F ¼ l0 M̄M Il� �

    H̄H ð19Þ

    so that ferrofluids move in the direction of increasing

    magnetic field strength. In a uniform magnetic field there is

    no magnetic force. When M̄ and H̄ are collinear, this force

    density modifies Bernoulli’s equation for inviscid and

    irrotational steady flow to [1&&]

    pþ 12

    qjv̄vj2 � qḡg Ir̄r � l0Z H0

    MdH ¼ constant ð20Þ

    where |v̄| is the magnitude of the fluid velocity, ḡ is the

    gravitational acceleration vector, and r̄=xīx+yīy + zīz is the

    position vector. The magnetic term is the magnetic

    contribution to fluid pressure and describes such magnetic

    Fig. 2. Ferrofluid instabilities in DC magnetic fields for an Isopar-M based ferroflu

    diameter magnet behind a small ferrofluid droplet surrounded by propanol to pre

    magnetic field perpendicular to ferrofluid layer. The peaks initiate in a hexagonal a

    weight and surface tension—(b) 200 Gauss, (c) 330 Gauss, (d) 400 Gauss; (e– f)

    Gauss vertical magnetic field [71,72].

    field effects as the shape of a ferrofluid meniscus; flow and

    instabilities of a ferrofluid jet such as for magnetic fluid

    inkjet printing; operation of magnetic fluid seals, bearings,

    load supports; sink-float separations; magnetic fluid

    nozzles; and magnet self-levitation in a ferrofluid

    [1&&,3&,4&].

    Ferrofluids exhibit a wide range of very interesting lines,

    patterns, and structures that can develop from ferrohydro-

    dynamic instabilities as shown in Fig. 2 [20&,71,72].

    The magnetic field gradient force density in Eq. (19)

    has been used for precise positioning and transport of

    ferrofluid for sealing, damping, heat transfer, and liquid

    delivery systems [73–75]. Smooth and continuous pump-

    ing of ferrofluid in a tube or channel can be achieved by

    a traveling wave non-uniform magnetic field generated by

    a spatially traveling current varying sinusoidally in time

    with a sinusoidal spatial variation along the duct axis

    [76,77]. Magnetic forces on ferrofluids have been used

    for adaptive optics to shape deformable mirrors [78,79];

    as a ferrofluid cladding layer in development of a tunable

    in-line optical-fiber modulator [80]; use as a flat panel

    display cell where the luminance is magnetically con-

    id with saturation magnetization of about 400 Gauss. (a) 1200 Gauss 5 mm

    vent ferrofluid wetting of 1 mm gap glass plates; (b–d) peak pattern with

    rray when the magnetic surface force exceeds the stabilizing effects of fluid

    labyrinth instability with ferrofluid between 1 mm gap glass plates in 250

  • C. Rinaldi et al. / Current Opinion in Colloid & Interface Science 10 (2005) 141–157 147

    trolled through the ferrofluid film thickness [81]; as a

    microfluidic MEMS-based light modulator [82]; for

    magnetic control of the Plateau rule of the angle between

    contacting films in 2D foams [83]; for magnetic control

    of bubble size and transport in ferrofluid foams for

    microfluidic and ‘‘lab-on-a-chip’’ applications and for

    possible zero gravity experiments using the magnetic

    field gradient force to precisely balance gravity so that

    bubbles are free floating in ferrofluid [84]; for magnetic

    field control of ferrofluid in a cavitating flow in a

    converging–diverging nozzle [85]; and for studies of

    longitudinal and transverse tangential magnetic fields on

    ferrofluid capillary rise [86]. Applied electric fields can

    also exert forces in ferrofluids to control suspension

    rheological properties [87–89].

    3.2. Torque-driven flows

    In a rotating magnetic field, the magnetization relaxation

    time constants of Eqs. (6) and (7) create a phase difference

    between magnetization and magnetic field so that M̄ and H̄

    are not in the same direction. This causes a magnetic torque

    density

    T̄T ¼ l0M̄M � H̄H ð21Þ

    which causes the magnetic nanoparticles and surrounding

    fluid to spin as given in Eq. (17). The concerted action

    of order 1022–1023 spinning nanoparticles/m3 can cause

    fluid pumping [90,91&,92&] and other interesting ferrofluid

    flows [93&,94] and microdrop behavior shown in Fig. 3

    [95&,96&,97&].

    To emphasize magnetic torque effects, the effective

    viscosity of ferrofluid was studied using the rotating uniform

    Fig. 3. A ferrofluid drop between Hele–Shaw cell glass plates with 1.1 mm gap has

    DC (0–250 Gauss) magnetic fields. The ferrofluid is surrounded by equal propor

    smearing. (a) The vertical DC field is first applied to form the labyrinth pattern and

    spiral pattern; (b) the counter-clockwise rotating field is applied first to hold the dr

    to about 100 Gauss, the continuous fluid drop abruptly transitions to discrete drop

    pattern further; (c) various end-states of spirals, and (d) droplet patterns [95&].

    magnetic field generated by a two-pole three-phase motor

    stator winding with a Couette viscometer used as a torque

    meter [47&,48&]. When a fixed spindle rotation speed is

    selected, the viscometer applies the necessary torque in order

    to keep it rotating at the specified speed. When the magnetic

    field co-rotates with the spindle immersed in ferrofluid, the

    magnetic-field-induced shear stress on the spindle is in the

    direction opposite to spindle rotation, making it harder to

    turn the spindle at the specified speed so that the viscometer

    applies a higher torque and it records an increase of effective

    ferrofluid viscosity as shown in Fig. 4 [48&]. When the

    magnetic field counter-rotates relative to the spindle, the

    magnetic-field-induced shear stress on the spindle is in the

    same direction as spindle rotation, so that it is easier to rotate

    the spindle at the specified speed; therefore the viscometer

    applies a lower torque, and the viscometer records a

    decrease of effective ferrofluid viscosity as also shown in

    Fig. 4. When the torque in Fig. 4 is negative, the effective

    viscosity is negative [48&,98–106]. Other torque measure-

    ments used a stationary spindle with ferrofluid inside the

    spindle, outside the spindle, or simultaneously inside and

    outside the spindle [48&]. The approximate magnetic-field-

    induced time average torque on the spindle for ferrofluid

    entirely outside the spindle is [48&]

    T,� 8kR2Lc2l0v0 1þ v0ð ÞH2Xs

    v0 þ 2ð Þ2 þ c2v0 4þ 2v0 þ v0c2ð Þ

    ð22Þ

    where R is the radius of the outer cylinder, c is the ratio ofthe inner cylinder radius to the outer radius of the

    ferrofluid container, Xf is the radian frequency of theapplied magnetic field with rms amplitude H, s is theeffective magnetization relaxation time, v0 is the equili-

    simultaneous applied in-plane rotating (20 Gauss rms at 25 Hz) and vertical

    tions by volume of water and isopropyl alcohol (propanol) to prevent glass

    then a counter-clockwise rotating field is applied to form a slowly rotating

    op together with no labyrinth and then as the DC magnetic field is increased

    lets, while further increase of the magnetic field to 250 Gauss changes the

  • Fig. 4. Torque required to rotate a spindle surrounded with ferrofluid as a

    function of rotating magnetic field amplitude, frequency, and rotation

    direction with respect to the 100 RPM rotation of the spindle [48&].

    C. Rinaldi et al. / Current Opinion in Colloid & Interface Science 10 (2005) 141–157148

    brium ferrofluid magnetic susceptibility, and l0=4p�10�7 henries/meter is the magnetic permeability of free

    space. The approximate magnetic field induced torque on

    the spindle for ferrofluid entirely inside the spindle is

    [107,48&]

    T,� 2Xf sv0l0H2V 1�sv0l0H

    2

    2f

    � �ð23Þ

    where f =1.5g/ is the vortex viscosity and g is the fluidviscosity. Magnetic field and shear dependent changes in

    ferrofluid viscosity have many possible applications in

    damping technologies [108&,109].

    Fig. 5. The change in magnetoviscosity versus related DC magnetic field Hx or ma

    imposed magnetic flux density Bx, magnetic susceptibility vm =1.55 was used in

    4. Ferrohydrodynamics

    4.1. Magnetoviscosity

    Magnetic fluids have been used in microfluidic and

    microchannel devices [110–112]. Theoretical analysis

    calculates the shear stress for the planar Couette flow in

    Fig. 1 with a uniform x directed DC magnetic field Hx or

    magnetic flux density Bx, to evaluate the DC effective

    magnetoviscosity [106]. If the ferrofluid fills an air-gap of

    length s and area A of an infinitely magnetically permeable

    magnetic circuit excited by an N turn coil carrying a DC

    current I0, then Hx=NI0 / s is spatially uniform for planar

    Couette flow only as the spin velocity xy is also uniform.For other flow profiles, Hx and xy will be a function ofposition inside the ferrofluid flow, while magnetic flux

    density Bx will always be spatially uniform in the planar

    channel as l I B̄ =0, no matter the flow profile in thegeometry of Fig. 1 [104,106]. Extensions of this work

    include AC magnetic fields for Couette and Poiseuille flows

    with magnetic field components along and perpendicular to

    the duct axis [104,113&].

    For the planar Couette flow shown in Fig. 1, with x

    directed magnetic field only (Hz =0), the change in

    magnetoviscosity

    Dg ¼ f 1þ 2xydV

    � �ð24Þ

    is plotted as the multivalued functions versus Hx or Bxmagnetic parameters in Fig. 5 [106]. For a DC magnetic

    field, the effective viscosity always increases with magnetic

    gnetic flux density Bx parameters PH or PB for various values of Vs /d. Forthe plots for oil-based ferrofluid Ferrotec EFH1 [106].

  • C. Rinaldi et al. / Current Opinion in Colloid & Interface Science 10 (2005) 141–157 149

    field amplitude. For AC magnetic fields, the effective

    viscosity can decrease, even through zero and negative

    values [48&,104].

    The size distribution of nanoparticles in ferrofluid is very

    important to various shear flows in a magnetic field. A

    gradient magnetic field causes large particles to diffuse to

    the strong field region while smaller particles have a weaker

    effect. Rheometer measurements of viscosity show that

    large particles and agglomerates have a strong magnetic

    field dependence on viscosity increase which depends on

    shear rate, while the smaller particle fluid has only a small

    viscosity increase and smaller shear-rate dependence

    [114,115].

    Another rotating magnetic probe was developed to

    measure the local viscoelasticity on microscopic scales

    based on the alignment of dipolar chains of submicron

    magnetic particles in the direction of an applied magnetic

    field [116]. Proposed continuing work is to investigate the

    rheological properties of the interior of a cell using this

    magnetic rotational microrheology method [117].

    4.2. Ultrasound velocity profile measurements

    Because ferrofluids are opaque, an ultrasound velocity

    profile method was developed for use with ferrofluids

    [118&,119&,120&,121&]. The sound velocity in ferrofluid

    increases slightly with magnetic field strength and ultra-

    sound frequency while the velocity slightly decreases with

    increasing angle between magnetic field and ultrasound

    propagation direction and with weight concentration of

    magnetic particles [122&].

    An Ultrasonic Doppler Velocimeter (UDV) measured the

    bulk velocity of ferrofluid in a cylindrical container, with

    and without a top, with flow driven by a uniform rotating

    magnetic field, as shown in Fig. 6. The measurements show

    that a counter-clockwise rotating magnetic field causes a

    Fig. 6. Spin-up flow profiles at mid-height excited by a magnetic field rotating co

    free top surface and (b) with a cover on the container, so that there is no free top su

    linear profiles of a fluid in rigid body co-rotation with the applied magnetic field,

    velocity of the bulk flows in the central region co-rotate with the respective mag

    surface angular velocities are larger and counter-rotate with the magnetic field at ¨

    central region co-rotate with the magnetic field at larger velocities than in (a) with

    rigid-body-like counter-clockwise rotation of ferrofluid at a

    constant angular velocity. Only in a thin layer near the

    cylinder wall does the no-slip condition force the fluid flow

    to differ from rigid-body rotation.

    The free ferrofluid/air interface at the surface of the vessel

    with no top is observed to rotate in the clockwise direction,

    opposite to the bulk rotation. The experiments show that

    while the ferrofluid bulk slowly co-rotates with the applied

    magnetic field, the ferrofluid free surface counter-rotates

    with the applied magnetic field at higher rotational speed

    than the bulk. With a fixed top on the ferrofluid free surface,

    the volume flow velocity is generally higher than for a free

    surface which counter-rotates to the volume flow.

    Rosensweig et al. observed the velocity of the free

    surface of ferrofluid flow in a uniform rotating magnetic

    field driven by magnetic shear stresses in meniscus regions

    [123&]. When the meniscus of the ferrofluid free surface was

    concave, the ferrofluid surface flow counter-rotated to the

    driven magnetic field; when the meniscus of ferrofluid free

    surface was flat, the ferrofluid surface flow was stationary;

    and when the meniscus of the ferrofluid free surface was

    convex, the ferrofluid surface flow co-rotated with the

    driven magnetic field. Because ferrofluid is opaque,

    measurement of the bulk flow profiles was not possible.

    All the surface flow observations were made by placing

    tracer particles on the top rotating surface. Based on these

    observations, Rosensweig concluded that volume flow

    effects were negligible. However, the ultrasound velocity

    measurements in Fig. 6 show that there is a significant

    volume flow, even without a free top interface.

    4.3. Interfacial phenomena

    Further experimental measurements of meniscus height,

    h =n(0), and shape n(x), were performed using narrowdiameter laser beam reflections from the meniscus interface

    unter-clockwise at 200 Hz for various magnetic field amplitudes; (a) with a

    rface. The central flow velocity profiles for (a) and (b) are approximately the

    dropping to zero at the wall over a thin boundary layer. In (a), the angular

    netic fields from high to low at ¨14, 8, 4, and 1 rpm while the respective

    48, 36, 22, and 9 rpm. In (b), the angular velocity of the bulk flows in the

    the respective magnetic fields from high to low at ¨20, 12, 5, and 1 rpm.

  • Fig. 7. The experimental optical setup for measuring ferrofluid meniscus

    shape [124&].

    C. Rinaldi et al. / Current Opinion in Colloid & Interface Science 10 (2005) 141–157150

    as a function of DC applied magnetic fields, as shown in

    Fig. 7. The angle of deflection of the laser beam from the

    vertical allows the computation of the shape of the ferrofluid

    meniscus [124&].

    Three applied magnetic field configurations were used,

    as shown in Fig. 8. A meniscus forms on the sides of a glass

    slide immersed in the center of a vessel with negligibly

    small field gradients. Fig. 9 shows for both oil and water-

    based ferrofluids that applied magnetic field in configura-

    tion a has practically no effect on the meniscus shape. As

    expected, with a uniform magnetic field tangent to the

    interface, the shape of the interface matches the non-

    magnetic meniscus shape

    x

    a¼ 1ffiffiffi

    2p cosh�1

    ffiffiffi2

    p

    n=a� cosh�1

    ffiffiffi2

    p

    h=a

    þ

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2� h

    a

    � 2s�

    ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2� n

    a

    � 2sð25Þ

    where a =(2r /qg)1 / 2 is the capillary length for ferrofluid ofmass density q and surface tension r, with g being theacceleration of gravity. Fig. 9 shows that applied horizontal

    magnetic field in configuration b lowers the height of a

    Fig. 8. The three magnetic field configurations of the ferr

    meniscus, while in vertical magnetic fields the meniscus

    rises (configuration c). An approximate minimization of free

    energy model was used including magnetization, surface

    tension, and gravitation energies, to calculate meniscus

    height change with magnetic field which compared well to

    measurements [124&]. Other work used an optical reflection

    method to measure the height of the perpendicular field

    instability peaks, like those shown in Fig. 2b–d [125].

    The surface tension at an air/ferrofluid interface or at a

    non-magnetic liquid (propanol)/ferrofluid interface was

    measured using the perpendicular field instability of Fig.

    2b–d by measuring the fluid peak spacing and magnetic

    field strength at the incipience of instability [126].

    Other work used a 2D lattice Boltzmann model to account

    for the competition between interfacial tension and dipolar

    forces in ferrofluids [127]; experiments were performed on

    the time evolution of the breakup of a liquid bridge of

    ferrofluid in an external magnetic field as it disintegrates,

    including study of the dynamics of satellite drops emanating

    from the liquid bridge [128]; while other measurements

    compared the shape response of ferrofluid droplets, a

    magnetorheological fluid, and a composite magnetic fluid

    with nanometer and micron sized particles in low frequency

    alternating magnetic fields up to 5 Hz [129].

    The Weissenberg effect is the rise of a free surface of a

    viscoelastic fluid at a rotating rod due to normal stress forces

    [130]. For magnetic fluids the height of the fluid surface at

    the rod also depends on the length and quantity of fluid

    particle chains, the strength of the applied magnetic field

    parallel to the surface, the concentration of larger magnetic

    particles, and the shear rate.

    4.4. Hele–Shaw cell flows

    In a porous medium, the local average fluid velocity is

    described by Darcy’s law

    0 ¼ �lp� bv̄v þ F̄F ð26Þ

    where p is the hydrodynamic pressure, b =g /j is the ratioof fluid viscosity g to the permeability j which depends onthe geometry of the particles and interstitial space, and F̄ is

    the total external force density, typically due to gravity and

    magnetization given by Eq. (19).

    A Hele–Shaw cell consists of two parallel walls a small

    distance d apart, and is used for two-dimensional flows

    where the flow velocity only has components parallel to the

    ofluid meniscus experimentally investigated [124&].

  • Fig. 9. Measured meniscus shape for the three configurations of applied magnetic field. The top row shows oil-based ferrofluid (v0�1.55) results while thebottom row shows water-based ferrofluid (v0�0.67) results [124&].

    C. Rinaldi et al. / Current Opinion in Colloid & Interface Science 10 (2005) 141–157 151

    walls. The mean flow between the walls is then also

    described by Eq. (26) where b =12g /d2 and j =d2 /12. Aferrofluid drop in a Hele–Shaw cell forms an intricate

    labyrinth when a uniform DC magnetic field is applied

    perpendicular to the walls as shown in Fig. 2e–f. Typical

    labyrinth analyses are usually approximately based on a

    minimization of the free energy as a sum of magnetostatic

    and interfacial tension energies, including the effects of

    demagnetizing magnetic fields typically assuming that the

    magnetization is linear with H̄, as in Eq. (2). Recent work

    has generalized the free energy analysis to nonlinear

    magnetization characteristics, particularly using the Lange-

    vin magnetization of Eq. (1) [131].

    Fig. 3 shows interesting Hele–Shaw cell ferrofluid

    drop responses to simultaneous vertical DC and horizontal

    rotating uniform magnetic fields [95&,96&,97&]. Other

    ferrofluid Hele–Shaw cell flows investigated azimuthal

    magnetic fields from a long straight current-carrying wire

    along the axis perpendicular to the walls with a time-

    dependent gap [132,133]; labyrinthine instability in mis-

    cible magnetic fluids in a horizontal Hele–Shaw cell with

    a vertical magnetic field [134]; theory and experiments of

    interacting ferrofluid drops [135&]; numerical simulations

    [136] and measurements [137] of viscous fingering

    labyrinth instabilities; fingering instability of an expanding

    air bubble in a horizontal Hele–Shaw cell [138] and of a

    rising bubble in a vertical Hele–Shaw cell [139]; finger-

    ing instabilities of a miscible magnetic fluid droplet in a

    rotating Hele–Shaw cell [140]; theory and experiments of

    the Rayleigh–Taylor instability [141&]; and natural con-

    vection of magnetic fluid in a bottom-heated square

    Hele–Shaw cell with two insulated side walls with heat

    transfer measurements and liquid crystal thermography

    [142].

    4.5. Flow instabilities

    4.5.1. Jet and sheet flows

    The behavior of an initially circular cross-section

    ferrofluid jet impacting a solid circular surface to create

    an expanding sheet flow in the presence of a magnetic field

    was investigated [143]. A horizontal magnetic field trans-

    verse to a vertical ferrofluid jet axis changes the jet cross-

    section from circular to elliptical, with long axis in the

    direction of the applied magnetic field. As the transverse

    magnetic field is increased, the expanding sheet also

    becomes approximately elliptically shaped but with long

    axis perpendicular to the magnetic field, which is transverse

    to the long axis of the jet cross-section, as shown in Fig. 10.

    At large magnetic fields, the sheet forms sharp tips and fluid

    chains emerge from its corners. The prime cause of the

    change in sheet shape is the influence of the magnetic field

    on the jet shape. If an initially elliptical non-magnetic jet

    strikes a circular impactor, the expanding sheet is similarly

    elliptical but with its long axis also perpendicular to that of

    the jet.

    If the applied magnetic field is vertical, and thus

    perpendicular to the expanding circular sheet, the sheet

  • (a) B≈0 Gauss (b) B≈200 Gauss

    (c) B≈600 Gauss (d) B≈1200 Gauss

    Fig. 10. A vertical oil-based ferrofluid jet, with diameter 2.5 mm, impacts a small circular horizontal plate of 10 mm diameter creating a radially expanding thin

    sheet flow. (a) In zero magnetic field, a circular jet will create a circular sheet; (b) application of the magnetic field transverse to the jet, in the direction of the

    arrow, causes the jet cross-section to elongate in the direction of the applied field while the sheet distorts to an approximately elliptical shape, but with long-axis

    perpendicular to the applied magnetic field. (c, d) For large magnetic fields, the sheet becomes very thin and is characterized by sharp tips and fluid chains

    emerging from its corners [143].

    C. Rinaldi et al. / Current Opinion in Colloid & Interface Science 10 (2005) 141–157152

    diameter decreases with increasing magnetic field B. A

    simple approximate Bernoulli analysis including magnetic

    surface forces predicts the sheet radius to be

    r ¼ rT 1�B2

    qU21

    l0� 1

    l

    � � �; rT ¼

    qUQ4kr

    ð27Þ

    where rT is the Taylor radius with B =0, l is the ferrofluidmagnetic permeability, l0 is the magnetic permeability ofthe free space surrounding the sheet, q is the ferrofluiddensity, U is the ferrofluid sheet velocity, Q is the jet flow

    rate, and r is the interfacial surface tension [143].

    4.5.2. Kelvin–Helmholtz instability

    The classic instability of the planar interface between

    two superposed fluids with a relative horizontal velocity is

    called the Kelvin–Helmholtz instability. The moving

    ferrofluid sheet flow of Section 4.5.1 has magnetically

    coupled upper and lower interfaces which can become

    Kelvin–Helmholtz unstable. A planar sheet model,

    generalized to include a magnetic field B perpendicular

    to both interfaces, is formulated to examine the effects of

    magnetic fields on interfacial stability for a fluid layer of

    thickness d moving with horizontal velocity U [143].Letting n1 be the upwards deflection of the upperinterface and n2 be the upwards deflection of the lowerinterface, a linear stability analysis shows that there are

    two interfacial modes with n1 /n2=T1. For interfacial

    deflections of the form n1,2=Re[n̂1,2ej(xt-kz)] the disper-

    sion relation is

    q0x2

    kþ q x � kUð Þ

    2

    k

    coth kd=2ð Þtanh kd=2ð Þ

    � � rk2

    þk l � l0ð Þ2B2 coshkd þ ll0 sinhkdF1

    h ilsinhkd l2 þ l20 þ 2ll0cothkd

    � �¼ 0 n1=n2 ¼ k1 ð28Þ

    where q0 is the density of the surrounding air. Interfacialdeflections are unstable if x22. Application of a mag-netic field component parallel to the sheet interfaces tends

    to stabilize the sheet flow [143].

    The stability of Kelvin–Helmholtz waves propagating

    on the interface between two magnetic fluids stressed by

    an oblique magnetic field has also been analyzed for free

    fluids [144] and for fluids streaming through porous media

    [145&].

    4.5.3. Other flow instabilities

    The Rayleigh–Maragoni–Bénard instability of a ferro-

    fluid in a vertical magnetic field was studied combining

  • & of special interest.&& of outstanding interest.

    C. Rinaldi et al. / Current Opinion in Colloid & Interface Science 10 (2005) 141–157 153

    aspects of volume and surface forces due to heating, flow

    rheology, and magnetism [146].

    The Faraday instability is the parametric generation of

    standing waves on the free surface of a fluid subjected to

    vertical vibrations. The initially flat free surface of the fluid

    becomes unstable at a critical intensity of the vertical

    vibrations. The Faraday instability has been used to study

    the rheological properties of ferrofluids caused by strong

    dipole interactions that lead to the formation of magnetic

    particle chains elongated in the direction of the horizontal

    magnetic field [147]; and used with a vertical oscillating

    magnetic field to parametrically excite standing surface

    waves with no mechanically imposed vertical vibrations

    together with a vertical DC magnetic field to tune the fluid

    parameters [148]. Later work used both a vertical DC

    magnetic field with a mechanical exciter to vertically shake

    the test cell [149] to obtain a novel pattern of standing twin

    peaks due to the simultaneous excitation of two different

    wave numbers.

    4.6. Thermodiffusion (Soret effect)

    In a multicomponent mixture, a temperature gradient

    leads to diffusive concentration gradients due to the

    coupling between heat and mass transport, known as the

    Soret effect. The Soret effect has a strong dependence on

    magnetic field in ferrofluids [150–157&] and the sign is

    controlled by the electrostatic charge or surfactant at the

    nanoparticle interface [158,159]. A temperature gradient

    imposed across a flat ferrofluid layer with a transverse

    magnetic field causes a concentration gradient of magnetic

    nanoparticles. The mass and temperature gradient causes a

    magnetization gradient resulting in a magnetic field

    gradient which redistributes the magnetic particles with

    mixing. Unlike a homogeneous single component fluid

    which has a stationary instability, the two component

    ferrofluid system has a double-diffusive oscillatory insta-

    bility [160].

    Recent work used a vertical laser beam to create a hot

    spot on a thin horizontal ferrofluid layer [161]. With a

    positive Soret coefficient the magnetic nanoparticles dif-

    fused to colder regions leaving the hot spot transparent.

    With a vertical magnetic field the initially round hot spot

    changed to various polygon shapes, said to be caused by

    convectively unstable roll cells. Shliomis [162] offered

    another explanation where the hot spot is initially a non-

    magnetic ‘‘bubble’’ within a magnetic fluid which is the

    inverse of the labyrinth instability shown in Figs. 3a and

    2e–f. Using the magnetic Bond number criterion for

    labyrinth instability, Shliomis obtained very good agreement

    with the labyrinth theory.

    Calculation of the thermomagnetic force on a prolate

    spheroidal body immersed in ferrofluid was used as a model

    of long magnetic nanoparticle aggregates in a non-uniformly

    heated ferrofluid in an applied magnetic field with temper-

    ature gradient at an angle to the magnetic force [163].

    5. Conclusions

    This magnetic fluid rheology and flows review has

    presented recent advances in fundamental theory; ferrofluid

    flows and instabilities; thermal effects and applications; and

    practical device and biomedical applications. Of special

    importance are the better understanding of the significant

    effects of magnetic particle aggregation; modern electrical,

    mechanical, optical, and radiation measurement methods to

    determine important rheological parameters; extensions of

    the governing magnetization and ferrohydrodynamic equa-

    tions including theoretical and experimental investigations

    and applications of the antisymmetric part of the viscous

    stress tensor; magnetic control of the magnetic susceptibility

    tensor for heating applications with ferrofluids with non-

    zero spin velocity; recent advances in biomedical applica-

    tions of drug delivery, hyperthermia, and magnetic reso-

    nance imaging; effective magnetoviscosity analysis and

    measurements, including zero and negative values and not

    just viscosity reduction; ultrasound velocity profile meas-

    urements of spin-up flow showing counter-rotating surface

    and volume flows in a uniform rotating magnetic field;

    theory and measurements of ferrofluid meniscus shape for

    tangential and perpendicular magnetic fields; and new

    analysis and measurements of ferrohydrodynamic flows,

    instabilities, and thermodiffusion (Soret effect) phenomena.

    Acknowledgments

    For the MIT authors, this work was partially supported

    by the Thomas and Gerd Perkins Professorship at MIT and

    by generous MIT alumnus Thomas F. Peterson. Ferrotec

    Corporation is acknowledged for providing ferrofluids used

    in MIT experiments and Brookfield Corporation is acknowl-

    edged for providing a viscometer. The University of Puerto

    Rico at Mayaguez authors were supported by the US

    National Science Foundation (CTS-0331379) and by the

    Petroleum Research Fund (ACS-PRF 40867-G 9).

    References and recommended readings

    [1]&&

    R.E. Rosensweig, Ferrohydrodynamics. New York: Cambridge

    University Press, 1985: republished by New York: Dover Publica-

    tions, 1997.

    [2]&&

    S. Odenbach, Magnetoviscous Effects in Ferrofluids, Springer,

    Berlin, 2002.

    [3]&

    B. Berkovski, V. Bashtovoy (Eds.), Magnetic Fluids and Applica-

    tions, Begell House, New York, 1996.

    [4]&

    B.M. Berkovsky, V.F. Medvedev, M.S. Krakov, Magnetic Fluids

    Engineering Applications, Oxford University Press, Oxford, 1993.

    [5]&

    E. Blums, A. Cebers, M.M. Maiorov, Magnetic Fluids, Walter de

    Gruyter, Berlin, 1997.

  • C. Rinaldi et al. / Current Opinion in Colloid & Interface Science 10 (2005) 141–157154

    [6]&

    S.W. Charles, The preparation of magnetic fluids, in: S. Odenbach

    (Ed.), Ferrofluids, Magnetically Controllable Fluids and Their

    Applications, Springer, Berlin, 2002, pp. 3–18.

    [7] I. Bolshakova, M. Bolshakov, A. Zaichenko, A. Egorov, The

    investigation of the magnetic fluid stability using the devices with

    magnetic field microsensors, JMMM 289 (2005) 108–110.

    [8] G. Bossis, O. Volkova, S. Lacis, A. Meunier, Magnetorheology:

    fluids, structures and rheology, in: S. Odenbach (Ed.), Ferrofluids,

    Magnetically Controllable Fluids and Their Applications, Springer,

    Berlin, 2002, pp. 202–230.

    [9] A. Zubarev, Statistical physics of non-dilute ferrofluids, in: S.

    Odenbach (Ed.), Ferrofluids, Magnetically Controllable Fluids and

    Their Applications, Springer, Berlin, 2002, pp. 143–161.

    [10] M.I. Morozov, M.I. Shliomis, Magnetic fluid as an assembly of

    flexible chains, in: S. Odenbach (Ed.), Ferrofluids, Magnetically

    Controllable Fluids and Their Applications, Springer, Berlin, 2002,

    pp. 162–184.

    [11] A.Y. Zubarev, L.Y. Iskakova, On the theory of structural trans-

    formations in magnetic fluids, Colloid J 65 (6) (2003) 703–710.

    [12] A.O. Ivanov, S.S. Kantorovich, Structure of chain aggregates in

    ferrocolloids, Colloid J 65 (2) (2003) 166–176.

    [13] A. Wiedenmann, Magnetic and crystalline nanostructures in ferro-

    fluids as probed by small angle neutron scattering, in: S. Odenbach

    (Ed.), Ferrofluids, Magnetically Controllable Fluids and Their

    Applications, Springer, Berlin, 2002, pp. 33–58.

    [14] F. Gazeau, E. Dubois, J.-C. Bacri, Boué, A. Cebers, R. Perzynski,

    Anisotropy of the structure factor of magnetic fluids under a field

    probed by small-angle neutron scattering, Phys Rev E 65 (2002)

    1–15 (031403).

    [15] M.H. Sousa, E. Hasmonay, J. Depeyrot, F.A. Tourinho, J.-C.

    Bacri, E. Dubois, et al., NiFe2O4 nanoparticles in ferrofluids:

    evidence of spin disorder in the surface layer, JMMM 242-245

    (2002) 572–574.

    [16] Y.L. Raikher, V.I. Stepanov, J.-C. Bacri, R. Perzynski, Orientational

    dynamics in magnetic fluids under strong coupling of external and

    internal relaxations, JMMM 289 (2005) 222–225.

    [17] Y.L. Raikher, V.I. Stepanov, J.-C. Bacri, R. Perzynski, Orientational

    dynamics of ferrofluids with finite magnetic anisotropy of the

    particles: relaxation of magneto-birefringence in crossed fields, Phys

    Rev E 66 (2002) 1–14 (021203).

    [18] D.K. Kim, D. Kan, T. Veres, F. Normadin, J.K. Liao, H.H. Kim, et

    al., Monodispersed Fe–Pt nanoparticles for biomedical applica-

    tions, J Appl Phys 97 (2005) 1–3 (10Q918).

    [19] F. Shen, C. Poncet-Legrand, S. Somers, A. Slade, C. Yip, A.M.

    Duft, et al., Properties of a novel magnetized alginate for

    magnetic resonance imaging, Biotechnol Bioeng 83 (3) (2003)

    282–292.

    [20]&

    C. Rinaldi, T. Franklin, M. Zahn, T. Cader, Magnetic nanoparticles

    in fluid suspension: ferrofluid applications, Dekker Encyclopedia of

    Nanoscience and Nanotechnology, Marcel Dekker, New York, 2004,

    pp. 1731–1748.

    [21]&&

    S. Mornet, S. Vasseur, F. Grasset, E. Duguet, Magnetic nanoparticle

    design for medical diagnosis and therapy, J Mater Chem 14 (2004)

    2161–2175.

    [22] R. Ganguly, A.P. Gaind, S. Sen, I.K. Puri, Analyzing ferrofluid

    transport for magnetic drug targeting, JMMM 289 (2005) 331–334.

    [23] R. Müller, R. Hergt, M. Zeisberger, W. Gawalek, Preparation of

    magnetic nanoparticles with large specific loss power for heating

    applications, JMMM 289 (2005) 13–16.

    [24] J.W.M. Bulte, Magnetic nanoparticles as markers for cellular MR

    imaging, JMMM 289 (2005) 423–427.

    [25] L.F. Gamarra, G.E.S. Brito, W.M. Pontuschka, E. Amaro, A.H.C.

    Parma, G.F. Goya, Biocompatible superparamagnetic iron oxide

    nanoparticles used for contrast agents: a structural and magnetic

    study, JMMM 289 (2005) 439–441.

    [26] G. Glockl, V. Brinkmeier, K. Aurich, E. Romanus, P. Weber, W.

    Weitschies, Development of a liquid phase immunoassay by time-

    dependent measurements of the transient magneto-optical birefrin-

    gence using functionalized magnetic nanoparticles, JMMM 289

    (2005) 480–483.

    [27] R. Aquino, J.A. Gomes, F.A. Tourinho, E. Dubois, R. Perzynski,

    G.J. da Silva, et al., Sm and Y radiolabeled magnetic fluids:

    magnetic and magneto-optical characterization, JMMM 289 (2005)

    431–434.

    [28] O. Brunke, S. Odenbach, C. Fritsche, I. Hilger, W.A. Kaiser,

    Determination of magnetic particle distribution in biomedical

    applications by X-ray microtomography, JMMM 289 (2005)

    428–430.

    [29] T.A. Osman, G.S. Nada, Z.S. Safar, Effect of using current-carrying-

    wire models in the design of hydrodynamic journal bearings

    lubricated with ferrofluids, Tribol Lett 11 (2001) 61–70.

    [30] S. Jakabsky, M. Lovas, A. Mockovciakova, S. Hredzak, Utilization

    of ferromagnetic fluids in mineral processing and water treatment, J

    Radioanal Nucl Chem 246 (3) (2000) 543–547.

    [31]&&

    M.I. Shliomis, Magnetic fluids, Soviet Phys Uspekhi 17 (2) (1974)

    153–169.

    [32] R.W. Chantrell, J. Popplewell, S.W. Charles, Measurements of

    particle size distribution parameters in ferrofluids, IEEE Trans Magn

    14 (1978) 975–977.

    [33] P. Kop*anský, M. Timko, I. Poto*ova, M. Koneracká, A. Juŕiková, N.Tomaxovi*ova, et al., The determination of the hydrodynamicdiameter of magnetic particles using FRS experiment, JMMM 289

    (2005) 97–100.

    [34]&&

    M.I. Shliomis, Effective viscosity of magnetic suspensions, Soviet

    Phys JETP 34 (6) (1972) 1291–1294.

    [35]&&

    M.A. Martsenyuk, Y.L. Raikher, M.I. Shliomis, On the kinetics of

    magnetization of suspensions of ferromagnetic particles, Soviet

    Phys JETP 38 (2) (1974) 413–416.

    [36]&

    B.U. Felderhof, H.J. Kroh, Hydrodynamics of magnetic and

    dielectric fluid in interaction with the electromagnetic field, J Chem

    Phys 110 (15) (1999) 7403–7411.

    [37]&

    B.U. Felderhof, Magnetoviscosity and relaxation in ferrofluids, Phys

    Rev E 62 (3) (2000) 3848–3854.

    [38]&

    M. Shliomis, Comment on Fmagnetoviscosity and relaxation inferrofluids_, Phys Rev E 64 (2001) 1–6 (063501).

    [39]&

    M.I. Shliomis, Ferrohydrodynamics: retrospective and issues, in:

    (Ed.), Ferrofluids, Magnetically Controllable Fluids and Their

    Applications, Springer, Berlin, 2002, pp. 85–111.

    [40]&

    B.U. Felderhof, Reply to ‘‘comment on Fmagnetoviscosity and

    relaxation in ferrofluids_’’, Phys Rev E 6406 (2001) 1–4 (063502).

    [41]&

    R. Patel, R.V. Upadhyay, R.V. Mehta, Viscosity measurements of a

    ferrofluid: comparison with various hydrodynamic equations, J Col-

    loid Interface Sci 263 (2003) 661–664.

    [42] A. Zeuner, R. Richter, I. Rehberg, Weak periodic excitation of a

    magnetic fluid capillary flow, JMMM 201 (1999) 321–323.

    [43]&

    M. Shliomis, Ferrohydrodynamics: testing a third magnetization

    equation, Phys Rev E 64 (R) (2001) 1–4 (060501).

    [44] H.W. Müller, M. Liu, Ferrofluid dynamics, in: S. Odenbach (Ed.),

    Ferrofluids, Magnetically Controllable Fluids and Their Applica-

    tions, Springer, Berlin, 2002, pp. 112–123.

    [45]&

    A.F. Lehlooh, S.H. Mahmood, J.M. Williams, On the particle size

    dependence of the magnetic anisotropy energy, Physica B 321 (2002)

    159–162.

    [46] P.C. Fannin, I. Malaescu, C.N. Marin, The effective anisotropy

    constant of particles within magnetic fluids as measured by magnetic

    resonance, JMMM 289 (2005) 162–164.

    [47]&

    A.D. Rosenthal, C. Rinaldi, T. Franklin, M. Zahn, Torque

    measurements in spin-up flow of ferrofluids, J Fluids Eng 126

    (2004) 198–205.

    [48]&

    C. Rinaldi, F. Gutman, X. He, A.D. Rosenthal, M. Zahn, Torque

    measurements on ferrofluid cylinders in rotating magnetic fields,

    JMMM 289 (2005) 307–310.

    [49] A. Engel, P. Reimann, Thermal ratchet effects in ferrofluids, Phys

    Rev E 70 (2004) 1–15 (051107).

  • C. Rinaldi et al. / Current Opinion in Colloid & Interface Science 10 (2005) 141–157 155

    [50] A. Engel, H.W. Müller, P. Reimann, A. Jung, Ferrofluids as thermal

    ratchets, Phys Rev Lett 91 (6) (2003) 1–4 (060602).

    [51] M.I. Shliomis, Comment on ‘‘ferrofluids as thermal ratchets’’, Phys

    Rev Lett 92 (18) (2003) 1, (188901).

    [52] P.C. Fannin, C.N. Marin, I. Malaescu, A.T. Giannitsis, Micro-

    wave absorption of composite magnetic fluids, JMMM 289

    (2005) 78–80.

    [53] A.T. Giannitsis, P.C. Fannin, S.W. Charles, Nonlinear effects in

    magnetic fluids, JMMM 289 (2005) 165–167.

    [54]&

    P.C. Fannin, Characterisation of magnetic fluids, J Alloys Compd

    369 (2004) 43–51.

    [55] P.C. Fannin, Y.P. Kalmykov, S.W. Charles, Investigation of subsid-

    iary loss-peaks in the frequency dependent susceptibility profiles of

    magnetic fluids, JMMM 289 (2005) 133–135.

    [56] P.C. Fannin, Magnetic spectroscopy as an aide in understanding

    magnetic fluids, in: S. Odenbach (Ed.), Ferrofluids, Magnetically

    Controllable Fluids and Their Applications, Springer, Berlin, 2002,

    pp. 19–32.

    [57] B. Fischer, B. Huke, M. Lücke, R. Hempelmann, Brownian

    relaxation of magnetic colloids, JMMM 289 (2005) 74–77.

    [58] Y.L. Raikher, V.V. Rusakov, Theoretical models for AC-field

    probing of complex magnetic fluids, JMMM 258–259 (2003)

    459–463.

    [59] D.K. Kim, M.S. Amin, S. Elborai, S.-H. Lee, Y. Koseoglu, M. Zahn,

    et al., Energy absorption of superparamagnetic iron oxide nano-

    particles by microwave irradiation, J Appl Phys 97 (2005) 1–3

    (10J510).

    [60] R.E. Rosensweig, Heating magnetic fluid with an alternating

    magnetic field, JMMM 252 (2002) 370–374.

    [61] R.E. Rosensweig, Role of internal rotations in selected magnetic

    fluid applications, Magnetohydrodynamics 36 (4) (2000) 303–316.

    [62] P.C. Fannin, C.N. Marin, I. Malaescu, A.T. Giannitsis, Micro-

    wave absorption of composite magnetic fluids, JMMM 289

    (2005) 78–80.

    [63] H. Brenner, Rheology of two-phase systems, Annual Rev Fluid Mech

    2 (1970) 137–176.

    [64] D.M. Ramos, F.R. Cunha, Y.D. Sobral, J.L.A. Fontoura Rodrigues,

    Computer simulations of magnetic fluids in laminar pipe flows,

    JMMM 289 (2005) 238–241.

    [65] F.R. Cunha, Y.D. Sobral, Asymptotic solution for pressure-driven

    flows of magnetic fluids in pipes, JMMM 289 (2005) 314–317.

    [66] N.T. Nguyen, Z.G. Wu, X.Y. Huang, C.-Y. Wen, The application of

    APIV technique in the study of magnetic flows in a micro-channel,JMMM 289 (2005) 396–398.

    [67]&

    R. Rosensweig, Basic equations for magnetic fluids with internal

    rotations, in: S. Odenbach (Ed.), Ferrofluids, Magnetically Controllable

    Fluids and Their Applications, Springer, Berlin, 2002, pp. 61–84.

    [68]&

    R. Rosensweig, Continuum equations for magnetic and dielec-

    tric fluids with internal rotations, J Chem Phys 121 (3) (2004)

    1228–1242.

    [69] S.R. de Groot, P. Mazur, Non-equilibrium Thermodynamics, North-

    Holland Amsterdam, 1962, also re-published by New York: Dover,

    1984.

    [70]&

    B.U. Felderhof, Relativistic hydrodynamics of magnetic and dielec-

    tric fluids in interaction with the electromagnetic field, J Chem Phys

    120 (8) (2004) 3598–3603.

    [71] C. Rinaldi, M. Zahn, Ferrohydrodynamic instabilities in dc magnetic

    fields, J Vis 7 (1) (2004) 8.

    [72] M. Zahn, Magnetic fluid and nanoparticle applications to nano-

    technology, J Nanopart Res 3 (2001) 73–78.

    [73] W. He, S.J. Lee, D.C. Jiles, D.H. Schimidt, D. Porter, R. Shinar,

    Design of high-magnetic field gradient sources for controlling

    magnetically induced flow of ferrofluid in microfluidic systems, J

    Appl Phys 93 (10) (2003) 7459–7461.

    [74] Y. Melikhov, S.J. Lee, D.C. Jiles, D.H. Schimidt, M.D. Porter, R.

    Shinar, Microelectromagnetic ferrofluid-based actuator, J Appl Phys

    93 (10) (2003) 8438–8440.

    [75] N. Saga, T. Nakamura, Elucidation of propulsive force of

    microrobot using magnetic fluid, J Appl Phys 91 (10) (2002)

    7003–7005.

    [76]&

    L. Mao, H. Koser, Ferrohydrodynamic pumping in spatially

    traveling sinusiodally time-varying magnetic fields, JMMM 289

    (2005) 199–202.

    [77] K. Zimmermann, I. Zeidis, V.A. Naletova, V.A. Turkov, V.E.

    Bachurin, Locomotion based on a two-layers flow of magnetizable

    nanosuspensions, JMMM 290-291 (part2) (2005) 808–810.

    [78] P.R. Laird, R. Bergamasco, V. Bérubé, E.F. Borra, J. Gingras, A.

    Ritcey, et al., Ferrofluid based deformable mirrors—a new approach

    to adaptive optics using liquid mirrors, in: P.L. Wizinowich, D.

    Bonaccini (Eds.), Adaptive Optical System. Technologies II, Pro-

    ceedings of SPIEvol. 4839, SPIE, WA, 2003, pp. 733–740.

    [79] P.R. Laird, E.F. Borra, R. Begamasco, J. Gingras, L. Truong, A.M.

    Ritcey, Deformable mirrors based on magnetic liquids, in: D.B.

    Calia, B.L. Ellerbroek, (Eds.), Advancements in Adaptive Optics,

    Proceedings of SPIEvol. 5490, PIE, WA, 2004, pp. 1493–1501.

    [80] J.J. Chieh, S.Y. Yang, Y.H. Chao, H.E. Horng, C.-Y. Hong, H.C.

    Yang, Dynamic response of optical-fiber modulator by using

    magnetic fluid as a cladding layer, J Appl Phys 97 (2005) 1–4

    (043104).

    [81] J.-W. Seo, S.-M. Jeon, S.H. Park, H.-S. Lee, An experimental and

    numerical investigation of flat panel display cell using magnetic

    fluid, JMMM 252 (2002) 353–355.

    [82] J.-W. Seo, H. Kim, S. Sung, Design and fabrication of a magnetic

    microfluidic light modulator using magnetic fluid, JMMM 272-276

    (2004) e1787–e1789.

    [83] E. Janiaud, J.-C. Bacri, A. Cebers, F. Elias, Magnetic forces in 2D

    foams, JMMM 289 (2005) 215–218.

    [84] W. Drenckhan, F. Elias, S. Hutzler, D. Weaire, E. Janiaud, J.-C.

    Bacri, Bubble size control and measurement in the generation of

    ferrofluid foams, J Appl Phys 93 (12) (2003) 10078–10083.

    [85] J. Ishimoto, S. Kamiyama, Numerical study of cavitating flow of

    magnetic fluid in a vertical converging–diverging nozzle, JMMM

    289 (2005) 26–263.

    [86] V. Bashtovoi, G. Bossis, P. Kuzhir, A. Reks, Magnetic field effect on

    capillary rise of magnetic fluids, JMMM 289 (2005) 376–378.

    [87] J.G. Veguera, Y.I. Dikansky, Periodical structure in a magnetic fluid

    under the action of an electric field and with a shear flow, JMMM

    289 (2005) 87–89.

    [88] V.A. Naletova, V.A. Turkov, V.V. Sokolov, A.N. Tyatyuskin, The

    interaction of the particles of magnetizable suspensions with a wall

    (wall-adjacent effect) in uniform electric and magnetic fields,

    JMMM 289 (2005) 367–369.

    [89] V.A. Naletova, V.A. Turkov, A.N. Tyatyuskin, Spherical body in a

    magnetic fluid in uniform electric and magnetic fields, JMMM 289

    (2005) 370–372.

    [90] R. Krauh, B. Reimann, R. Richter, I. Rehberg, Fluid pumped bymagnetic stress, Appl Phys Lett 86 (2005) 1–3 (024102).

    [91]&

    A.P. Krekhov, M.I. Shliomis, S. Kamiyama, Ferrofluid pipe flow in

    an oscillating magnetic field, Phys Fluids 17 (2005) 1–8 (033105).

    [92]&

    K.R. Schumacher, I. Sellien, G.S. Knoeke, T. Cader, B.A. Finlayson,

    Experiment and simulation of laminar and turbulent ferrofluid pipe

    flow in an oscillating magnetic field, Phys Rev E 67 (2003) 1–11

    ([026308]).

    [93]&

    B.U. Felderhof, Flow of a ferrofluid down a tube in an oscillating

    magnetic field, Phys Rev E 64 (2001) 1–7 ([021508]).

    [94] M.I. Shliomis, M.A. Zaks, Nonlinear dynamics of a ferrofluid

    pendulum, Phys Rev Lett 93 (4) (2004) 1–4 (047202).

    [95]&

    C. Lorenz, M. Zahn, Hele–Shaw ferrohydrodynamics for rotating

    and dc axial magnetic fields, Phys Fluids Gallery Fluid Motion 15 (S4

    No. 9) (2003) (see also S4, http://pof.aip.org/pof/gallery/2003toc.jsp).

    [96]&

    S. Rhodes, J. Perez, S. Elborai, S.-H. Lee, M. Zahn, Ferrofluid spiral

    formations and continuous-to-discrete phase transitions under simul-

    taneously applied DC axial and AC in-plane rotating magnetic fields,

    JMMM 289 (2005) 353–355.

    http://pof.aip.org/pof/gallery/2003toc.jsp

  • C. Rinaldi et al. / Current Opinion in Colloid & Interface Science 10 (2005) 141–157156

    [97]&

    S.M. Elborai, D.K. Kim, X. He, S.-H. Lee, S. Rhodes, M. Zahn, Self-

    forming, quasi-two-dimensional magnetic fluid patterns with applied

    in-phase-rotating and DC-axial magnetic fields, J Appl Phys 97

    (2005) 1–3 (10Q303).

    [98] M.I. Shliomis, K.I. Morozov, Negative viscosity of ferrofluid under

    alternating magnetic-field, Phys Fluids 6 (1994) 2855–2861.

    [99] A. Zeuner, R. Richter, I. Rehberg, Experiments on negative and

    positive magnetoviscosity in an alternating magnetic field, Phys Rev

    E 58 (1998) 6287–6293.

    [100] M. Zahn, L.L. Pioch, Magnetizable fluid behavior with effective

    positive, zero or negative dynamic viscosity, Indian J Eng Mat Sci 5

    (1998) 400–410.

    [101] M. Zahn, L.L. Pioch, Ferrofluid flows in ac and traveling wave

    magnetic fields with effective positive, zero or negative dynamic

    viscosity, JMMM 201 (1999) 144–148.

    [102] J.-C. Bacri, R. Perzynski, M.I. Shliomis, G.I. Burde, Negative-

    viscosity effect in a magnetic fluid, Phys Rev Lett 75 (1) (1995)

    2128–2131.

    [103] R.E. Rosensweig, ‘‘Negative viscosity’’ in a magnetic fluid, Science

    271 (5249) (1996) 614–615.

    [104] M. Zahn, D.R. Greer, Ferrohydrodynamic pumping in spatially

    uniform sinusoidally time-varying magnetic fields, JMMM 149 (1)

    (1995) 165–173.

    [105] A. Zeuner, R. Richter, I. Rehberg, On the consistency of the standard

    model for magnetoviscosity in an alternating magnetic field, JMMM

    201 (1999) 191–194.

    [106] X. He, S.M. Elborai, D.K. Kim, S.H. Lee, M. Zahn, Effective

    magnetoviscosity of planar-couette magnetic fluid flow, J Appl Phys

    97 (2005) 1–3 (10Q302).

    [107] C. Rinaldi, Continuum modeling of polarizable systems. PhD thesis

    in the Department of Chemical Engineering, Massachusetts Institute

    of Technology, 2002.

    [108]&

    S. Odenbach, S. Thurm, Magnetoviscous effects in ferrofluids, in: S.

    Odenbach (Ed.), Ferrofluids, magnetically controllable fluids and

    their applications, Springer, Berlin, 2002, pp. 185–201.

    [109] V. Bashtovoi, O. Lavrova, T. Mitkova, V. Polevikov, L. Tobiska,

    Flow and energy dissipation in a magnetic fluid drop around a

    permanent magnet, JMMM 289 (2005) 207–210.

    [110] H. Kikura, J. Matsushita, O. Hirashima, M. Aritomi, I. Nakatani,

    Flow visualization and particle size determination of primary

    clusters in a water-based magnetic fluid, JMMM 289 (2005)

    392–395.

    [111] S. Odenbach, Ferrofluids-magnetically controlled suspensions, Col-

    loids Surf A Physicochem Eng Asp 217 (2003) 171–178.

    [112] M. Kroell, M. Pridoehl, G. Zimmermann, L. Pop, S. Odenbach, A.

    Hartwig, Magnetic and rheological characterization of novel ferro-

    fluids, JMMM 289 (2005) 21–24.

    [113]&

    C. Rinaldi, M. Zahn, Effects of spin viscosity on ferrofluid flow

    profiles in alternating and rotating magnetic fields, Phys Fluids 14 (8)

    (2002) 2847–2870.

    [114] S. Thurm, S. Odenbach, Magnetic separation of ferrofluids, JMMM

    252 (2002) 247–249.

    [115] L.M. Pop, S. Odenbach, A. Wiedenmann, N. Matoussevitch, H.

    Bönnemann, Microstructure and rheology of ferrofluids, JMMM 289

    (2005) 303–306.

    [116] C. Wilhelm, J. Browaeys, A. Ponton, J.-C. Bacri, Rotational

    magnetic particles microrheology: the Maxwellian case, Phys Rev

    E 67 (2003) 1–10 (011504).

    [117] C. Wilhelm, F. Gazeau, J.-C. Bacri, Rotational magnetic endosome

    microrheolgy: viscoelastic architecture inside living cells, Phys Rev

    E 67 (2003) 1–12 (061908).

    [118]&

    Y. Takeda, Velocity profile measurement by ultrasonic Doppler

    method, Exp Therm Fluid Sci 10 (1995) 444–453.

    [119]&

    H. Kikura, Y. Takeda, T. Sawada, Velocity profile measurements of

    magnetic fluid flow using ultrasonic Doppler method, JMMM 201

    (1999) 276–280.

    [120]&

    H. Kikura, M. Aritomi, Y. Takeda, Velocity measurement on Taylor–

    Couette flow of a magnetic fluid with small aspect ratio, JMMM 289

    (2005) 342–345.

    [121]&

    T. Sawada, H. Nishiyama, T. Tabata, Influence of a magnetic field

    on ultrasound propagation in a magnetic fluid, JMMM 252 (2002)

    186–188.

    [122]&

    M. Motozawa, T. Sawada, Influence of magnetic field on

    ultrasonic propagation velocity in magnetic fluids, JMMM 289

    (2005) 66–69.

    [123]&

    R.E. Rosensweig, J. Popplewell, R.J. Johnston, Magnetic fluid

    motion in rotating field, JMMM 85 (1990) 171–180.

    [124]&

    R.E. Rosensweig, S. Elborai, S.-H. Lee, M. Zahn, Theory and

    measurements of ferrofluid meniscus shape in applied uniform

    horizontal and vertical magnetic fields, JMMM 289 (2005)

    192–195.

    [125] E.G. Megalios, N. Kapsalis, J. Paschalidis, A.G. Papathanasiou, A.G.

    Boudouvis, A simple optical device for measuring free surface

    deformations of nontransparent liquids, J Colloid Interface Sci 288

    (2005) 508–512.

    [126] M.S. Amin, S.M. Elborai, S.-H. Lee, X. He, M. Zahn, Surface

    tension measurement techniques of magnetic fluids at an interface

    between different fluids using perpendicular field instability, J Appl

    Phys 97 (2005) 1–3 (10R308).

    [127] V. Sofonea, W.-G. Früh, A. Cristea, Lattice Boltzmann model for the

    simulation of interfacial phenomena in magnetic fluids, JMMM 252

    (2002) 144–146.

    [128] A. Rothert, R. Richter, Experiments on the breakup of a liquid bridge

    of magnetic fluid, JMMM 201 (1999) 324–327.

    [129] S. Sudo, A. Nakagawa, K. Shimada, H. Nishiyama, Shape response

    of functional drops in alternating magnetic fields, JMMM 289 (2005)

    321–324.

    [130] K. Melzner, S. Odenbach, Investigation of the Weissenberg effect

    in ferrofluids under microgravity conditions, JMMM 252 (2002)

    250–252.

    [131] J. Richardi, M.P. Pileni, Nonlinear theory of pattern formation in

    ferrofluid films at high field strengths, Phys Rev E 69 (2004) 1–9

    (016304).

    [132] R.M. Oliveira, J.A. Miranda, Magnetic fluid in a time-dependent gap

    Hele–Shaw cell, JMMM 289 (2005) 360–363.

    [133] J.A. Miranda, R.M. Oliveira, Time dependent gap Hele–Shaw cell

    with a ferrofluid: evidence for an interfacial singularity inhibition by

    a magnetic field, Phys Rev E 69 (2004) 1–5 (066312).

    [134] M. Igonin, A. Cebers, Labyrinthine instability of miscible magnetic

    fluids, JMMM 252 (2002) 293–295.

    [135]&

    D.P. Jackson, Theory, experiment, and simulations of a symmetric

    arrangement of quasi-two-dimensional magnetic fluid drops, JMMM

    289 (2005) 188–191.

    [136] C.-Y. Chen, H.-J. Wu, L. Hsu, Numerical simulations of labyrinthine

    instabilities on a miscible elliptical magnetic droplet, JMMM 289

    (2005) 364–366.

    [137] W. Herreman, P. Molho, S. Neveu, Magnetic field effects on viscous

    fingering of a ferrofluid in a radial Hele–Shaw cell, JMMM 289

    (2005) 356–359.

    [138] I. Drikis, A. Cebers, Viscous fingering in magnetic fluids: numerical

    simulation of radial Hele–Shaw flow, JMMM 201 (1999) 339–342.

    [139] A. Tatulchenkov, A. Cebers, Shapes of a gas bubble rising in the

    vertical Hele–Shaw cell with magnetic liquid, JMMM 289 (2005)

    373–375.

    [140] C.-Y. Chen, H.-J. Wu, Fingering instabilities of a miscible

    magnetic droplet on a rotating Hele–Shaw cell, JMMM 289

    (2005) 339–341.

    [141]&

    G. Pacitto, C. Flament, J.-C. Bacri, M. Widom, Rayleigh–Taylor

    instability with magnetic fluids: experiment and theory, Phys Rev E

    62 (6) (2000) 7941–7948.

    [142] C.-Y. Wen, W.-P. Su, Natural convection of magnetic fluid in a

    rectangular Hele–Shaw cell, JMMM 289 (2005) 299–302.

  • C. Rinaldi et al. / Current Opinion in Colloid & Interface Science 10 (2005) 141–157 157

    [143] T.A. Franklin, Ferrofluid flow phenomena. Master’s Thesis in the

    Department of Electrical Engineering and Computer Science;

    Massachusetts Institute of Technology: Cambridge, MA, 2003.

    [144] K. Zakaria, Nonlinear dynamics of magnetic fluids with a relative

    motion in the presence of an oblique magnetic field, Physica A 327

    (2003) 221–248.

    [145]&

    Y.O. El-Dib, A.Y. Ghaly, Destabilizing effect of time-dependent

    oblique magnetic field on magnetic fluids streaming in porous media,

    J Colloid Interface Sci 269 (2004) 224–239.

    [146] M. Hennenberg, B. Weyssow, S. Slavtchev, V. Alexandrov, T.

    Desaive, Rayleigh–Marangoni–Bénard instability of a ferrofluid

    layer in a vertical magnetic field, JMMM 289 (2005) 268–271.

    [147] V.V. Mekhonoshin, A. Lange, Chain-induced effects in the Faraday

    instability on ferrofluids in a horizontal magnetic field, Phys Fluids

    16 (4) (2004) 925–935.

    [148] T. Mahr, I. Rehberg, Magnetic Faraday instability, Europhys Lett 43

    (1) (1998) 23–28.

    [149] B. Reimann, T. Mahr, R. Richter, I. Rehberg, Standing twin peaks

    due to non-monotonic dispersion of Faraday waves, JMMM 201

    (1999) 303–305.

    [150] A. Lange, Magnetic Soret effect: application of the ferrofluid

    dynamics theory, Phys Rev E 70 (2004) 1–6 ([046308]).

    [151] T. Völker, S. Odenbach, Thermodiffusion in magnetic fluids, JMMM

    289 (2005) 289–291.

    [152] S. Odenbach, T. Völker, Thermal convection in a ferrofluid supported

    by thermodiffusion, JMMM 289 (2005) 122–125.

    [153] T. Völker, E. Blums, S. Odenbach, Thermodiffusion in magnetic

    fluids, JMMM 252 (2002) 218–220.

    [154] A.J. Silva, M. Gonçalves, S.M. Shibli, Thermodiffusion study in

    ferrofluids through collinear mirage effect, JMMM 289 (2005)

    295–298.

    [155] E. Blums, New transport properties of ferrocolloids: magnetic Soret

    effect and thermomagnetoosmosis, JMMM 289 (2005) 246–249.

    [156] A. Ryskin, H. Pleiner, Influence of a magnetic field on the Soret-

    effect-dominated thermal convection in ferrofluids, Phys Rev E 69

    (2004) 1–10 (046301).

    [157]&

    E. Blums, Heat and mass transfer phenomena, in: S. Odenbach (Ed.),

    Ferrofluids, Magnetically Controllable Fluids and Their Applica-

    tions, Springer, Berlin, 2002, pp. 124–142.

    [158] S.I. Alves, A. Bourdon, A.M. Figueiredo Neto, Investigation of the

    Soret coefficient in magnetic fluids using the Z-scan technique,

    JMMM 289 (2005) 285–288.

    [159] J. Lenglet, A. Bourdon, J.-C. Bacri, G. Demouchy, Thermodiffu-

    sion in magnetic colloids evidenced and studied by forced

    Rayleigh scattering experiments, Phys Rev E 65 (2002) 1–14

    (031408).

    [160] M.I. Shliomis, B.I. Smorodin, Convective instability of magnetized

    ferrofluids, JMMM 252 (2002) 197–202.

    [161] W. Luo, T. Du, J. Huang, Novel convective instabilities in a magnetic

    fluid, Phys Rev Lett 82 (20) (1999) 4134–4137.

    [162] M.I. Shliomis, Comment on ‘‘novel convective instabilities in a

    magnetic fluid’’, Phys Rev Lett 87 (5) (2001) 1, (059801).

    [163] V.A. Naletova, A.S. Kvitantsev, Thermomagnetic force acting on

    a spheroidal body in a magnetic fluid, JMMM 289 (2005)

    250–252.

    Magnetic fluid rheology and flowsIntroduction to magnetic fluidsFerrofluid compositionApplications

    Governing ferrohydrodynamic equationsMagnetizationLangevin magnetization equilibriumMagnetization relaxationMagnetization relaxation time constants

    Complex magnetic susceptibilityHeatingConservation of linear and angular momentum equations

    Magnetic forces and torquesMagnetic force densityTorque-driven flows

    FerrohydrodynamicsMagnetoviscosityUltrasound velocity profile measurementsInterfacial phenomenaHele-Shaw cell flowsFlow instabilitiesJet and sheet flowsKelvin-Helmholtz instabilityOther flow instabilities

    Thermodiffusion (Soret effect)

    ConclusionsAcknowledgmentsReferences and recommended readings,•


Recommended